Sample records for observed brightness temperature

  1. Intrinsic Brightness Temperatures of AGN Jets

    NASA Astrophysics Data System (ADS)

    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.

    2006-05-01

    We present a new method for studying the intrinsic brightness temperatures of the parsec-scale jet cores of active galactic nuclei (AGNs). Our method uses observed superluminal motions and observed brightness temperatures for a large sample of AGNs to constrain the characteristic intrinsic brightness temperature of the sample as a whole. To study changes in intrinsic brightness temperature, we assume that the Doppler factors of individual jets are constant in time, as justified by their relatively small changes in observed flux density. We find that in their median-low brightness temperature state, the sources in our sample have a narrow range of intrinsic brightness 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 brightness state, we find that sources in our sample have a characteristic intrinsic brightness 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 brightness 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 observed brightness temperature, the assumption of equipartition, or both may lead to large scatter or systematic errors in the derived values.

  2. 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.

  3. Landcover Based Optimal Deconvolution of PALS L-band Microwave Brightness Temperature

    NASA Technical Reports Server (NTRS)

    Limaye, Ashutosh S.; Crosson, William L.; Laymon, Charles A.; Njoku, Eni G.

    2004-01-01

    An optimal de-convolution (ODC) technique has been developed to estimate microwave brightness temperatures of agricultural fields using microwave radiometer observations. 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 brightness temperatures of corn and soybeans were observed to be significantly different because of large differences in vegetation biomass. PALS observations have significant over-sampling; observations were made about 100 m apart and the sensor footprint extends to about 400 m. Conventionally, observations of this type are averaged to produce smooth spatial data fields of brightness temperatures. However, the conventional approach is in contrast to reality in which the brightness temperatures are in fact strongly dependent on landcover, which is characterized by sharp boundaries. In this study, we mathematically de-convolve the observations into brightness temperature at the field scale (500-800m) using the sensor antenna response function. The result is more accurate spatial representation of field-scale brightness temperatures, which may in turn lead to more accurate soil moisture retrieval.

  4. [Monitoring of brightness temperature fluctuation of water in SHF range].

    PubMed

    Ivanov, Yu D; Kozlov, A F; Galiullin, R A; Tatu, V Yu; Vesnin, S G; Ziborov, V S; Ivanova, N D; Pleshakova, T O

    2017-01-01

    The purpose of the research consisted in detection of fluctuation of brightness 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 brightness 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 brightness temperature measurements were carried out using radiometer. Results: Fluctuation with maximum of brightness 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 brightness 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 observed. At that, the environment temperature remained constant. Conclusion: The significant increasing in brightness 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 observed. At the same time, thermodynamic temperature virtually does not change. The observed effects can be used in the development of the systems for diadnostics of pathologies in human and analytical system.

  5. Microwave Brightness Temperatures of Tilted Convective Systems

    NASA Technical Reports Server (NTRS)

    Hong, Ye; Haferman, Jeffrey L.; Olson, William S.; Kummerow, Christian D.

    1998-01-01

    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 observe 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 brightness 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 brightness 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 brightness temperature. The radiative transfer results indicate that brightness temperature varies when the viewing angle changes because of the different optical depth. The tilt increases the displacements between high 19 GHz brightness temperature (Tb(sub 19)) due to liquid emission from lower level of cloud and the low 85 GHz brightness temperature (Tb(sub 85)) due to ice scattering from upper level of cloud. As the resolution degrades, the difference of brightness temperature due to the change of viewing angle decreases dramatically. The dislocation between Tb(sub 19) and Tb(sub 85), however, remains prominent.

  6. 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.

  7. Extremely Low Passive Microwave Brightness Temperatures Due to Thunderstorms

    NASA Technical Reports Server (NTRS)

    Cecil, Daniel J.

    2015-01-01

    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 observed over northern Argentina with 85 GHz brightness temperatures below 50 K and 37 GHz brightness temperatures below 70 K (Zipser et al. 2006). These values are well below what had previously been observed from satellite sensors with lower resolution. The 37 GHz brightness 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 brightness temperatures below 150 K would normally be interpreted as the result of a low-emissivity water surface (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 brightness temperatures resulting from thunderstorms as seen by TRMM have been in Argentina in November and December. For

  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. The Influence of Microphysical Cloud Parameterization on Microwave Brightness Temperatures

    NASA Technical Reports Server (NTRS)

    Skofronick-Jackson, Gail M.; Gasiewski, Albin J.; Wang, James R.; Zukor, Dorothy J. (Technical Monitor)

    2000-01-01

    The microphysical parameterization of clouds and rain-cells plays a central role in atmospheric forward radiative transfer models used in calculating passive microwave brightness 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 brightness temperatures with respect to the microphysical cloud parameterization. Cloud parameterizations for wideband (6-410 GHz observations of baseline brightness 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 brightness 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 brightness 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) brightness temperatures shows that in general all but, two parameterizations produce calculated T(sub B)'s that fall within the observed clear-air minima and maxima. The exceptions are for parameterizations that, enhance the scattering characteristics of frozen hydrometeors.

  10. Synthesizing SMOS Zero-Baselines with Aquarius Brightness Temperature Simulator

    NASA Technical Reports Server (NTRS)

    Colliander, A.; Dinnat, E.; Le Vine, D.; Kainulainen, J.

    2012-01-01

    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 Surface 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 brightness temperature observations 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 brightness temperature level is the measurements performed with the so-called zero-baselines [3]; SMOS employs an interferometric measurement technique which forms a brightness temperature image from several baselines constructed by combination of multiple receivers in an array; zero-length baseline defines the overall brightness temperature level. The basis of the Aquarius brightness temperature level is resolved from the brightness 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.

  11. 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.

  12. 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

  13. 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.

  14. Arctic sea ice signatures: L-band brightness temperature sensitivity comparison using two radiation transfer models

    NASA Astrophysics Data System (ADS)

    Richter, Friedrich; Drusch, Matthias; Kaleschke, Lars; Maaß, Nina; Tian-Kunze, Xiangshan; Mecklenburg, Susanne

    2018-03-01

    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 brightness 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 brightness 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 brightness temperatures. We use two different radiative transfer models to generate top of atmosphere brightness 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 brightness temperatures over sea ice. The simulated brightness 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 observe the largest differences of more than 20 K over sea ice between simulated and observed brightness temperatures. We conclude that the assimilation of SMOS brightness 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 brightness temperatures should be used simultaneously.

  15. Microwave brightness temperature features of lunar craters: observation from Chang'E-1 mission

    NASA Astrophysics Data System (ADS)

    Hu, Guo-Ping; Chen, Ke; Guo, Wei; Li, Qing-Xia; Su, Hong-Yan

    2013-01-01

    Topographic features of lunar craters have been found from the brightness temperature (TB) observed 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 observation 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.

  16. 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.

  17. 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.

  18. 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.

  19. Calculation of gyrosynchrotron radiation brightness temperature for outer bright loop of ICME

    NASA Astrophysics Data System (ADS)

    Sun, Weiying; Wu, Ji; Wang, C. B.; Wang, S.

    :Solar polar orbit radio telescope (SPORT) is proposed to detect the high density plasma clouds of outer bright 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 bright loop of ICMEs. Moreover, we calculate and analyze the emission brightness 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.

  20. L band brightness temperature observations over a corn canopy during the entire growth cycle

    USDA-ARS?s Scientific Manuscript database

    During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided brightness temperature (T¬B) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characte...

  1. Effect of Different Tree canopies on the Brightness Temperature of Snowpack

    NASA Astrophysics Data System (ADS)

    Mousavi, S.; De Roo, R. D.; Brucker, L.

    2017-12-01

    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 observations 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 observe continuously the snowpack surrounding the truck. Observations 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 brightness temperature of snow is discussed. It is shown that Engelmann Spruce trees increases the brightness temperature of the snowpack more than Aspen trees do. Moreover, the elevation angular dependence of the measured brightness 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 brightness temperature change as snow depth evolved. Also, our study takes advantage of the snowpit measurements that were collected near the radiometers' field of view.

  2. Brightness temperature - obtaining the physical properties of a non-equipartition plasma

    NASA Astrophysics Data System (ADS)

    Nokhrina, E. E.

    2017-06-01

    The limit on the intrinsic brightness 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 observed cores of active galactic nuclei (AGNs) sustained the limit lower than the `Compton catastrophe'. Recent observations with unprecedented high resolution by the RadioAstron have revealed systematic exceed in the observed brightness 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 brightness 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.

  3. Assimilation of SMOS brightness temperatures in the ECMWF EKF for the analysis of soil moisture

    NASA Astrophysics Data System (ADS)

    Munoz-Sabater, Joaquin

    2012-07-01

    Since November 2nd 2009, the European Centre for Medium-Range Weather Forecasts (ECMWF) has being monitoring, in Near Real Time (NRT), L-band brightness 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 observed brightness 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 brightness 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 observed brightness 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 brightness temperatures which match the ECMWF T511 (~40 km) reduced Gaussian Grid. Secondly, SMOS observational 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 observations to ensure an unbiased dataset ready for assimilation in the ECMWF surface analysis system. The current ECMWF operational soil moisture analysis

  4. 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.

  5. 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.

  6. Relationship of magnetic field strength and brightness of fine-structure elements in the solar temperature minimum region

    NASA Technical Reports Server (NTRS)

    Cook, J. W.; Ewing, J. A.

    1990-01-01

    A quantitative relationship was determined between magnetic field strength (or magnetic flux) from photospheric magnetograph observations and the brightness temperature of solar fine-structure elements observed at 1600 A, where the predominant flux source is continuum emission from the solar temperature minimum region. A Kitt Peak magnetogram and spectroheliograph observations 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 brightness temperature in the quiet sun at 1600 A, and absolute value of magnetic field strength in the same area were determined from these observations. 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 brightness 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.

  7. Microwave brightness temperature of a windblown sea

    NASA Technical Reports Server (NTRS)

    Hall, F. G.

    1972-01-01

    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 surface or atmosphere. It is shown that a variation of microwave brightness temperature with zenith angle is necessary to produce the wind sensitivity of the horizontally polarized brightness temperature; the variation of sky temperature with frequency is sufficient to produce a frequency dependent wind sensitivity.

  8. 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.

  9. Characterization and Correction of Aquarius Long Term Calibration Drift Using On-Earth Brightness Temperature Refernces

    NASA Technical Reports Server (NTRS)

    Brown, Shannon; Misra, Sidharth

    2013-01-01

    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 surface salinity. One of the main mission objectives is to provide monthly global salinity maps for climate studies of ocean circulation, surface 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 brightness 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 brightness temperature stability at the warmer end of the brightness temperature dynamic range. Soon after launch, time variable drifts were observed in the Aquarius data compared to in-situ data from ARGO and models for the ocean surface 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 brightness temperature calibration. If arising from the brightness 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 brightness temperature calibration using on-Earth reference targets that were independent of the ocean model.

  10. Understanding Climate Trends Using IR Brightness Temperature Spectra from AIRS, IASI and CrIS

    NASA Astrophysics Data System (ADS)

    Deslover, D. H.; Nikolla, E.; Knuteson, R. O.; Revercomb, H. E.; Tobin, D. C.

    2016-12-01

    NASA's Atmospheric Infrared Sounder (AIRS) provides a data record that extends from its 2002 launch to the present. The Infrared Atmospheric Sounding Interferometer (IASI) onboard Metop- (A launched in 2006, B in 2012), as well as the Joint Polar Satellite System (JPSS) Cross-track Infrared Sounder (CrIS) launched in 2011, complement this data record. Future infrared sounders with similar capabilities will augment these measurements into the near future. We have created a global data set from these infrared measurements, using the nadir-most observations for each of the aforementioned instruments. We can filter the data based upon spatial, diurnal and seasonal properties to discern trends for a given spectral channel and, therefore, a specific atmospheric layer. Subtle differences between spectral sampling among the three instruments can lead significant differences in the resultant probability distribution functions for similar spectral channels. We take advantage of the higher (0.25 cm-1) IASI spectral resolution to subsample the IASI spectra onto AIRS and CrIS spectral grids to better compare AIRS/IASI and CrIS/IASI trends in the brightness temperature anomalies. To better understand the dependance of trace gases on the measured brightness temperature spectral time-series, a companion study has utilized coincident vertical profiles of stratospheric carbon dioxide, water vapor and ozone concentration are used to infer a correlation with the CrIS brightness temperatures. The goal was to investigate the role of ozone heating and carbon dioxide cooling on the observed brightness temperature spectra. Results from that study will be presented alongside the climate trend analysis.

  11. S193 radiometer brightness temperature precision/accuracy for SL2 and SL3

    NASA Technical Reports Server (NTRS)

    Pounds, D. J.; Krishen, K.

    1975-01-01

    The precision and accuracy with which the S193 radiometer measured the brightness temperature of ground scenes is investigated. Estimates were derived from data collected during Skylab missions. Homogeneous ground sites were selected and S193 radiometer brightness temperature data analyzed. The precision was expressed as the standard deviation of the radiometer acquired brightness temperature. Precision was determined to be 2.40 K or better depending on mode and target temperature.

  12. 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.

  13. Extreme Brightness Temperatures and Refractive Substructure in 3C273 with RadioAstron

    NASA Astrophysics Data System (ADS)

    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.

    2016-03-01

    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 brightness 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 brightness 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 brightness 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 brightness temperatures, in excess of {10}13 {{K}}. We also demonstrate that for a source having a Gaussian surface brightness profile, a single long-baseline estimate of refractive substructure determines an absolute minimum brightness temperature, if the scattering properties along a given line of sight are known, and that this minimum accurately approximates the apparent brightness temperature over a wide range of total flux densities.

  14. 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.

  15. IRTM brightness temperature maps of the Martian south polar region during the polar night: The cold spots don't move

    NASA Technical Reports Server (NTRS)

    Paige, D. A.; Crisp, D.; Santee, M. L.; Richardson, M. I.

    1993-01-01

    A series of infrared thermal mapper (IRTM) south polar brightness temperature maps obtained by Viking Orbiter 2 during a 35-day period during the southern fall season in 1978 was examined. The maps show a number of phenomena that have been identified in previous studies, including day to day brightness temperature variations in individual low temperature regions and the tendency for IRTM 11-micron channel brightness temperatures to also decrease in regions where low 20-micron channel brightness temperatures are observed. The maps also show new phenomena, the most striking of which is a clear tendency for the low brightness temperature regions to occur at fixed geographic regions. During this season, the coldest low brightness temperatures appear to be concentrated in distinct regions, with spatial scales ranging from 50 to 300 km. There are approximately a dozen of these concentrations, with the largest centered near the location of the south residual polar cap. Other concentrations are located at Cavi Angusti and close to the craters Main, South, Lau, and Dana. Broader, less intense regions appear to be well correlated with the boundaries of the south polar layered deposits and the Mountains of Mitchell. No evidence for horizontal motion of any of these regions has been detected.

  16. Simulation of the brightness temperatures observed by the visible infrared imaging radiometer suite instrument

    NASA Astrophysics Data System (ADS)

    Evrard, Rebecca L.; Ding, Yifeng

    2018-01-01

    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 brightness 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 observations 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.

  17. Soil Moisture Active/Passive (SMAP) Forward Brightness Temperature Simulator

    NASA Technical Reports Server (NTRS)

    Peng, Jinzheng; Peipmeier, Jeffrey; Kim, Edward

    2012-01-01

    The SMAP is one of four first-tier missions recommended by the US National Research Council's Committee on Earth Science and Applications from Space (Earth Science and Applications from Space: National Imperatives for the Next Decade and Beyond, Space Studies Board, National Academies Press, 2007) [1]. It is to measure the global soil moisture and freeze/thaw from space. One of the spaceborne instruments is an L-band radiometer with a shared single feedhorn and parabolic mesh reflector. While the radiometer measures the emission over a footprint of interest, unwanted emissions are also received by the antenna through the antenna sidelobes from the cosmic background and other error sources such as the Sun, the Moon and the galaxy. Their effects need to be considered accurately, and the analysis of the overall performance of the radiometer requires end-to-end performance simulation from Earth emission to antenna brightness temperature, such as the global simulation of L-band brightness temperature simulation over land and sea [2]. To assist with the SMAP radiometer level 1B algorithm development, the SMAP forward brightness temperature simulator is developed by adapting the Aquarius simulator [2] with necessary modifications. This poster presents the current status of the SMAP forward brightness simulator s development including incorporating the land microwave emission model and its input datasets, and a simplified atmospheric radiative transfer model. The latest simulation results are also presented to demonstrate the ability of supporting the SMAP L1B algorithm development.

  18. Effects of evening bright light exposure on melatonin, body temperature and sleep.

    PubMed

    Bunnell; Treiber; Phillips; Berger

    1992-03-01

    Five male subjects were exposed to a single 2-h period of bright (2500 lux) or dim (<100 lux) light prior to sleep on two consecutive nights. The two conditions were repeated the following week in opposite order. Bright light significantly suppressed salivary melatonin and raised rectal temperature 0.3 degrees C (which remained elevated during the first 1.5 h of sleep), without affecting tympanic temperature. Bright light also increased REM latency, NREM period length, EEG spectral power in low frequency, 0.75-8 Hz and sigma, 12-14 Hz (sleep spindle) bandwidths during the first hour of sleep, and power of all frequency bands (0.5-32 Hz) within the first NREMP. Potentiation of EEG slow wave activity (0.5-4.0 Hz) by bright light persisted through the end of the second NREMP. The enhanced low-frequency power and delayed REM sleep after bright light exposure could represent a circadian phase-shift and/or the effect of an elevated rectal temperature, possibly mediated by the suppression of melatonin.

  19. Exposure to bright light for several hours during the daytime lowers tympanic temperature

    NASA Astrophysics Data System (ADS)

    Aizawa, Seika; Tokura, H.

    The present study investigates the effect on thympanic temperature of exposure to different light intensities for several hours during the daytime. Nine healthy young adult volunteers (two male, seven female) were exposed to bright light of 4000 lx or dim light of 100 lx during the daytime from 0930 to 1800 hours; the light condition was then kept at 100 lx for a further hour. Tympanic temperature was measured continuously at a neutral condition (28° C, 60% relative humidity) from 1000 to 1800 hours. Urinary samples were collected from 1100 to 1900 hours every 2 h, and melatonin excretion rate was measured by enzyme immunoassay. Of nine subjects, six showed clearly lower tympanic temperatures in the bright compared with the dim condition from 1400 to 1800 hours. Average tympanic temperatures were significantly lower in the bright than in the dim condition from 1645 to 1800 hours. Melatonin excretion rate tended to be higher in the bright than in the dim condition. It was concluded that exposure to bright light of 4000 lx during the daytime for several hours could reduce tympanic temperature, compared with that measured in dim light of 100 lx.

  20. Exposure to bright light for several hours during the daytime lowers tympanic temperature.

    PubMed

    Aizawa, S; Tokura, H

    1997-11-01

    The present study investigates the effect on thympanic temperature of exposure to different light intensities for several hours during the daytime. Nine healthy young adult volunteers (two male, seven female) were exposed to bright light of 4000 lx or dim light of 100 lx during the daytime from 0930 to 1800 hours; the light condition was then kept at 100 lx for a further hour. Tympanic temperature was measured continuously at a neutral condition (28 degrees C, 60% relative humidity) from 1000 to 1800 hours. Urinary samples were collected from 1100 to 1900 hours every 2 h, and melatonin excretion rate was measured by enzyme immunoassay. Of nine subjects, six showed clearly lower tympanic temperatures in the bright compared with the dim condition from 1400 to 1800 hours. Average tympanic temperatures were significantly lower in the bright than in the dim condition from 1645 to 1800 hours. Melatonin excretion rate tended to be higher in the bright than in the dim condition. It was concluded that exposure to bright light of 4000 lx during the daytime for several hours could reduce tympanic temperature, compared with that measured in dim light of 100 lx.

  1. 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.

  2. 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.

  3. 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

  4. 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

  5. L Band Brightness Temperature Observations over a Corn Canopy during the Entire Growth Cycle

    PubMed Central

    Joseph, Alicia T.; van der Velde, Rogier; O’Neill, Peggy E.; Choudhury, Bhaskar J.; Lang, Roger H.; Kim, Edward J.; Gish, Timothy

    2010-01-01

    During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided brightness temperature (TB) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characterization of land surface variables including soil moisture, soil temperature, vegetation biomass, and surface 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 surface emissions are investigated by means of a semi-empirical radiative transfer model. Additionally, the impact of roughness on the surface 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

  6. The GPM Common Calibrated Brightness Temperature Product

    NASA Technical Reports Server (NTRS)

    Stout, John; Berg, Wesley; Huffman, George; Kummerow, Chris; Stocker, Erich

    2005-01-01

    The Global Precipitation Measurement (GPM) project will provide a core satellite carrying the GPM Microwave Imager (GMI) and will use microwave observations from a constellation of other satellites. Each partner with a satellite in the constellation will have a calibration that meets their own requirements and will decide on the format to archive their brightness temperature (Tb) record in GPM. However, GPM multi-sensor precipitation algorithms need to input intercalibrated Tb's in order to avoid differences among sensors introducing artifacts into the longer term climate record of precipitation. The GPM Common Calibrated Brightness Temperature Product is intended to address this problem by providing intercalibrated Tb data, called "Tc" data, where the "c" stands for common. The precipitation algorithms require a Tc file format that is both generic and flexible enough to accommodate the different passive microwave instruments. The format will provide detailed information on the processing history in order to allow future researchers to have a record of what was done. The format will be simple, including the main items of scan time, latitude, longitude, and Tc. It will also provide spacecraft orientation, spacecraft location, orbit, and instrument scan type (cross-track or conical). Another simplification is to store data in real numbers, avoiding the ambiguity of scaled data. Finally, units and descriptions will be provided in the product. The format is built on the concept of a swath, which is a series of scans that have common geolocation and common scan geometry. Scan geometry includes pixels per scan, sensor orientation, scan type, and incidence angles. The Tc algorithm and data format are being tested using the pre-GPM Precipitation Processing System (PPS) software to generate formats and 1/0 routines. In the test, data from SSM/I, TMI, AMSR-E, and WindSat are being processed and written as Tc products.

  7. The correlation of Skylab L-band brightness temperatures with antecedent precipitation

    NASA Technical Reports Server (NTRS)

    Mcfarland, M. J.

    1975-01-01

    The S194 L-band radiometer flown on the Skylab mission measured terrestrial radiation at the microwave wavelength of 21.4 cm. The terrain emissivity at this wavelength is strongly dependent on the soil moisture content, which can be inferred from antecedent precipitation. For the Skylab data acquisition pass from the Oklahoma panhandle to southeastern Texas on 11 June 1973, the S194 brightness temperatures are highly correlated with antecedent precipitation from the preceding eleven day period, but very little correlation was apparent for the preceding five day period. The correlation coefficient between the averaged antecedent precipitation index values and the corresponding S194 brightness temperatures between 230 K and 270 K, the region of apparent response to soil moisture in the data, was -0.97. The equation of the linear least squares line fitted to the data was: API (cm) = 31.99 -0.114 TB (K), where API is the antecedent precipitation index and TB is the S194 brightness temperature.

  8. 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

  9. Quantifying spatial and temporal variabilities of microwave brightness temperature over the U.S. Southern Great Plains

    NASA Technical Reports Server (NTRS)

    Choudhury, B. J.; Owe, M.; Ormsby, J. P.; Chang, A. T. C.; Wang, J. R.; Goward, S. N.; Golus, R. E.

    1987-01-01

    Spatial and temporal variabilities of microwave brightness temperature over the U.S. Southern Great Plains are quantified in terms of vegetation and soil wetness. The brightness temperatures (TB) are the daytime observations from April to October for five years (1979 to 1983) obtained by the Nimbus-7 Scanning Multichannel Microwave Radiometer at 6.6 GHz frequency, horizontal polarization. The spatial and temporal variabilities of vegetation are assessed using visible and near-infrared observations by the NOAA-7 Advanced Very High Resolution Radiometer (AVHRR), while an Antecedent Precipitation Index (API) model is used for soil wetness. The API model was able to account for more than 50 percent of the observed variability in TB, although linear correlations between TB and API were generally significant at the 1 percent level. The slope of the linear regression between TB and API is found to correlate linearly with an index for vegetation density derived from AVHRR data.

  10. 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.

  11. 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.

  12. Predicting Near Real-Time Inundation Occurrence from Complimentary Satellite Microwave Brightness Temperature Observations

    NASA Astrophysics Data System (ADS)

    Fisher, C. K.; Pan, M.; Wood, E. F.

    2017-12-01

    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 observations 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 brightness temperature observations (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 observations 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

  13. L Band Brightness Temperature Observations Over a Corn Canopy During the Entire Growth Cycle

    NASA Technical Reports Server (NTRS)

    Joseph, Alicia T.; O'Neill, Peggy E.; Choudhury, Bhaskar J.; vanderVelde, Rogier; Lang, Roger H.; Gish, Timothy

    2011-01-01

    During a field campaign covering the 2002 corn growing season, a dual polarized tower mounted L-band (1.4 GHz) radiometer (LRAD) provided brightness temperature (T(sub B)) measurements at preset intervals, incidence and azimuth angles. These radiometer measurements were supported by an extensive characterization of land surface variables including soil moisture, soil temperature, vegetation biomass, and surface 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 surface emissions are investigated by means of a semi-empirical radiative transfer model. Additionally, the impact of roughness on the surface 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.

  14. 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.

  15. SMOS brightness temperature assimilation into the Community Land Model

    NASA Astrophysics Data System (ADS)

    Rains, Dominik; Han, Xujun; Lievens, Hans; Montzka, Carsten; Verhoest, Niko E. C.

    2017-11-01

    SMOS (Soil Moisture and Ocean Salinity mission) brightness temperatures at a single incident angle are assimilated into the Community Land Model (CLM) across Australia to improve soil moisture simulations. Therefore, the data assimilation system DasPy is coupled to the local ensemble transform Kalman filter (LETKF) as well as to the Community Microwave Emission Model (CMEM). Brightness temperature climatologies are precomputed to enable the assimilation of brightness temperature anomalies, making use of 6 years of SMOS data (2010-2015). Mean correlation R with in situ measurements increases moderately from 0.61 to 0.68 (11 %) for upper soil layers if the root zone is included in the updates. A reduced improvement of 5 % is achieved if the assimilation is restricted to the upper soil layers. Root-zone simulations improve by 7 % when updating both the top layers and root zone, and by 4 % when only updating the top layers. Mean increments and increment standard deviations are compared for the experiments. The long-term assimilation impact is analysed by looking at a set of quantiles computed for soil moisture at each grid cell. Within hydrological monitoring systems, extreme dry or wet conditions are often defined via their relative occurrence, adding great importance to assimilation-induced quantile changes. Although still being limited now, longer L-band radiometer time series will become available and make model output improved by assimilating such data that are more usable for extreme event statistics.

  16. Lunar brightness temperature from Microwave Radiometers data of Chang'E-1 and Chang'E-2

    NASA Astrophysics Data System (ADS)

    Feng, J.-Q.; Su, Y.; Zheng, L.; Liu, J.-J.

    2011-10-01

    Both of the Chinese lunar orbiter, Chang'E-1 and Chang'E-2 carried Microwave Radiometers (MRM) to obtain the brightness temperature of the Moon. Based on the different characteristics of these two MRMs, modified algorithms of brightness temperature and specific ground calibration parameters were proposed, and the corresponding lunar global brightness temperature maps were made here. In order to analyze the data distributions of these maps, normalization method was applied on the data series. The second channel data with large deviations were rectified, and the reasons of deviations were analyzed in the end.

  17. Long-term observations minus background monitoring of ground-based brightness temperatures from a microwave radiometer network

    NASA Astrophysics Data System (ADS)

    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

    2017-10-01

    Ground-based microwave radiometers (MWRs) offer the capability to provide continuous, high-temporal-resolution observations 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 observations 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 observations, 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 observations and their background counterparts (O-B). Monitoring of O-B statistics is crucial to detect and remove systematic errors coming from the measurements, the observation operator, and/or the NWP model. This work illustrates a 1-year O-B analysis for MWR observations in clear-sky conditions for an European-wide network of six MWRs. Observations include MWR brightness 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 observation 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 observations show a similar trend, though values increase with increasing air mass. O-B statistics for temperature channels show different

  18. 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.

  19. Coronal bright points at 6cm wavelength

    NASA Technical Reports Server (NTRS)

    Fu, Qijun; Kundu, M. R.; Schmahl, E. J.

    1988-01-01

    Results are presented from observations of bright points at a wavelength of 6-cm using the VLA with a spatial resolution of 1.2 arcsec. During two hours of observations, 44 sources were detected with brightness 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 bright points.

  20. Absolute brightness temperature measurements at 3.5-mm wavelength. [of sun, Venus, Jupiter and Saturn

    NASA Technical Reports Server (NTRS)

    Ulich, B. L.; Rhodes, P. J.; Davis, J. H.; Hollis, J. M.

    1980-01-01

    Careful observations have been made at 86.1 GHz to derive the absolute brightness temperatures of the sun (7914 + or - 192 K), Venus (357.5 + or - 13.1 K), Jupiter (179.4 + or - 4.7 K), and Saturn (153.4 + or - 4.8 K) with a standard error of about three percent. This is a significant improvement in accuracy over previous results at millimeter wavelengths. A stable transmitter and novel superheterodyne receiver were constructed and used to determine the effective collecting area of the Millimeter Wave Observatory (MWO) 4.9-m antenna relative to a previously calibrated standard gain horn. The thermal scale was set by calibrating the radiometer with carefully constructed and tested hot and cold loads. The brightness temperatures may be used to establish an absolute calibration scale and to determine the antenna aperture and beam efficiencies of other radio telescopes at 3.5-mm wavelength.

  1. Australia 31-GHz brightness temperature exceedance statistics

    NASA Technical Reports Server (NTRS)

    Gary, B. L.

    1988-01-01

    Water vapor radiometer measurements were made at DSS 43 during an 18 month period. Brightness temperatures at 31 GHz were subjected to a statistical analysis which included correction for the effects of occasional water on the radiometer radome. An exceedance plot was constructed, and the 1 percent exceedance statistics occurs at 120 K. The 5 percent exceedance statistics occurs at 70 K, compared with 75 K in Spain. These values are valid for all of the three month groupings that were studied.

  2. The Atacama Cosmology Telescope: Beam Measurements and the Microwave Brightness Temperatures of Uranus and Saturn

    NASA Technical Reports Server (NTRS)

    Hasselfield, Matthew; Moodley, Kavilan; Bond, J. Richard; Das, Sudeep; Devlin, Mark J.; Dunkley, Joanna; Dunner, Rolando; Fowler, Joseph W.; Gallardo, Patricio; Gralla, Megan B.; hide

    2013-01-01

    We describe the measurement of the beam profiles and window functions for the Atacama Cosmology Telescope (ACT), which operated from 2007 to 2010 with kilopixel bolometer arrays centered at 148, 218, and 277 GHz. Maps of Saturn are used to measure the beam shape in each array and for each season of observations. Radial profiles are transformed to Fourier space in a way that preserves the spatial correlations in the beam uncertainty to derive window functions relevant for angular power spectrum analysis. Several corrections are applied to the resulting beam transforms, including an empirical correction measured from the final cosmic microwave background (CMB) survey maps to account for the effects of mild pointing variation and alignment errors. Observations of Uranus made regularly throughout each observing season are used to measure the effects of atmospheric opacity and to monitor deviations in telescope focus over the season. Using the WMAP-based calibration of the ACT maps to the CMB blackbody, we obtain precise measurements of the brightness temperatures of the Uranus and Saturn disks at effective frequencies of 149 and 219 GHz. For Uranus we obtain thermodynamic brightness temperatures T(149/U) = 106.7 +/- 2.2 K and T(219/U) = 100.1 +/- 3.1 K. For Saturn, we model the effects of the ring opacity and emission using a simple model and obtain resulting (unobscured) disk temperatures of T(149/S) = 137.3 +/- 3.2 K and T(219/S) = 137.3 +/- 4.7 K.

  3. The prediction of tropopause height from clusters of brightness temperatures and its application in the stratified regression temperature retrievals using microwave and infrared satellite measurements

    NASA Technical Reports Server (NTRS)

    Munteanu, M. J.; Piraino, P.; Jakubowicz, O.

    1984-01-01

    A total of 1575 radiosondes and the corresponding simulated brightness temperatures were used in an effort to derive a temperature retrieval based on the clusters of brightness temperatures. The 8 simulated channels, namely, 3 MSU and 5 IR of the TIROS-N satellite are used by the GLAS temperature retrieval method. The 3 MSU and 5 IR brightness temperatures were clustered into 17 cluster groups and a regression for the prediction of the tropopause height in mb was generated. The overall r.m.s. for the tropopause prediction is excellent, namely, around 16 mb for the summer and 23 mb for the winter. The correct cluster of brightness temperatures can be identified 98% of the time by the method of discriminatory classification if it is approximately a normal distribution or, in general, by the method of the nearest neighbor.

  4. Preliminary Evaluation of Influence of Aerosols on the Simulation of Brightness Temperature in the NASA's Goddard Earth Observing System Atmospheric Data Assimilation System

    NASA Technical Reports Server (NTRS)

    Kim, Jong; Akella, Santha; da Silva, Arlindo M.; Todling, Ricardo; McCarty, William

    2018-01-01

    This document reports on preliminary results obtained when studying the impact of aerosols on the calculation of brightness temperature (BT) for satellite infrared (IR) instruments that are currently assimilated in a 3DVAR configuration of Goddard Earth Observing 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 observation 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 observations over regions with significant amounts of dust. The cooling effect obtained with the aerosol-affected BT leads to a slight warming of the analyzed surface 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.

  5. 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.

  6. The high brightness temperature of B0529+483 revealed by RadioAstron and implications for interstellar scattering

    NASA Astrophysics Data System (ADS)

    Pilipenko, S. V.; Kovalev, Y. Y.; Andrianov, A. S.; Bach, U.; Buttaccio, S.; Cassaro, P.; Cimò, G.; Edwards, P. G.; Gawroński, M. P.; Gurvits, L. I.; Hovatta, T.; Jauncey, D. L.; Johnson, M. D.; Kovalev, Yu A.; Kutkin, A. M.; Lisakov, M. M.; Melnikov, A. E.; Orlati, A.; Rudnitskiy, A. G.; Sokolovsky, K. V.; Stanghellini, C.; de Vicente, P.; Voitsik, P. A.; Wolak, P.; Zhekanis, G. V.

    2018-03-01

    The high brightness temperatures, Tb ≳ 1013 K, detected in several active galactic nuclei by RadioAstron space VLBI observations challenge theoretical limits. Refractive scattering by the interstellar medium may affect such measurements. We quantify the scattering properties and the sub-mas scale source parameters for the quasar B0529+483. Using RadioAstron correlated flux density measurements at 1.7, 4.8, and 22 GHz on projected baselines up to 240 000 km we find two characteristic angular scales in the quasar core, about 100 and 10 μas. Some indications of scattering substructure are found. Very high brightness temperatures, Tb ≥ 1013 K, are estimated at 4.8 and 22 GHz even taking into account the refractive scattering. Our findings suggest a clear dominance of the particle energy density over the magnetic field energy density in the core of this quasar.

  7. Post-shock temperatures in minerals. [infrared detection of brightness temperature

    NASA Technical Reports Server (NTRS)

    Raikes, S. A.; Ahrens, T. J.

    1978-01-01

    Post-shock temperatures were measured in a wide variety of materials, including those of geophysical interest such as silicates by using an infrared detector to determine the brightness temperature of samples shocked to pressures in the range 5 to approximately 30 GPa. Measurements were made in the 4.5 to 5.75 micron and in the 7 to 14 micron wavelength ranges. Reproducible results, withe the temperatures in the two wavelength bands generally in excellent agreement, were obtained for aluminum-2024 (10.5 to 33 GPa; 125 to 260 C), stainless steel-304 (11.5 to 50 GPa; 80 to 350 C), crystalline quartz (5.0 to 21.5 GPa; 80 to 250 C) forsterite (7.5 to 28.0 GPa; approximately 30 to 160 C) and Bamble bronzite (6.0 to 26.0 GPa; approximately 30 to 225 C). Results are generally much higher at low pressures than the values calculated assuming a hydrodynamic rheology and isentropic release parallel to the Hugoniot but tend towards them at higher pressures.

  8. 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.

  9. The correlation of Skylab L-band brightness temperatures with antecedent precipitation

    NASA Technical Reports Server (NTRS)

    Mcfarland, M. J.

    1975-01-01

    The S194 L-band radiometer flown on the Skylab mission measured terrestrial radiation at the microwave wavelength of 21.4 cm. The terrain emissivity at this wavelength is strongly dependent on the soil moisture content, which can be inferred from antecedent precipitation. For the Skylab data acquisition pass from the Oklahoma panhandle to southeastern Texas on 11 June 1973, the S194 brightness temperatures are highly correlated with antecedent precipitation from the preceding eleven day period, but very little correlation was apparent for the preceding five day period. The correlation coefficient between the averaged antecedent precipitation index values and the corresponding S194 brightness temperatures between 230 K and 270 K, the region of apparent response to soil moisture in the data, was -0.97. The equation of the linear least squares line is given.

  10. 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

  11. Nist Microwave Blackbody: The Design, Testing, and Verification of a Conical Brightness Temperature Source

    NASA Astrophysics Data System (ADS)

    Houtz, Derek Anderson

    Microwave radiometers allow remote sensing of earth and atmospheric temperatures from space, anytime, anywhere, through clouds, and in the dark. Data from microwave radiometers are high-impact operational inputs to weather forecasts, and are used to provide a vast array of climate data products including land and sea surface temperatures, soil moisture, ocean salinity, cloud precipitation and moisture height profiles, and even wind speed and direction, to name a few. Space-borne microwave radiometers have a major weakness when it comes to long-term climate trends due to their lack of traceability. Because there is no standard, or absolute reference, for microwave brightness temperature, nationally or internationally, individual instruments must each rely on their own internal calibration source to set an absolute reference to the fundamental unit of Kelvin. This causes each subsequent instrument to have a calibration offset and there is no 'true' reference. The work introduced in this thesis addresses this vacancy by proposing and introducing a NIST microwave brightness temperature source that may act as the primary reference. The NIST standard will allow pre-launch calibration of radiometers across a broad range of remote sensing pertinent frequencies between 18 GHz and 220 GHz. The blackbody will be capable of reaching temperatures ranging between liquid nitrogen boiling at approximately 77 K and warm-target temperature of 350 K. The brightness temperature of the source has associated standard uncertainty ranging as a function of frequency between 0.084 K and 0.111 K. The standard can be transferred to the calibration source in the instrument, providing traceability of all subsequent measurements back to the primary standard. The development of the NIST standard source involved predicting and measuring its brightness temperature, and minimizing the associated uncertainty of this quantity. Uniform and constant physical temperature along with well characterized and

  12. Hi-C OBSERVATIONS OF SUNSPOT PENUMBRAL BRIGHT DOTS

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

    Alpert, Shane E.; Tiwari, Sanjiv K.; Moore, Ronald L.

    We report observations of bright 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 observation 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

  13. Red-emission phosphor's brightness deterioration by x-ray and brightness recovery phenomenon by heating.

    PubMed

    Nakamura, Masaaki; Chida, Koichi; Inaba, Yohei; Kobayashi, Ryota; Zuguchi, Masayuki

    2017-06-26

    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 brightness 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 brightness 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 brightness 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 brightness. After the irradiation, brightness changes occurring at room temperature and at 80 °C were investigated. The irradiation reduced the brightness of all the phosphors by 5%-10%, but the brightness 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 brightness remained at 95%-98% of the brightness before the x-ray irradiation. The brightness recovery phenomena of Y 2 O 3 :Eu, (Y, Gd, Eu)BO 3 , and YVO 4 :Eu phosphors occurring after brightness 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 brightness changes caused by x-rays were elucidated.

  14. 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.

  15. Benchmark Transiting Brown Dwarf LHS 6343 C: Spitzer Secondary Eclipse Observations Yield Brightness Temperature and Mid-T Spectral Class

    NASA Astrophysics Data System (ADS)

    Montet, Benjamin T.; Johnson, John Asher; Fortney, Jonathan J.; Desert, Jean-Michel

    2016-05-01

    There are no field brown dwarf analogs with measured masses, radii, and luminosities, precluding our ability to connect the population of transiting brown dwarfs with measurable masses and radii and field brown dwarfs with measurable luminosities and atmospheric properties. LHS 6343 C, a weakly irradiated brown dwarf transiting one member of an M+M binary in the Kepler field, provides the first opportunity to probe the atmosphere of a non-inflated brown dwarf with a measured mass and radius. Here, we analyze four Spitzer observations of secondary eclipses of LHS 6343 C behind LHS 6343 A. Jointly fitting the eclipses with a Gaussian process noise model of the instrumental systematics, we measure eclipse depths of 1.06 ± 0.21 ppt at 3.6 μm and 2.09 ± 0.08 ppt at 4.5 μm, corresponding to brightness temperatures of 1026 ± 57 K and 1249 ± 36 K, respectively. We then apply brown dwarf evolutionary models to infer a bolometric luminosity {log}({L}\\star /{L}⊙ )=-5.16+/- 0.04. Given the known physical properties of the brown dwarf and the two M dwarfs in the LHS 6343 system, these depths are consistent with models of a 1100 K T dwarf at an age of 5 Gyr and empirical observations of field T5-6 dwarfs with temperatures of 1070 ± 130 K. We investigate the possibility that the orbit of LHS 6343 C has been altered by the Kozai-Lidov mechanism and propose additional astrometric or Rossiter-McLaughlin measurements of the system to probe the dynamical history of the system.

  16. Study of the model of calibrating differences of brightness temperature from geostationary satellite generated by time zone differences

    NASA Astrophysics Data System (ADS)

    Li, Weidong; Shan, Xinjian; Qu, Chunyan

    2010-11-01

    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 surface). For a geostationary satellite panorama graph at a point of time, the brightness temperature of different zones is unable to represent the thermal radiation information of the surface at the same point of time because of the effect of different sun solar radiation. So it is necessary to calibrate brightness temperature of different zones with respect to the same point of time. A model of calibrating the differences of the brightness 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 brightness 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 brightness temperature rising or dropping caused by time zone differences.

  17. 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

  18. 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.

  19. Hinode observations and 3D magnetic structure of an X-ray bright point

    NASA Astrophysics Data System (ADS)

    Alexander, C. E.; Del Zanna, G.; Maclean, R. C.

    2011-02-01

    Aims: We present complete Hinode Solar Optical Telescope (SOT), X-Ray Telescope (XRT)and EUV Imaging Spectrometer (EIS) observations of an X-ray bright point (XBP) observed 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 bright 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 observed 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 observed 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 bright point of ±15 km s-1 which are observed to change

  20. Simultaneous Assimilation of AMSR-E Brightness Temperature and MODIS LST to Improve Soil Moisture with Dual Ensemble Kalman Smoother

    NASA Astrophysics Data System (ADS)

    Huang, Chunlin; Chen, Weijin; Wang, Weizhen; Gu, Juan

    2017-04-01

    Uncertainties in model parameters can easily cause systematic differences between model states and observations 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 brightness temperature (TB) and MODIS Land Surface 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 observation 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 observed brightness 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.

  1. Organic Nanocrystals with Bright Red Persistent Room-Temperature Phosphorescence for Biological Applications.

    PubMed

    Fateminia, S M Ali; Mao, Zhu; Xu, Shidang; Yang, Zhiyong; Chi, Zhenguo; Liu, Bin

    2017-09-25

    Persistent room-temperature phosphorescence (RTP) in pure organic materials has attracted great attention because of their unique optical properties. The design of organic materials with bright red persistent RTP remains challenging. Herein, we report a new design strategy for realizing high brightness and long lifetime of red-emissive RTP molecules, which is based on introducing an alkoxy spacer between the hybrid units in the molecule. The spacer offers easy Br-H bond formation during crystallization, which also facilitates intermolecular electron coupling to favor persistent RTP. As the majority of RTP compounds have to be confined in a rigid environment to quench nonradiative relaxation pathways for bright phosphorescence emission, nanocrystallization is used to not only rigidify the molecules but also offer the desirable size and water-dispersity for biomedical applications. © 2017 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

  2. 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.

  3. Far infrared and submillimeter brightness temperatures of the giant planets

    NASA Technical Reports Server (NTRS)

    Hildebrand, R. H.; Loewenstein, R. F.; Harper, D. A.; Orton, G. S.; Keene, J.; Whitcomb, S. E.

    1985-01-01

    The brightness temperatures of Jupiter, Saturn, Uranus, and Neptune in the range 35 to 1000 micron. The effective temperatures derived from the measurements, supplemented by shorter wavelength Voyager data for Jupiter and Saturn, are 126.8 + or - 4.5 K, 93.4 + or - 3.3 K, 58.3 + or - 2.0 K, and 60.3 + or - 2.0 K, respectively. The implications of the measurements for bolometric output and for atmospheric structure and composition are discussed. The temperature spectrum of Jupiter shows a strong peak at approx. 350 microns followed by a deep valley at approx. 450 to 500 microns. Spectra derived from model atmospheres qualitatively reproduced these features but do not fit the data closely.

  4. 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.

  5. 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.

  6. Estimation of wind speeds inside Super Typhoon Nepartak from AMSR2 low-frequency brightness temperatures

    NASA Astrophysics Data System (ADS)

    Zhang, Lei; Yin, Xiaobin; Shi, Hanqing; Wang, Zhenzhan; Xu, Qing

    2018-04-01

    Accurate estimations of typhoon-level winds are highly desired over the western Pacific Ocean. A wind speed retrieval algorithm is used to retrieve the wind speeds within Super Typhoon Nepartak (2016) using 6.9- and 10.7-GHz brightness temperatures from the Japanese Advanced Microwave Scanning Radiometer 2 (AMSR2) sensor on board the Global Change Observation Mission-Water 1 (GCOM-W1) satellite. The results show that the retrieved wind speeds clearly represent the intensification process of Super Typhoon Nepartak. A good agreement is found between the retrieved wind speeds and the Soil Moisture Active Passive wind speed product. The mean bias is 0.51 m/s, and the root-mean-square difference is 1.93 m/s between them. The retrieved maximum wind speeds are 59.6 m/s at 04:45 UTC on July 6 and 71.3 m/s at 16:58 UTC on July 6. The two results demonstrate good agreement with the results reported by the China Meteorological Administration and the Joint Typhoon Warning Center. In addition, Feng-Yun 2G (FY-2G) satellite infrared images, Feng-Yun 3C (FY-3C) microwave atmospheric sounder data, and AMSR2 brightness temperature images are also used to describe the development and structure of Super Typhoon Nepartak.

  7. X-ray Observations of the Bright Old Nova V603 Aquilae

    NASA Technical Reports Server (NTRS)

    Mukai, K.; Orio, M.

    2004-01-01

    We report on our Chandra and RXTE observations of the bright 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.

  8. 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.

  9. Development of Yellow Sand Image Products Using Infrared Brightness Temperature Difference Method

    NASA Astrophysics Data System (ADS)

    Ha, J.; Kim, J.; Kwak, M.; Ha, K.

    2007-12-01

    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 surface as well as during nighttime. There had been suggestion of using brightness temperature difference (BTD) between 11 and 12¥ìm. We have found that the technique is highly depends on surface temperature, emissivity, and zenith angle, which results in changing the threshold of BTD. In order to overcome these problems, we have constructed the background brightness 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.

  10. 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.

  11. 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

  12. 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.

  13. Variations in West Antarctic Ice Front and Passive Microwave Brightness Temperature for 8 Years Duration in 2000s

    NASA Astrophysics Data System (ADS)

    Kim, J.; Yu, J.; Wang, L.; Liu, H.

    2017-12-01

    Changes in Antarctic ice sheet are caused by various reasons such as changes in Holocene climate, precipitation, and ocean temperature. Such issues of changes in ice sheet has been mainly focused on the Antarctic peninsula, and it is known that ice retreat of the area is caused by changes in atmospheric and ocean temperatures. For the case of West Antarctica, ice front change research is relatively rarely conducted except the Pine island glacier area. This study has monitored ice front changes of West Antarctica and compared the patterns with the changes in brightness temperature based on remote sensing techniques. We used 2000 Radarsat-1 and 2008 Rasarsat-2 SAR data to delineate coastlines of whole West Antarctica based on the locally thresholding adaptive algorithm. The delineated coast lines are analyzed to figure out ice front change patterns between the duration. The variations in brightness temperature for the same duration are calculated based on Defense Meteorological Satellite Program (DMSP)'s Special Sensor Microwave/Images-Special Sensor Microwave Imager/Sounder (SSM/I-SSMIS) passive microwave data. The results show ice front of West Antarctica shows advancing trend except the pine island glacier area. The brightness temperature had decreasing trend during the study period. It infers that changes in ice front and brightness temperature of West Antarctica have considerable relationships. It is expected that a long term monitoring of the relationship would contribute understanding ice dynamics of West Antarctica significantly.

  14. Atmospheric correction for retrieving ground brightness temperature at commonly-used passive microwave frequencies.

    PubMed

    Han, Xiao-Jing; Duan, Si-Bo; Li, Zhao-Liang

    2017-02-20

    An analysis of the atmospheric impact on ground brightness temperature (Tg) is performed for numerous land surface 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 surfaces with emissivities greater than 0.7, at 6.93 GHz for land surfaces with emissivities greater than 0.8, and at 10.65 GHz for land surfaces 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 brightness 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 observation direction Lwv, total cloud liquid water content Lclw and mean temperature of cloud Tclw), which is important for simultaneously determining the land surface parameters using multi-frequency passive microwave satellite data.

  15. Sparkling extreme-ultraviolet bright dots observed with Hi-C

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

    Régnier, S.; Alexander, C. E.; Walsh, R. W.

    Observing 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 observed 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 bright 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 observations 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

  16. Assessment of mesoscale convective systems using IR brightness temperature in the southwest of Iran

    NASA Astrophysics Data System (ADS)

    Rafati, Somayeh; Karimi, Mostafa

    2017-07-01

    In this research, the spatial and temporal distribution of Mesoscale Convective Systems was assessed in the southwest of Iran using Global merged satellite IR brightness temperature (acquired from Meteosat, GOES, and GMS geostationary satellites) and synoptic station data. Event days were selected using a set of storm reports and precipitation criteria. The following criteria are used to determine the days with occurrence of convective systems: (1) at least one station reported 6-h precipitation exceeding 10 mm and (2) at least three stations reported phenomena related to convection (thunderstorm, lightning, and shower). MCSs were detected based on brightness temperature, maximum areal extent, and duration thresholds (228 K, 10,000 km2, and 3 h, respectively). An MCS occurrence classification system is developed based on mean sea level, 850 and 500 hPa pressure patterns.

  17. Control of dispersed-phase temperature in plasma flows by the spectral-brightness pyrometry method

    NASA Astrophysics Data System (ADS)

    Dolmatov, A. V.; Gulyaev, I. P.; Gulyaev, P. Yu; Iordan, V. I.

    2016-02-01

    In the present work, we propose a new method for measuring the distribution of temperature in the ensembles of condensed-phase particles in plasma spray flows. Interrelation between the spectral temperature of the particles and the distribution of camera brightness signal is revealed. The established inter-relation enables an in-situ calibration of measuring instruments using the objects under study. The spectral-brightness pyrometry method was approbated on a Plazer plasma-arc wire spraying facility at the Paton Institute of Electrical Welding (Ukrainian Academy of Sciences, Kiev) and on the Thermoplasma 50-1 powder spraying facility at the Institute of Theoretical and Applied Mechanics (Russian Academy of Sciences, Siberian Branch, Novosibirsk). The work was supported by the Russian Foundation for Basic Research (Grants Nos. 14-08-90428 and 15-48-00100).

  18. Sensitivity of Support Vector Machine Predictions of Passive Microwave Brightness Temperature Over Snow-covered Terrain in High Mountain Asia

    NASA Astrophysics Data System (ADS)

    Ahmad, J. A.; Forman, B. A.

    2017-12-01

    High Mountain Asia (HMA) serves as a water supply source for over 1.3 billion people, primarily in south-east Asia. Most of this water originates as snow (or ice) that melts during the summer months and contributes to the run-off downstream. In spite of its critical role, there is still considerable uncertainty regarding the total amount of snow in HMA and its spatial and temporal variation. In this study, the NASA Land Information Systems (LIS) is used to model the hydrologic cycle over the Indus basin. In addition, the ability of support vector machines (SVM), a machine learning technique, to predict passive microwave brightness temperatures at a specific frequency and polarization as a function of LIS-derived land surface model output is explored in a sensitivity analysis. Multi-frequency, multi-polarization passive microwave brightness temperatures as measured by the Advanced Microwave Scanning Radiometer - Earth Observing System (AMSR-E) over the Indus basin are used as training targets during the SVM training process. Normalized sensitivity coefficients (NSC) are then computed to assess the sensitivity of a well-trained SVM to each LIS-derived state variable. Preliminary results conform with the known first-order physics. For example, input states directly linked to physical temperature like snow temperature, air temperature, and vegetation temperature have positive NSC's whereas input states that increase volume scattering such as snow water equivalent or snow density yield negative NSC's. Air temperature exhibits the largest sensitivity coefficients due to its inherent, high-frequency variability. Adherence of this machine learning algorithm to the first-order physics bodes well for its potential use in LIS as the observation operator within a radiance data assimilation system aimed at improving regional- and continental-scale snow estimates.

  19. Hi-C Observations of Penumbral Bright Dots: Comparison with the IRIS Results

    NASA Technical Reports Server (NTRS)

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

    2014-01-01

    We observed bright 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 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 and Hi-C had a short observation time. We use 193 A 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 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 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, 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 observed by IRIS.

  20. 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.

  1. Brightness temperature and attenuation statistics at 20.6 and 31.65 GHz

    NASA Technical Reports Server (NTRS)

    Westwater, Edgeworth R.; Falls, M. J.

    1991-01-01

    Attenuation and brightness temperature statistics at 20.6 and 31.65 GHz are analyzed for a year's worth of data. The data were collected in 1988 at Denver and Platteville, Colorado. The locations are separated by 49 km. Single-station statistics are derived for the entire year. Quality control procedures are discussed and examples of their application are given.

  2. 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.

  3. 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.

  4. Height formation of bright points observed by IRIS in Mg II line wings during flux emergence

    NASA Astrophysics Data System (ADS)

    Grubecka, M.; Schmieder, B.; Berlicki, A.; Heinzel, P.; Dalmasse, K.; Mein, P.

    2016-09-01

    Context. A flux emergence in the active region AR 111850 was observed on September 24, 2013 with the Interface Region Imaging Spectrograph (IRIS). Many bright 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 bright points (CBs) observed 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 observed 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 bright points observed 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.

  5. 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.

  6. Effect of bright light at night on core temperature, subjective alertness and performance as a function of exposure time.

    PubMed

    Foret, J; Daurat, A; Tirilly, G

    1998-01-01

    This simulated night shift study measured the effects of moderate bright light (a 4-hour pulse starting at 2000 or 0400) during the exposure night and subsequent night (dim light). Eight young males remained confined with little physical activity to a laboratory in groups of 4. After a night of reference, they were active for 24 hours; then after a morning recovery sleep, they were active again for 16 hours. Continuously measured rectal temperature proved to be immediately sensitive to 4 hours of bright light, particularly when given at the end of the night. Self-assessed alertness and also performance on a task with a high requirement for short-term memory were improved by the exposure to bright light. During the subsequent night the subjects were exposed only to dim light. Core temperature, subjective alertness and performance continued to show a time course depending on the preceding bright light exposure. Probably because evening exposure to bright light and morning sleep both had a phase-delaying effect, the effects on the circadian pacemaker were more pronounced. Thus, for practical applications in long night shifts, bright light can be considered to improve mood and alertness immediately but the possibility of modifying the circadian "clock" during subsequent nights should be taken into consideration, in particular after exposure to bright light in the evening.

  7. Acute effects of bright light and caffeine on nighttime melatonin and temperature levels in women taking and not taking oral contraceptives

    NASA Technical Reports Server (NTRS)

    Wright, K. P. Jr; Myers, B. L.; Plenzler, S. C.; Drake, C. L.; Badia, P.; Czeisler, C. A. (Principal Investigator)

    2000-01-01

    Caffeine and bright light effects on nighttime melatonin and temperature levels in women were tested during the luteal phase of the menstrual cycle (n=30) or the pseudo luteal phase for oral contraceptive users (n=32). Participants were randomly assigned to receive either bright (5000 lux) or dim room light (<88 lux) between 20:00 and 08:00 h under a modified constant routine protocol. Half the subjects in each lighting condition were administered either caffeine (100 mg) or placebo in a double-blind manner at 20:00, 23:00, 02:00 and 05:00 h. Results showed that the combination of bright light and caffeine enhanced nighttime temperature levels to a greater extent than did either caffeine or bright light alone. Both of the latter groups had higher temperature levels relative to the dim light placebo condition and the two groups did not differ. Temperature levels in the bright light caffeine condition were maintained at near peak circadian levels the entire night in the luteal and pseudo luteal phase. Melatonin levels were reduced throughout the duration of bright light exposure for all women. Caffeine reduced the onset of melatonin levels for women in the luteal phase, but it had little effect on melatonin levels for oral contraceptive users. The results for women in the luteal phase of the menstrual cycle are consistent with our previous findings in men. The results also suggest that oral contraceptives may alter the effects of caffeine on nighttime melatonin levels.

  8. Exospheric temperatures deduced from 7320- to 7330-A /O/+//2P/ - O/+//2D// twilight observations

    NASA Technical Reports Server (NTRS)

    Yee, J. H.; Abreu, V. J.

    1982-01-01

    A technique developed to deduce exospheric temperatures from the 7320- to 7330-A emission measured by the visible airglow experiment on board the AE-E satellite is considered. An excess emission in the measured 7320- to 7330-A brightness is noticed as a result of the interaction between the spacecraft and the atmosphere. The observed brightnesses are corrected for this effect. The galactic background emission is also carefully subtracted. The deduced temperatures exhibit a positive correlation with solar activity. It varies from approximately 700 K in late 1976 to approximately 1700 K at the peak of this solar cycle. The presence of a nonthermal oxygen corona is considered inconclusive.

  9. Evaluation of AIRS, MODIS, and HIRS 11 Micron Brightness Temperature Difference Changes from 2002 through 2006

    NASA Technical Reports Server (NTRS)

    Broberg, Steven E.; Aumann, Hartmut H.; Gregorich, David T.; Xiong, X.

    2006-01-01

    In an effort to validate the accuracy and stability of AIRS data at low scene temperatures (200-250 K range), we evaluated brightness temperatures at 11 microns with Aqua MODIS band 31 and HIRS/3 channel 8 for Antarctic granules between September 2002 and May 2006. We found excellent agreement with MODIS (at the 0.2 K level) over the full emperature range in data from early in the Aqua mission. However, in more recent data, starting in April 2005, we found a scene temperature dependence in MODIS-AIRS brightness temperature differences, with a discrepancy of 1- 1.5 K at 200 K. The comparison between AIRS and HIRS/3 (channel 8) on NOAA 16 for the same time period yields excellent agreement. The cause and time dependence of the disagreement with MODIS is under evaluation, but the change was coincident with a change in the MODIS production software from collection 4 to 5.

  10. Three-Dimensional Structure and Evolution of Extreme-Ultraviolet Bright Points Observed by STEREO/SECCHI/EUVI

    NASA Technical Reports Server (NTRS)

    Kwon, Ryun Young; Chae, Jongchul; Davila, Joseph M.; Zhang, Jie; Moon, Yong-Jae; Poomvises, Watanachak; Jones, Shaela I.

    2012-01-01

    We unveil the three-dimensional structure of quiet-Sun EUV bright 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 bright points and their temporal evolutions. Among them we present three bright 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 bright points. We conclude that EUV bright points are flaring loops formed by magnetic reconnection and their geometry may represent the reconnected magnetic field lines rather than the separator field lines.

  11. Effect of morning bright light on body temperature, plasma cortisol and wrist motility measured during 24 hour of constant conditions.

    PubMed

    Foret, J; Aguirre, A; Touitou, Y; Clodoré, M; Benoit, O

    1993-06-11

    Using 24 h constant conditions, time course of body temperature, plasma cortisol and wrist motility was measured in response to a 3 day morning 2 h bright light pulse. This protocol demonstrated that a 2000 lux illumination was sufficient to elicit a shift of about 2 h of temperature minimum and cortisol peak. In reference session, actimetric recordings showed a circadian time course, closely in relation with core temperature. Bright light pulse resulted in a decrease of amplitude and a disappearance of circadian pattern of actimetry.

  12. 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

  13. 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

  14. 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.

  15. Evaluation of brightness temperature from a forward model of ground-based microwave radiometer

    NASA Astrophysics Data System (ADS)

    Rambabu, S.; Pillai, J. S.; Agarwal, A.; Pandithurai, G.

    2014-06-01

    Ground-based microwave radiometers are getting great attention in recent years due to their capability to profile the temperature and humidity at high temporal and vertical resolution in the lower troposphere. The process of retrieving these parameters from the measurements of radiometric brightness temperature ( T B ) includes the inversion algorithm, which uses the back ground information from a forward model. In the present study, an algorithm development and evaluation of this forward model for a ground-based microwave radiometer, being developed by Society for Applied Microwave Electronics Engineering and Research (SAMEER) of India, is presented. Initially, the analysis of absorption coefficient and weighting function at different frequencies was made to select the channels. Further the range of variation of T B for these selected channels for the year 2011, over the two stations Mumbai and Delhi is discussed. Finally the comparison between forward-model simulated T B s and radiometer measured T B s at Mahabaleshwar (73.66 ∘E and 17.93∘N) is done to evaluate the model. There is good agreement between model simulations and radiometer observations, which suggests that these forward model simulations can be used as background for inversion models for retrieving the temperature and humidity profiles.

  16. 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.

  17. Possible Bright Starspots on TRAPPIST-1

    NASA Astrophysics Data System (ADS)

    Morris, Brett M.; Agol, Eric; Davenport, James R. A.; Hawley, Suzanne L.

    2018-04-01

    The M8V star TRAPPIST-1 hosts seven roughly Earth-sized planets and is a promising target for exoplanet characterization. Kepler/K2 Campaign 12 observations 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 bright 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 observed more often when the brightest spot is facing the observer, suggesting a correlation between the position of the bright spots and flare events. In addition, these flares may occur preferentially when the spots are increasing in brightness, which suggests that the 3.3-day periodicity may not be a rotational signal, but rather a characteristic timescale of active regions.

  18. Comparison of measured brightness temperatures from SMOS with modelled ones from ORCHIDEE and H-TESSEL over the Iberian Peninsula

    NASA Astrophysics Data System (ADS)

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

    2017-01-01

    L-band radiometry is considered to be one of the most suitable techniques to estimate surface soil moisture (SSM) by means of remote sensing. Brightness temperatures are key in this process, as they are the main input in the retrieval algorithm which yields SSM estimates. The work exposed compares brightness temperatures measured by the SMOS mission to two different sets of modelled ones, over the Iberian Peninsula from 2010 to 2012. The two modelled sets were estimated using a radiative transfer model and state variables from two land-surface models: (i) ORCHIDEE and (ii) H-TESSEL. The radiative transfer model used is the CMEM. 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 at the moment. Further hypotheses are proposed and will be explored in a forthcoming paper. 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.

  19. 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.

  20. Implementation of an Ultra-Bright Thermographic Phosphor for Gas Turbine Engine Temperature Measurements

    NASA Technical Reports Server (NTRS)

    Eldridge, Jeffrey I.; Bencic, Timothy J.; Zhu, Dongming; Cuy, Michael D.; Wolfe, Douglas E.; Allison, Stephen W.; Beshears, David L.; Jenkins, Thomas P.; Heeg, Bauke; Howard, Robert P.; hide

    2014-01-01

    The overall goal of the Aeronautics Research Mission Directorate (ARMD) Seedling Phase II effort was to build on the promising temperature-sensing characteristics of the ultrabright thermographic phosphor Cr-doped gadolinium aluminum perovskite (Cr:GAP) demonstrated in Phase I by transitioning towards an engine environment implementation. The strategy adopted was to take advantage of the unprecedented retention of ultra-bright luminescence from Cr:GAP at temperatures over 1000 C to enable fast 2D temperature mapping of actual component surfaces as well as to utilize inexpensive low-power laser-diode excitation suitable for on-wing diagnostics. A special emphasis was placed on establishing Cr:GAP luminescence-based surface temperature mapping as a new tool for evaluating engine component surface cooling effectiveness.

  1. Microwave attenuation and brightness temperature due to the gaseous atmosphere: A comparison of JPL and CCIR values

    NASA Technical Reports Server (NTRS)

    Smith, E. K.; Waters, J. W.

    1981-01-01

    A sophisticated but flexible radiative transfer program designed to assure internal consistency was used to produce brightness temperature (sky noise temperature in a given direction) and gaseous attenuation curves. The curves, derived from atmospheric models, were compared and a new set was derived for a specified frequency range.

  2. 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.

  3. The First ALMA Observation of a Solar Plasmoid Ejection from an X-Ray Bright Point

    NASA Astrophysics Data System (ADS)

    Shimojo, M.; Hudson, H. S.; White, S. M.; Bastian, T.; Iwai, K.

    2017-12-01

    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 observations is not straightforward. We have observed a plasmoid ejection from an X-ray bright 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 observations with ALMA will be able to match the spatial resolution of space-borne and other high-resolution telescopes.

  4. 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

  5. Preliminary results of radiometric measurements of clear air and cloud brightness (antenna) temperatures at 37GHz

    NASA Astrophysics Data System (ADS)

    Arakelyan, A. K.; Hambaryan, A. K.; Arakelyan, A. A.

    2012-05-01

    In this paper the results of polarization measurements of clear air and clouds brightness temperatures at 37GHz are presented. The results were obtained during the measurements carried out in Armenia from the measuring complex built under the framework of ISTC Projects A-872 and A-1524. The measurements were carried out at vertical and horizontal polarizations, under various angles of sensing by Ka-band combined scatterometric-radiometric system (ArtAr-37) developed and built by ECOSERV Remote Observation Centre Co.Ltd. under the framework of the above Projects. In the paper structural and operational features of the utilized system and the whole measuring complex will be considered and discussed as well.

  6. Passive Microwave Remote Sensing of Colorado Watersheds Using Calibrated, Enhanced-Resolution Brightness Temperatures (CETB) from AMSR-E and SSM/I for Estimation of Snowmelt Timing

    NASA Astrophysics Data System (ADS)

    Johnson, M.; Ramage, J. M.; Troy, T. J.; Brodzik, M. J.

    2017-12-01

    Understanding the timing of snowmelt is critical for water resources management in snow-dominated watersheds. Passive microwave remote sensing has been used to estimate melt-refreeze events through brightness temperature satellite observations taken with sensors like the Special Sensor Microwave Imager (SSM/I) and the Advanced Microwave Scanning Radiometer - Earth Observing System (AMSR-E). Previous studies were limited to lower resolution ( 25 km) datasets, making it difficult to quantify the snowpack in heterogeneous, high-relief areas. This study investigates the use of newly available passive microwave calibrated, enhanced-resolution brightness temperatures (CETB) produced at the National Snow and Ice Data Center to estimate melt timing at much higher spatial resolution ( 3-6 km). CETB datasets generated from SSM/I and AMSR-E records will be used to examine three mountainous basins in Colorado. The CETB datasets retain twice-daily (day/night) observations of brightness temperatures. Therefore, we employ the diurnal amplitude variation (DAV) method to detect melt onset and melt occurrences to determine if algorithms developed for legacy data are valid with the improved CETB dataset. We compare melt variability with nearby stream discharge records to determine an optimum melt onset algorithm using the newly reprocessed data. This study investigates the effectiveness of the CETB product for several locations in Colorado (North Park, Rabbit Ears, Fraser) that were the sites of previous ground/airborne surveys during the NASA Cold Land Processes Field Experiment (CLPX 2002-2003). In summary, this work lays the foundation for the utilization of higher resolution reprocessed CETB data for snow evolution more broadly in a range of environments. Consequently, the new processing methods and improved spatial resolution will enable hydrologists to better analyze trends in snow-dominated mountainous watersheds for more effective water resources management.

  7. Effect of a single 3-hour exposure to bright light on core body temperature and sleep in humans.

    PubMed

    Dijk, D J; Cajochen, C; Borbély, A A

    1991-01-02

    Seven human subjects were exposed to bright light (BL, approx. 2500 lux) and dim light (DL, approx. 6 lux) during 3 h prior to nocturnal sleep, in a cross-over design. At the end of the BL exposure period core body temperature was significantly higher than at the end of the DL exposure period. The difference in core body temperature persisted during the first 4 h of sleep. The latency to sleep onset was increased after BL exposure. Rapid-eye movement sleep (REMS) and slow-wave sleep (SWS; stage 3 + 4 of non-REMS) were not significantly changed. Eight subjects were exposed to BL from 20.30 to 23.30 h while their eyes were covered or uncovered. During BL exposure with uncovered eyes, core body temperature decreased significantly less than during exposure with covered eyes. We conclude that bright light immediately affects core body temperature and that this effect is mediated via the eyes.

  8. The Herschel Bright Sources (HerBS): sample definition and SCUBA-2 observations

    NASA Astrophysics Data System (ADS)

    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.

    2018-01-01

    We present the Herschel Bright Sources (HerBS) sample, a sample of bright, 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) observations 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.

  9. Comparison of solar photospheric bright points between Sunrise observations and MHD simulations

    NASA Astrophysics Data System (ADS)

    Riethmüller, T. L.; Solanki, S. K.; Berdyugina, S. V.; Schüssler, M.; Martínez Pillet, V.; Feller, A.; Gandorfer, A.; Hirzberger, J.

    2014-08-01

    Bright points (BPs) in the solar photosphere are thought to be the radiative signatures (small-scale brightness 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 observations 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 observations. 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 observed BPs are compared with synthetic ones. Again, these are found to match relatively well, except that the observations 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 observation 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 observations, we used the simulations to explore the properties of BPs further. The Stokes V asymmetries increase with the distance to the

  10. Bright light and thermoregulatory responses to exercise.

    PubMed

    Atkinson, G; Barr, D; Chester, N; Drust, B; Gregson, W; Reilly, T; Waterhouse, J

    2008-03-01

    The thermoregulatory responses to morning exercise after exposure to different schedules of bright light were examined. At 07:00 h, six males ran on two occasions in an environmental chamber (temperature = 31.4 +/- 1.0 degrees C, humidity = 66 +/- 6 %) for 40 min at 60 % of maximal oxygen uptake. Participants were exposed to bright light (10,000 lux) either between 22:00 - 23:00 h (BT (low)) or 06:00 - 07:00 h (BT (high)). Otherwise, participants remained in dim light (< 50 lux). It was hypothesized that BT (low) attenuates core temperature during morning exercise via the phase-delaying properties of evening bright light and by avoiding bright light in the morning. Evening bright light in BT (low) suppressed (p = 0.037) the increase in melatonin compared to dim light (1.1 +/- 11.4 vs. 15.2 +/- 19.7 pg x ml (-1)) and delayed (p = 0.034) the core temperature minimum by 1.46 +/- 1.24 h. Core temperature was 0.20 +/- 0.17 degrees C lower in BT (low) compared to BT (high) during the hour before exercise (p = 0.036), with evidence (p = 0.075) that this difference was maintained during exercise. Conversely, mean skin temperature was 1.0 +/- 1.7 degrees C higher during the first 10 min of exercise in BT (low) than in BT (high) (p = 0.030). There was evidence that the increase in perceived exertion was attenuated in BT (low) (p = 0.056). A chronobiologically-based light schedule can lower core temperature before and during morning exercise in hot conditions.

  11. 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> <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('https://ntrs.nasa.gov/search.jsp?R=20060046588&hterms=Antarctic+icebergs&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAntarctic%2Bicebergs','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060046588&hterms=Antarctic+icebergs&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAntarctic%2Bicebergs"><span>Antarctic Iceberg Tracking Based on Time Series of Aqua AMSRE Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Blonski, Slawomir; Peterson, Craig</p> <p>2006-01-01</p> <p><span class="hlt">Observations</span> of icebergs are identified as one of the requirements for the GEOSS (Global Earth <span class="hlt">Observation</span> System of Systems) in the area of reducing loss of life and property from natural and human-induced disasters. However, iceberg <span class="hlt">observations</span> are not included among targets in the GEOSS 10-Year Implementation Plan, and thus there is an unfulfilled need for iceberg detection and tracking in the near future. Large Antarctic icebergs have been tracked by the National Ice Center and by the academic community using a variety of satellite sensors including both passive and active microwave imagers, such as SSM/I (Special Sensor Microwave/Imager) deployed on the DMSP (Defense Meteorological Satellite Program) spacecraft. Improvements provided in recent years by NASA and non-NASA satellite radars, scatterometers, and radiometers resulted in an increased number of <span class="hlt">observed</span> icebergs and even prompted a question: Is The Number of Antarctic Icebergs Really Increasing? [D.G. Long, J. Ballantyne, and C. Bertoia, Eos, Transactions of the American Geophysical Union 83 (42): 469 & 474, 15 October 2002]. AMSR-E (Advanced Microwave Scanning Radiometer for the Earth <span class="hlt">Observing</span> System) represents an improvement over SSM/I, its predecessor. AMSR-E has more measurement channels and higher spatial resolution than SSM/I. For example, the instantaneous field of view of the AMSR-E s 89-GHz channels is 6 km by 4 km versus 16 km by 14 km for SSM/I s comparable 85-GHz channels. AMSR-E, deployed on the Aqua satellite, scans across a 1450-km swath and provides <span class="hlt">brightness</span> <span class="hlt">temperature</span> measurements with nearglobal coverage every one or two days. In polar regions, overlapping swaths generate coverage up to multiple times per day and allow for creation of image time series with high temporal resolution. Despite these advantages, only incidental usage of AMSR-E data for iceberg tracking has been reported so far, none in an operational environment. Therefore, an experiment was undertaken in the RPC</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 <span class="hlt">temperatures</span> 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> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130009956','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130009956"><span><span class="hlt">Observations</span> of C-Band <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> and Ocean Surface Wind Speed and Rain Rate from the Hurricane Imaging Radiometer (HIRAD) during GRIP and HS3</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.; Roberts, J. B.; Jones, W. L.; Biswas, S.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; Black, P.; Albers, C.</p> <p>2013-01-01</p> <p>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, <span class="hlt">observations</span> 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 <span class="hlt">observations</span> within a single footprint at nadir angle. Results from the flights during the GRIP and HS3 campaigns will be shown, including images of <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, wind speed, and rain rate. Comparisons will be made with <span class="hlt">observations</span> from other instruments on the campaigns, for which HIRAD <span class="hlt">observations</span> 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.</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 surface, 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> <span class="hlt">temperature</span>, which we call an umbral <span class="hlt">brightness</span> enhancement. This enhanced emission corresponds to a <span class="hlt">temperature</span> 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://hdl.handle.net/2060/20140017807','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140017807"><span>L-Band <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Variations at Dome C and Snow Metamorphism at the Surface</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brucker, Ludovic; Dinnat, Emmanuel; Picard, Ghislain; Champollion, Nicolas</p> <p>2014-01-01</p> <p>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 <span class="hlt">observations</span> 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 <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) (e.g. Macelloni et al., 2007, 2013, Brucker et al., 2011, Champollion et al., 2013). However, to analyze the <span class="hlt">observations</span> 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 <span class="hlt">observations</span> at L-band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150000372','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150000372"><span>Aquarius <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Variations at Dome C and Snow Metamorphism at the Surface. [29</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brucker, Ludovic; Dinnat, Emmanuel Phillippe; Picard, Ghislain; Champollion, Nicolas</p> <p>2014-01-01</p> <p>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 <span class="hlt">observations</span> 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 <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) (e.g. Macelloni et al., 2007, 2013, Brucker et al., 2011, Champollion et al., 2013). However, to analyze the <span class="hlt">observations</span> 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 <span class="hlt">observations</span> at L-band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20170007426','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170007426"><span>Integration of SMAP and SMOS L-Band <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>Bindlish, Rajat; Jackson, Thomas J.; Chan, Steven; Colliander, Andreas; Kerr, Yaan</p> <p>2017-01-01</p> <p>Soil Moisture Active Passive (SMAP) mission and the ESA Soil Moisture and Ocean Salinity (SMOS) missions provide <span class="hlt">brightness</span> <span class="hlt">temperature</span> and soil moisture estimates every 2-3 days. SMAP <span class="hlt">brightness</span> <span class="hlt">temperature</span> <span class="hlt">observations</span> were compared with SMOS <span class="hlt">observations</span> at 40 Degrees incidence angle. The <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from the two missions are not consistent and have a bias of about 2.7K over land with respect to each other. SMAP and SMOS missions use different retrieval algorithms and ancillary datasets which result in further inconsistencies between the soil moisture products. The reprocessed constant-angle SMOS <span class="hlt">brightness</span> <span class="hlt">temperatures</span> were used in the SMAP soil moisture retrieval algorithm to develop a consistent multi-satellite product. The integrated product will have an increased global revisit frequency (1 day) and period of record that would be unattainable by either one of the satellites alone. Results from the development and validation of the integrated product will be presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1912746P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912746P"><span>Using SMOS <span class="hlt">brightness</span> <span class="hlt">temperature</span> and derived surface-soil moisture to characterize surface conditions and validate land surface models.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Polcher, Jan; Barella-Ortiz, Anaïs; Piles, Maria; Gelati, Emiliano; de Rosnay, Patricia</p> <p>2017-04-01</p> <p>The SMOS satellite, operated by ESA, <span class="hlt">observes</span> the surface in the L-band. On continental surface these <span class="hlt">observations</span> are sensitive to moisture and in particular surface-soil moisture (SSM). In this presentation we will explore how the <span class="hlt">observations</span> 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 <span class="hlt">brightness</span> <span class="hlt">temperatures</span> show a good agreement in their temporal evolution, but their spatial structures are not consistent. An empirical orthogonal function analysis of the <span class="hlt">brightness</span> <span class="hlt">temperature</span>'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 <span class="hlt">observations</span> 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</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('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>Surface-induced <span class="hlt">brightness</span> <span class="hlt">temperature</span> 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> <span class="hlt">temperature</span> 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 surface 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> <span class="hlt">temperature</span> 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 surface emissivity variations. The large <span class="hlt">brightness</span> <span class="hlt">temperature</span> differences are sufficient to cause false detection of cirrus clouds from remote sensing data acquired over certain surface 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 surface emissivity effects. In addition, it is recommended that in the future the variation of surface emissivity with wavelength should be taken into account in algorithms for retrieving surface <span class="hlt">temperatures</span> and low-level atmospheric <span class="hlt">temperature</span> and water vapor profiles.</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 surface, 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> <span class="hlt">temperature</span>, which we call an umbral <span class="hlt">brightness</span> enhancement. This enhanced emission corresponds to a <span class="hlt">temperature</span> 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.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 <span class="hlt">temperature</span> 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('https://www.ncbi.nlm.nih.gov/pubmed/23027282','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23027282"><span>High-<span class="hlt">brightness</span> 1.3 μm InAs/GaAs quantum dot tapered laser with high <span class="hlt">temperature</span> stability.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cao, Yulian; Ji, Haiming; Xu, Pengfei; Gu, Yongxian; Ma, Wenquan; Yang, Tao</p> <p>2012-10-01</p> <p>We demonstrate high-<span class="hlt">brightness</span> 1.3 μm tapered lasers with high <span class="hlt">temperature</span> stability by using p-doped InAs/GaAs quantum dots (QDs) as the active region. It is found that the beam quality factor M(2) for the devices is almost unchanged as the light power and <span class="hlt">temperature</span> increase. The almost constant M(2) results from the p-doped QD active region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870020588','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870020588"><span>Satellite-derived ice data sets no. 2: Arctic monthly average microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> and sea ice concentrations, 1973-1976</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Parkinson, C. L.; Comiso, J. C.; Zwally, H. J.</p> <p>1987-01-01</p> <p>A summary data set for four years (mid 70's) of Arctic sea ice conditions is available on magnetic tape. The data include monthly and yearly averaged Nimbus 5 electrically scanning microwave radiometer (ESMR) <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, an ice concentration parameter derived from the <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, monthly climatological surface air <span class="hlt">temperatures</span>, and monthly climatological sea level pressures. All data matrices are applied to 293 by 293 grids that cover a polar stereographic map enclosing the 50 deg N latitude circle. The grid size varies from about 32 X 32 km at the poles to about 28 X 28 km at 50 deg N. The ice concentration parameter is calculated assuming that the field of view contains only open water and first-year ice with an ice emissivity of 0.92. To account for the presence of multiyear ice, a nomogram is provided relating the ice concentration parameter, the total ice concentration, and the fraction of the ice cover which is multiyear ice.</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 <span class="hlt">temperature</span>, 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/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/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 <span class="hlt">temperature</span> variations, and shear flow driven nonmodel MHD wave evolution (self-heating and/or overreflection mechanisms).</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/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/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 <span class="hlt">temperatures</span>, but not always at coronal <span class="hlt">temperatures</span>. 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://hdl.handle.net/2060/20140017808','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140017808"><span>In-situ Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Variability from Ground-based Radiometer Measurements at Dome C in Antarctica Induced by Wind-formed Features</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Royer, A.; Picard, G.; Arnaud, L.; Brucker, L.; Fily, M..</p> <p>2014-01-01</p> <p>Space-borne microwave radiometers are among the most useful tools to study snow and to collect information on the Antarctic climate. They have several advantages over other remote sensing techniques: high sensitivity to snow properties of interest (<span class="hlt">temperature</span>, grain size, density), subdaily coverage in the polar regions, and their <span class="hlt">observations</span> are independent of cloud conditions and solar illumination. Thus, microwave radiometers are widely used to retrieve information over snow-covered regions. For the Antarctic Plateau, many studies presenting retrieval algorithms or numerical simulations have assumed, explicitly or not, that the subpixel-scale heterogeneity is negligible and that the retrieved properties were representative of whole pixels. In this presentation, we investigate the spatial variations of <span class="hlt">brightness</span> <span class="hlt">temperature</span> over arange of a few kilometers in the Dome C area (Antarctic Plateau).</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.ncbi.nlm.nih.gov/pubmed/21196966','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21196966"><span><span class="hlt">Brightness</span> field distributions of microlens arrays using micro molding.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cheng, Hsin-Chung; Huang, Chiung-Fang; Lin, Yi; Shen, Yung-Kang</p> <p>2010-12-20</p> <p>This study describes the <span class="hlt">brightness</span> field distributions of microlens arrays fabricated by micro injection molding (μIM) and micro injection-compression molding (μICM). The process for fabricating microlens arrays used room-<span class="hlt">temperature</span> imprint lithography, photoresist reflow, electroforming, μIM, μICM, and optical properties measurement. Analytical results indicate that the <span class="hlt">brightness</span> field distribution of the molded microlens arrays generated by μICM is better than those made using μIM. Our results further demonstrate that mold <span class="hlt">temperature</span> is the most important processing parameter for <span class="hlt">brightness</span> field distribution of molded microlens arrays made by μIM or μICM.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19750021459','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19750021459"><span>Ground truth data for test sites (SL-4). [thermal radiation <span class="hlt">brightness</span> <span class="hlt">temperature</span> and solar radiation measurments</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1974-01-01</p> <p>Field measurements performed simultaneous with Skylab overpass in order to provide comparative calibration and performance evaluation measurements for the EREP sensors are presented. Wavelength region covered include: solar radiation (400 to 1300 nanometer), and thermal radiation (8 to 14 micrometer). Measurements consisted of general conditions and near surface meteorology, atmospheric <span class="hlt">temperature</span> and humidity vs altitude, the thermal <span class="hlt">brightness</span> <span class="hlt">temperature</span>, total and diffuse solar radiation, direct solar radiation (subsequently analyzed for optical depth/transmittance), and target reflectivity/radiance. The particular instruments used are discussed along with analyses performed. Detailed instrument operation, calibrations, techniques, and errors are given.</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=19910044887&hterms=electron+thomson&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Delectron%2Bthomson','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910044887&hterms=electron+thomson&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Delectron%2Bthomson"><span>Synoptic maps constructed from <span class="hlt">brightness</span> <span class="hlt">observations</span> of Thomson scattering by heliospheric electrons</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hick, P.; Jackson, B.; Schwenn, R.</p> <p>1991-01-01</p> <p><span class="hlt">Observations</span> of the Thomson scattering <span class="hlt">brightness</span> by electrons in the inner heliosphere provide a means of probing the heliospheric electron distributions. An extensive data base of Thomson scattering <span class="hlt">observations</span>, 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 <span class="hlt">observations</span> 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.</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/2060/19820003639','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19820003639"><span>SMMR data set development for GARP. [impact of cross polarization and Faraday rotation on SMMR derived <span class="hlt">brightness</span> <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kogut, J.</p> <p>1981-01-01</p> <p>The NIMBUS 7 Scanning Multichannel Microwave Radiometer (SMMR) data are analyzed. The impact of cross polarization and Faraday rotation on SMMR derived <span class="hlt">brightness</span> <span class="hlt">temperatures</span> is evaluated. The algorithms used to retrieve the geophysical parameters are tested, refined, and compared with values derived by other techniques. The technical approach taken is described and the results presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830064488&hterms=1041&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231041','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830064488&hterms=1041&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3D%2526%25231041"><span>Heavy thunderstorms <span class="hlt">observed</span> over land by the Nimbus 7 scanning multichannel microwave radiometer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Spencer, R. W.; Olson, W. S.; Martin, D. W.; Weinman, J. A.; Santek, D. A.; Wu, R.</p> <p>1983-01-01</p> <p><span class="hlt">Brightness</span> <span class="hlt">temperatures</span> obtained through examination of microwave data from the Nimbus 7 satellite are noted to be much lower than those expected on the strength of radiation emanating from rain-producing clouds. Very cold <span class="hlt">brightness</span> <span class="hlt">temperature</span> cases all coincided with heavy thunderstorm rainfall, with the cold <span class="hlt">temperatures</span> being attributable to scattering by a layer of ice hydrometeors in the upper parts of the storms. It is accordingly suggested that <span class="hlt">brightness</span> <span class="hlt">temperatures</span> <span class="hlt">observed</span> by satellite microwave radiometers can sometimes distinguish heavy rain over land.</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('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://adsabs.harvard.edu/abs/2017AIPC.1857g0006S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1857g0006S"><span>Detecting the <span class="hlt">brightness</span> <span class="hlt">temperature</span> from Landsat-8 thermal infra red scanner preceding the Rinjani strombolian eruption 2015</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suwarsono, Hidayat, Suprapto, Totok; Prasasti, Indah; Parwati, Rokhis Khomarudin, M.</p> <p>2017-07-01</p> <p>At the end of October to early November 2015, Rinjani Volcano that is located in Lombok Island was erupted and has catapulted the ash, pyroclastic and lava flow. The dispersion of this volcanic ash in the atmosphere has been disrupting flights and the three closest airports to be closed for a while. The existence of Rinjani Volcano geographically plays an important role in the survival of life on the island of Lombok, because large areas of land on the island are a part of the Rinjani Volcano landscape. Based on the experience of violent eruptions that have occurred in the 13th century ago, the monitoring of the volcanism activity of this volcano needs to be done intensively and continuously. This is something important to do an early detection efforts of the volcanic eruption. These efforts need to be done as a preparedness effort in order to minimize adverse impacts that may occur as a result of this eruption. This research tries to detect the volcanic eruption precursor based on changes in <span class="hlt">temperature</span> conditions of the crater and the surrounding area. We use the medium resolution satellite data, Thermal Infra Red Scanner (TIRS), on board Landsat-8, to monitor the <span class="hlt">brightness</span> <span class="hlt">temperature</span> as a representative of surface <span class="hlt">temperature</span> of the volcanic region. The results showed that the <span class="hlt">brightness</span> <span class="hlt">temperature</span> derived from Landsat-8 TIRS is very usefull to predict the strombolian eruption which will occur in the near future. The use of multitemporal image data is important to understand the dynamics of volcanism activity over time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..18.8576L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..18.8576L"><span>Simulations of a Canadian snowpack <span class="hlt">brightness</span> <span class="hlt">temperatures</span> using SURFEX-Crocus for Snow Water Equivalent (SWE) retrievals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Larue, Fanny; Royer, Alain; De Sève, Danielle; Langlois, Alexandre; Roy, Alexandre; Saint-Jean-Rondeau, Olivier</p> <p>2016-04-01</p> <p>In Quebec, the water associated to snowmelt represents 30% of the annual electricity production so that the snow cover evaluation in real time is of primary interest. The key variable is snow water equivalent (SWE) which describes the evolution of a global seasonal snow cover. However, the sparse distribution of meteorological stations in northern Québec generates great uncertainty in the extrapolation of SWE. On the contrary, the spatial and temporal coverage of satellite data offer a source of information with a high potential when considered as an alternative to the poor spatial distribution of in-situ information. Thus, this project aims to improve the prediction of SWE by assimilation of satellite passive microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (Tb) <span class="hlt">observations</span>, independently of any ground <span class="hlt">observations</span>. The snowpack evolution is simulated by the French snow model SURFEX-Crocus, driven by the Canadian atmospheric model GEM with a spatial resolution of 10 km. The bias of the atmospheric model and the impact of initialization errors on the simulated SWE were quantified from our ground measurements. To assimilate satellite <span class="hlt">observations</span>, the multi-layered snow model is first coupled with a radiative transfer model using the Dense Media Radiative transfer theory (the DMRT-ML model) to estimate the microwave snow emission of the simulated snowpack. In order to retrieve simulated Tb in frequencies of interest (i.e. sensitive to snow dielectric properties), the snow microstructure needs to be well parameterized. It was shown in previous studies that the specific surface area (SSA) of snow grains is a well-defined parameter to describe the size and the shape of snow grains and which allows reproducible field measurements. SURFEX-Crocus estimates a SSA for each simulated snow layer, however, the snow microstructure in DMRT-ML is defined per layer by monodisperse optical radius of grain (~ 1/SSA) and by the stickiness which is not known. It thus becomes necessary to introduce</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/2015JGRD..120.6145X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JGRD..120.6145X"><span>Polarization signatures and <span class="hlt">brightness</span> <span class="hlt">temperatures</span> caused by horizontally oriented snow particles at microwave bands: Effects of atmospheric absorption</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xie, Xinxin; Crewell, Susanne; Löhnert, Ulrich; Simmer, Clemens; Miao, Jungang</p> <p>2015-06-01</p> <p>This study analyzes the effects of atmospheric absorption and emission on the polarization difference (PD) and <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) generated by horizontally oriented snow particles. A three-layer plane-parallel atmosphere model is used in conjunction with a simplified radiative transfer (RT) scheme to illustrate the combined effects of dichroic and nondichroic media on microwave signatures <span class="hlt">observed</span> by ground-based and spaceborne sensors. Based on idealized scenarios which encompass a dichroic snow layer and adjacent nondichroic layers composed of supercooled liquid water (SCLW) droplets and water vapor, we demonstrate that the presence of atmospheric absorption/emission enhances TB and damps PD when <span class="hlt">observed</span> from the ground. From a spaceborne perspective, however, TB can be reduced or enhanced by an absorbing/emitting layer above the snow layer, while a strong absorbing/emitting layer below the dichroic snow layer may even enhance PD. The induced PD and TB, which rely on snow microphysical assumptions, can vary up to 2 K and 10 K, respectively, due to the <span class="hlt">temperature</span>-dependent absorption of SCLW. RT calculations based on 223 snowfall profiles selected from European Centre for Medium-Range Weather Forecasts data sets indicate that the existence of SCLW has a noticeable impact on PD and TB at three window frequencies (150 GHz, 243 GHz, and 664 GHz) during snowfall. Our results imply that while polarimetric channels at the three window channels have the potential for snowfall characterization, accurate information on liquid water is required to correctly interpret the polarimetric <span class="hlt">observations</span>.</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://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> <span class="hlt">temperatures</span> are examined; spatial variability in the <span class="hlt">brightness</span> <span class="hlt">temperatures</span> of consolidated ice in winter and spring cyclic increases and decrease in <span class="hlt">brightness</span> <span class="hlt">temperatures</span> 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 <span class="hlt">temperature</span> and surface characteristics in the variability of spring <span class="hlt">brightness</span> <span class="hlt">temperatures</span> 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('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> <span class="hlt">Temperature</span> <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> <span class="hlt">temperature</span> 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 surface of the neutron star and can even exceed the magnetic field energy density at that surface. 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://ntrs.nasa.gov/search.jsp?R=20160014922&hterms=soil&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsoil','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160014922&hterms=soil&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsoil"><span>Assimilation of SMOS <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> or Soil Moisture Retrievals into a Land Surface Model</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.; Reichle, Rolf H.</p> <p>2016-01-01</p> <p>Three different data products from the Soil Moisture Ocean Salinity (SMOS) mission are assimilated separately into the Goddard Earth <span class="hlt">Observing</span> 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 <span class="hlt">brightness</span> <span class="hlt">temperature</span> (Tb) <span class="hlt">observations</span> 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 <span class="hlt">observation</span>-minus-forecast residuals and analysis increments reveal large differences in how the <span class="hlt">observations</span> add value in the Tb and SM retrieval assimilation systems. The distinct patterns of these diagnostics in the two systems reflect <span class="hlt">observation</span> 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</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('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 surface 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 surface, 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 surface from below. The heat source may have been transient (triggered by impact heating). Alternatively, internal <span class="hlt">temperatures</span> may be above the eutectic <span class="hlt">temperature</span> of subsurface brines, in which case fluids may exist at depth on Ceres today.</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 <span class="hlt">temperature</span> 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('http://adsabs.harvard.edu/abs/2017AGUFM.C51C1000N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51C1000N"><span>Analyzing Snowpack Metrics Over Large Spatial Extents Using Calibrated, Enhanced-Resolution <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Data and Long Short Term Memory Artificial Neural Networks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Norris, W.; J Q Farmer, C.</p> <p>2017-12-01</p> <p>Snow water equivalence (SWE) is a difficult metric to measure accurately over large spatial extents; snow-tell sites are too localized, and traditional remotely sensed <span class="hlt">brightness</span> <span class="hlt">temperature</span> data is at too coarse of a resolution to capture variation. The new Calibrated Enhanced-Resolution <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> (CETB) data from the National Snow and Ice Data Center (NSIDC) offers remotely sensed <span class="hlt">brightness</span> <span class="hlt">temperature</span> data at an enhanced resolution of 3.125 km versus the original 25 km, which allows for large spatial extents to be analyzed with reduced uncertainty compared to the 25km product. While the 25km <span class="hlt">brightness</span> <span class="hlt">temperature</span> data has proved useful in past research — one group found decreasing trends in SWE outweighed increasing trends three to one in North America; other researchers used the data to incorporate winter conditions, like snow cover, into ecological zoning criterion — with the new 3.125 km data, it is possible to derive more accurate metrics for SWE, since we have far more spatial variability in measurements. Even with higher resolution data, using the 37 - 19 GHz frequencies to estimate SWE distorts the data during times of melt onset and accumulation onset. Past researchers employed statistical splines, while other successful attempts utilized non-parametric curve fitting to smooth out spikes distorting metrics. In this work, rather than using legacy curve fitting techniques, a Long Short Term Memory (LSTM) Artificial Neural Network (ANN) was trained to perform curve fitting on the data. LSTM ANN have shown great promise in modeling time series data, and with almost 40 years of data available — 14,235 days — there is plenty of training data for the ANN. LSTM's are ideal for this type of time series analysis because they allow important trends to persist for long periods of time, but ignore short term fluctuations; since LSTM's have poor mid- to short-term memory, they are ideal for smoothing out the large spikes generated in the melt</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPhCS.771a2033H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPhCS.771a2033H"><span>Sky <span class="hlt">brightness</span> and twilight measurements at Jogyakarta city, Indonesia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Herdiwijaya, Dhani</p> <p>2016-11-01</p> <p>The sky <span class="hlt">brightness</span> measurements were performed using a portable photometer. A pocket-sized and low-cost photometer has 20 degree area measurement, and spectral ranges between 320-720 nm with output directly in magnitudes per arc second square (mass) unit. The sky <span class="hlt">brightness</span> with 3 seconds temporal resolutions was recorded at Jogyakarta city (110° 25’ E; 70° 52’ S; elevation 100 m) within 136 days in years from 2014 to 2016. The darkest night could reach 22.61 mpass only in several seconds, with mean value 18.8±0.7 mpass and <span class="hlt">temperature</span> variation 23.1±1.2 C. The difference of mean sky <span class="hlt">brightness</span> between before and after midnight was about -0.76 mpass or 2.0 times brighter. Moreover, the sky <span class="hlt">brightness</span> and <span class="hlt">temperature</span> fluctuations were more stable in after midnight than in before midnight. It is suggested that city light pollution affects those variations, and subsequently duration of twilight. By comparing twilight <span class="hlt">brightness</span> for several places, we also suggest a 17° solar dip or about 66 minutes before sunrise for new time of Fajr prayer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890059295&hterms=deming&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Ddeming','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890059295&hterms=deming&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Ddeming"><span>A search for p-mode oscillations of Jupiter - Serendipitous <span class="hlt">observations</span> of nonacoustic thermal wave structure</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Deming, Drake; Mumma, Michael J.; Espenak, Fred; Jennings, Donald E.; Kostiuk, Theodor; Wiedemann, Gunter</p> <p>1989-01-01</p> <p>Frequencies for the p-mode oscillations of Jupiter have been determined, and infrared <span class="hlt">brightness</span> <span class="hlt">temperature</span> fluctuations are used to search for the modes. Measurements of the infrared intensity of the Jovian disk were obtained in a broad bandwidth using a 20-element linear array. No p-mode oscillations were <span class="hlt">observed</span> at the 0.07-K level in the 8-13-micron <span class="hlt">brightness</span> <span class="hlt">temperature</span>. The results suggest that Jovian p modes are not likely to have <span class="hlt">observable</span> amplitudes. A prominent nonacoustic wave-like structure in the 8-13-micron <span class="hlt">brightness</span> <span class="hlt">temperature</span> is found both at 20 deg N and at the equator.</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 <span class="hlt">temperatures</span> (in the range 3.2x10^4 - 3.5x10^5 K), but not always at coronal <span class="hlt">temperatures</span>. 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://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 <span class="hlt">temperature</span> 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('http://adsabs.harvard.edu/abs/2007AGUSMSM31B..05B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUSMSM31B..05B"><span>Extreme Magnetosphere-Ionosphere Coupling at the Plasmapause: a - In-A <span class="hlt">Bright</span> SAR Arc</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Baumgardner, J.; Wroten, J.; Semeter, J.; Mendillo, M.; Kozyra, J.</p> <p>2007-05-01</p> <p>Heat conduction from the ring current - plasmapause interaction region generates high electron <span class="hlt">temperature</span> within the ionosphere that drive stable auroral red (SAR) arc emission at 6300 A. On the night of 29 October 1991, a SAR arc was <span class="hlt">observed</span> using an all-sky imager and meridional imaging spectrograph at Millstone Hill. At xxxx UT, the SAR arc was south of Millstone at approximate L = 2 and reached emission levels of 13,000 rayleighs (R). Over two solar cycle of imaging <span class="hlt">observations</span> have been made at Millstone Hill, and SAR arc <span class="hlt">brightness</span> levels (excluding this event) averaged ~ 500 R. Simultaneous <span class="hlt">observations</span> using the incoherent scatter radar (ISR), a DMSP satellite pass, the MSIS neutral atmosphere and SAR arc modeling using the Rees and Roble formalism succeeded in simulations of the <span class="hlt">observed</span> emission. The reason for the unusual <span class="hlt">brightness</span> was not the extreme <span class="hlt">temperatures</span> achieved (and therefore heat conduction input), but the fact that the end of the plasmapause field line where the elevated Te values were measured did not occur in the ionospheric trough, but equatorward of it, thereby having far more ambient electrons to heat and subsequently collide with atomic oxygen. This unusual spatial geometry probably resulted from unusual convection patterns early in a superstorm scenario.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.H31G1489C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.H31G1489C"><span>Utilizing Machine Learning to Downscale SMAP L3_SM_P <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> in Iowa for Agricultural Applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chakrabarti, S.; Judge, J.; Bindlish, R.; Bongiovanni, T.; Jackson, T. J.</p> <p>2016-12-01</p> <p>The NASA Soil Moisture Active Passive (SMAP) mission provides global <span class="hlt">observations</span> of <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TB) at 36km. For these <span class="hlt">observations</span> to be relevant to studies in agricultural regions, the TB values need to be downscaled to finer resolutions. In this study, a machine learning algorithm is introduced for downscaling of TB from 36km to 9km. The algorithm uses image segmentation to cluster the study region based on meteorological and land cover similarity, followed by a support vector machine based regression that computes the value of the disaggregated TB at all pixels. High resolution remote sensing products such as land surface <span class="hlt">temperature</span>, normalized difference vegetation index, enhanced vegetation index, precipitation, soil texture, and land-cover were used for downscaling. The algorithm was implemented in Iowa, United States, during the growing season from April to July 2015 when the SMAP L3-SM_AP TB product at 9 km was available for comparison. In addition, the downscaled estimates from the algorithm are compared with 9km TB obtained by resampling SMAP L1B_TB product at 36km. It was found that the downscaled TB were very similar to the SMAP-L3_SM _AP TB product, even for vegetated areas with a mean difference ≤ 5K. However, the standard deviation of the downscaled was lower by 7K than that of the AP product. The probability density functions of the downscaled TB were similar to the SMAP- TB. The results indicate that these downscaling algorithms may be used for downscaling TB using complex non-linear correlations on a grid without using active microwave <span class="hlt">observations</span>.</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><span class="hlt">Temperatures</span> of the martian surface 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 <span class="hlt">temperatures</span> 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 <span class="hlt">temperatures</span> 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 <span class="hlt">temperature</span> of Arsia Mons varies by a factor of nearly two each day; large diurnal <span class="hlt">temperature</span> 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> <span class="hlt">temperatures</span> 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 surface scattering.</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://hdl.handle.net/2060/20020023443','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020023443"><span><span class="hlt">Temperature</span> Map of the Perseus Cluster of Galaxies <span class="hlt">Observed</span> with ASCA</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Furusho, T.; Yamasaki, N. Y.; Ohashi, T.; Shibata, R.; Ezawa, H.; White, Nicholas E. (Technical Monitor)</p> <p>2000-01-01</p> <p>We present two-dimensional <span class="hlt">temperature</span> map of the Perseus cluster based on multi-pointing <span class="hlt">observations</span> with the Advanced Spacecraft for Cosmology Astrophysics (ASCA) Gas Imaging Spectrometer (GIS), covering a region with a diameter of approximately 2 deg. By correcting for the effect of the X-ray telescope response, the <span class="hlt">temperatures</span> were estimated from hardness ratios and the complete <span class="hlt">temperature</span> structure of the cluster with a spatial resolution of about 100 kpc was obtained for the first time. There is an extended cool region with a diameter of approximately 20 arcmin and kT approx. 5 keV at about 20 arcmin east from the cluster center. This region also shows higher surface <span class="hlt">brightness</span> and is surrounded by a large ring-like hot region with kT approx. > 7 keV, and likely to be a remnant of a merger with a poor cluster. Another extended cool region is extending outward from the IC 310 subcluster. These features and the presence of several other hot and cool blobs suggest that this rich cluster has been formed as a result of a repetition of many subcluster mergers.</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('http://adsabs.harvard.edu/abs/2012AGUFMIN43C1538N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMIN43C1538N"><span>Low Latency Workflow Scheduling and an Application of Hyperspectral <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nguyen, P. T.; Chapman, D. R.; Halem, M.</p> <p>2012-12-01</p> <p>New system analytics for Big Data computing holds the promise of major scientific breakthroughs and discoveries from the exploration and mining of the massive data sets becoming available to the science community. However, such data intensive scientific applications face severe challenges in accessing, managing and analyzing petabytes of data. While the Hadoop MapReduce environment has been successfully applied to data intensive problems arising in business, there are still many scientific problem domains where limitations in the functionality of MapReduce systems prevent its wide adoption by those communities. This is mainly because MapReduce does not readily support the unique science discipline needs such as special science data formats, graphic and computational data analysis tools, maintaining high degrees of computational accuracies, and interfacing with application's existing components across heterogeneous computing processors. We address some of these limitations by exploiting the MapReduce programming model for satellite data intensive scientific problems and address scalability, reliability, scheduling, and data management issues when dealing with climate data records and their complex <span class="hlt">observational</span> challenges. In addition, we will present techniques to support the unique Earth science discipline needs such as dealing with special science data formats (HDF and NetCDF). We have developed a Hadoop task scheduling algorithm that improves latency by 2x for a scientific workflow including the gridding of the EOS AIRS hyperspectral <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> (BT). This workflow processing algorithm has been tested at the Multicore Computing Center private Hadoop based Intel Nehalem cluster, as well as in a virtual mode under the Open Source Eucalyptus cloud. The 55TB AIRS hyperspectral L1b <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> record has been gridded at the resolution of 0.5x1.0 degrees, and we have computed a 0.9 annual anti-correlation to the El Nino Southern oscillation in</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/2017AGUFM.C51C0996R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C51C0996R"><span>Testing Snow Melt Algorithms in High Relief Topography Using Calibrated Enhanced-Resolution <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span>, Hunza River Basin, Pakistan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramage, J. M.; Brodzik, M. J.; Hardman, M.; Troy, T. J.</p> <p>2017-12-01</p> <p>Snow is a vital part of the terrestrial hydrological cycle, a crucial resource for people and ecosystems. In mountainous regions snow is extensive, variable, and challenging to document. Snow melt timing and duration are important factors affecting the transfer of snow mass to soil moisture and runoff. Passive microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> (Tb) changes at 36 and 18 GHz are a sensitive way to detect snow melt onset due to their sensitivity to the abrupt change in emissivity. They are widely used on large icefields and high latitude watersheds. The coarse resolution ( 25 km) of historically available data has precluded effective use in high relief, heterogeneous regions, and gaps between swaths also create temporal data gaps at lower latitudes. New enhanced resolution data products generated from a scatterometer image reconstruction for radiometer (rSIR) technique are available at the original frequencies. We use these Calibrated Enhanced-resolution <span class="hlt">Brightness</span> (CETB) <span class="hlt">Temperatures</span> Earth System Data Records (ESDR) to evaluate existing snow melt detection algorithms that have been used in other environments, including the cross polarized gradient ratio (XPGR) and the diurnal amplitude variations (DAV) approaches. We use the 36/37 GHz (3.125 km resolution) and 18/19 GHz (6.25 km resolution) vertically and horizontally polarized datasets from the Special Sensor Microwave Imager (SSM/I) and Advanced Microwave Radiometer for EOS (AMSR-E) and evaluate them for use in this high relief environment. The new data are used to assess glacier and snow melt records in the Hunza River Basin [area 13,000 sq. km, located at 36N, 74E], a tributary to the Upper Indus Basin, Pakistan. We compare the melt timing results visually and quantitatively to the corresponding EASE-Grid 2.0 25-km dataset, SRTM topography, and surface <span class="hlt">temperatures</span> from station and reanalysis data. The new dataset is coarser than the topography, but is able to differentiate signals of melt/refreeze timing for</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 <span class="hlt">temperature</span> 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 <span class="hlt">temperatures</span> of Ceres' surface 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/2015AMT.....8..369R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AMT.....8..369R"><span>Adaptive neuro-fuzzy inference system for <span class="hlt">temperature</span> and humidity profile retrieval from microwave radiometer <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>Ramesh, K.; Kesarkar, A. P.; Bhate, J.; Venkat Ratnam, M.; Jayaraman, A.</p> <p>2015-01-01</p> <p>The retrieval of accurate profiles of <span class="hlt">temperature</span> and water vapour is important for the study of atmospheric convection. Recent development in computational techniques motivated us to use adaptive techniques in the retrieval algorithms. In this work, we have used an adaptive neuro-fuzzy inference system (ANFIS) to retrieve profiles of <span class="hlt">temperature</span> and humidity up to 10 km over the tropical station Gadanki (13.5° N, 79.2° E), India. ANFIS is trained by using <span class="hlt">observations</span> of <span class="hlt">temperature</span> and humidity measurements by co-located Meisei GPS radiosonde (henceforth referred to as radiosonde) and microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> <span class="hlt">observed</span> by radiometrics multichannel microwave radiometer MP3000 (MWR). ANFIS is trained by considering these <span class="hlt">observations</span> during rainy and non-rainy days (ANFIS(RD + NRD)) and during non-rainy days only (ANFIS(NRD)). The comparison of ANFIS(RD + NRD) and ANFIS(NRD) profiles with independent radiosonde <span class="hlt">observations</span> and profiles retrieved using multivariate linear regression (MVLR: RD + NRD and NRD) and artificial neural network (ANN) indicated that the errors in the ANFIS(RD + NRD) are less compared to other retrieval methods. The Pearson product movement correlation coefficient (r) between retrieved and <span class="hlt">observed</span> profiles is more than 92% for <span class="hlt">temperature</span> profiles for all techniques and more than 99% for the ANFIS(RD + NRD) technique Therefore this new techniques is relatively better for the retrieval of <span class="hlt">temperature</span> profiles. The comparison of bias, mean absolute error (MAE), RMSE and symmetric mean absolute percentage error (SMAPE) of retrieved <span class="hlt">temperature</span> and relative humidity (RH) profiles using ANN and ANFIS also indicated that profiles retrieved using ANFIS(RD + NRD) are significantly better compared to the ANN technique. The analysis of profiles concludes that retrieved profiles using ANFIS techniques have improved the <span class="hlt">temperature</span> retrievals substantially; however, the retrieval of RH by all techniques considered in this paper (ANN, MVLR and</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('http://hdl.handle.net/2060/19930017284','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930017284"><span><span class="hlt">Brightness</span> <span class="hlt">temperature</span> and attenuation diversity statistics at 20.6 and 31.65 GHz for the Colorado Research Network</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Westwater, Ed R.; Falls, M. J.; Fionda, E.</p> <p>1992-01-01</p> <p>A limited network of four dual-channel microwave radiometers, with frequencies of 20.6 and 31.65 GHz, was operated in the front range of eastern Colorado from 1985 to 1988. Data, from November 1987 through October 1988 are analyzed to determine both single-station and joint-station <span class="hlt">brightness</span> <span class="hlt">temperature</span> and attenuation statistics. Only zenith <span class="hlt">observations</span> were made. The spatial separations of the stations varied from 50 km to 190 km. Before the statistics were developed, the data were screened by rigorous quality control methods. One such method, that of 20.6 vs. 31.65 GHz scatter plots, is analyzed in detail, and comparisons are made of measured vs calculated data. At 20.6 and 31.65 GHz, vertical attenuations of 5 and 8 dB are exceeded 0.01 percent of the time. For these four stations and at the same 0.01 percent level, diversity gains from 6 to 8 dB are possible with the 50 to 190 km separations.</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 <span class="hlt">temperature</span> 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://ntrs.nasa.gov/search.jsp?R=19870037740&hterms=mass+wasting&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dmass%2Bwasting','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870037740&hterms=mass+wasting&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dmass%2Bwasting"><span>Lunar and Venusian radar <span class="hlt">bright</span> rings</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Thompson, T. W.; Saunders, R. S.; Weissman, D. E.</p> <p>1986-01-01</p> <p>Twenty-one lunar craters have radar <span class="hlt">bright</span> ring appearances which are analogous to eleven complete ring features in the earth-based 12.5 cm <span class="hlt">observations</span> of Venus. Radar ring diameters and widths for the lunar and Venusian features overlap for sizes from 45 to 100 km. Radar <span class="hlt">bright</span> 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 <span class="hlt">bright</span> areas of the Venusian rings are also associated with the slopes on the rims of craters. The lunar craters have evolved to radar <span class="hlt">bright</span> 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 <span class="hlt">bright</span> ring appearance. It is proposed that the Venusian features with complete <span class="hlt">bright</span> 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 <span class="hlt">bright</span> rings via analogous surface processes.</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> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7809567','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7809567"><span>Dark goggles and <span class="hlt">bright</span> light improve circadian rhythm adaptation to night-shift work.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Eastman, C I; Stewart, K T; Mahoney, M P; Liu, L; Fogg, L F</p> <p>1994-09-01</p> <p>We compared the contributions of <span class="hlt">bright</span> light during the night shift and dark goggles during daylight for phase shifting the circadian rhythm of <span class="hlt">temperature</span> to realign with a 12-hour shift of sleep. After 10 baseline days there were 8 night-work/day-sleep days. <span class="hlt">Temperature</span> was continuously recorded from 50 subjects. There were four groups in a 2 x 2 design: light (<span class="hlt">bright</span>, dim), goggles (yes, no). Subjects were exposed to <span class="hlt">bright</span> light (about 5,000 lux) for 6 hours on the first 2 night shifts. Dim light was < 500 lux. Both <span class="hlt">bright</span> light and goggles were significant factors for producing circadian rhythm phase shifts. The combination of <span class="hlt">bright</span> light plus goggles was the most effective, whereas the combination of dim light and no goggles was the least effective. The <span class="hlt">temperature</span> rhythm either phase advanced or phase delayed when it aligned with daytime sleep. However, when subjects did not have goggles only phase advances occurred. Goggles were necessary for producing phase delays. The most likely explanation is that daylight during the travel-home window after a night shift inhibits phase-delay shifts, and goggles can prevent this inhibition. Larger <span class="hlt">temperature</span>-rhythm phase shifts were associated with better subjective daytime sleep, less subjective fatigue and better mood.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.H21H1497K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.H21H1497K"><span>Spatial and Temporal Patterns of SMAP <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> for Use in Level 1 TB Characterization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, E. J.</p> <p>2015-12-01</p> <p>1. IntroductionThe recent launch of NASA's Soil Moisture Active Passive (SMAP) mission [Entekhabi, et al] has opened the door to improved <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) calibration of satellite L-band microwave radiometers, through the use of SMAP's lower noise performance and better immunity to man-made interference (vs. ESA's Soil Moisture Ocean Salinity (SMOS) mission [Kerr, et al]), better spatial resolution (vs. NASA's Aquarius sea surface salinity mission [Le Vine, et al]), and cleaner antenna pattern (vs. SMOS). All three radiometers use/used large homogeneous places on Earth's surface as calibration targets—parts of the ocean, Antarctica, and tropical forests. Despite the recent loss of Aquarius data, there is still hope for creating a longer-term L-band data set that spans the timeframe of all 3 missions. 2. Description of Analyses and Expected Results In this paper, we analyze SMAP <span class="hlt">brightness</span> <span class="hlt">temperature</span> data to quantify the spatial and temporal characteristics of external target areas in the oceans, Antarctica, forests, and other areas. Existing analyses have examined these targets in terms of averages, standard deviations, and other basic statistics (for Aquarius & SMOS as well). This paper will approach the problem from a signal processing perspective. Coupled with the use of SMAP's novel RFI-mitigated TBs, and the aforementioned lower noise and cleaner antenna pattern, it is expected that of the 3 L-band missions, SMAP should do the best job of characterizing such external targets. The resulting conclusions should be useful to extract the best possible TB calibration from all 3 missions, helping to inter-compare the TB from the 3 missions, and to eventually inter-calibrate the TBs into a single long-term dataset.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Icar..298...98S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Icar..298...98S"><span>Diviner lunar radiometer gridded <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from geodesic binning of modeled fields of view</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sefton-Nash, E.; Williams, J.-P.; Greenhagen, B. T.; Aye, K.-M.; Paige, D. A.</p> <p>2017-12-01</p> <p>An approach is presented to efficiently produce high quality gridded data records from the large, global point-based dataset returned by the Diviner Lunar Radiometer Experiment aboard NASA's Lunar Reconnaissance Orbiter. The need to minimize data volume and processing time in production of science-ready map products is increasingly important with the growth in data volume of planetary datasets. Diviner makes on average >1400 <span class="hlt">observations</span> per second of radiance that is reflected and emitted from the lunar surface, using 189 detectors divided into 9 spectral channels. Data management and processing bottlenecks are amplified by modeling every <span class="hlt">observation</span> as a probability distribution function over the field of view, which can increase the required processing time by 2-3 orders of magnitude. Geometric corrections, such as projection of data points onto a digital elevation model, are numerically intensive and therefore it is desirable to perform them only once. Our approach reduces bottlenecks through parallel binning and efficient storage of a pre-processed database of <span class="hlt">observations</span>. Database construction is via subdivision of a geodesic icosahedral grid, with a spatial resolution that can be tailored to suit the field of view of the <span class="hlt">observing</span> instrument. Global geodesic grids with high spatial resolution are normally impractically memory intensive. We therefore demonstrate a minimum storage and highly parallel method to bin very large numbers of data points onto such a grid. A database of the pre-processed and binned points is then used for production of mapped data products that is significantly faster than if unprocessed points were used. We explore quality controls in the production of gridded data records by conditional interpolation, allowed only where data density is sufficient. The resultant effects on the spatial continuity and uncertainty in maps of lunar <span class="hlt">brightness</span> <span class="hlt">temperatures</span> is illustrated. We identify four binning regimes based on trades between the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C44B..06K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C44B..06K"><span>Enhanced-Resolution Satellite Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Records for Mapping Boreal-Arctic Landscape Freeze-Thaw Heterogeneity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Y.; Du, J.; Kimball, J. S.</p> <p>2017-12-01</p> <p>The landscape freeze-thaw (FT) status derived from satellite microwave remote sensing is closely linked to vegetation phenology and productivity, surface energy exchange, evapotranspiration, snow/ice melt dynamics, and trace gas fluxes over land areas affected by seasonally frozen <span class="hlt">temperatures</span>. A long-term global satellite microwave Earth System Data Record of daily landscape freeze-thaw status (FT-ESDR) was developed using similar calibrated 37GHz, vertically-polarized (V-pol) <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (Tb) from SMMR, SSM/I, and SSMIS sensors. The FT-ESDR shows mean annual spatial classification accuracies of 90.3 and 84.3 % for PM and AM overpass retrievals relative surface air <span class="hlt">temperature</span> (SAT) measurement based FT estimates from global weather stations. However, the coarse FT-ESDR gridding (25-km) is insufficient to distinguish finer scale FT heterogeneity. In this study, we tested alternative finer scale FT estimates derived from two enhanced polar-grid (3.125-km and 6-km resolution), 36.5 GHz V-pol Tb records derived from calibrated AMSR-E and AMSR2 sensor <span class="hlt">observations</span>. The daily FT estimates are derived using a modified seasonal threshold algorithm that classifies daily Tb variations in relation to grid cell-wise FT thresholds calibrated using ERA-Interim reanalysis based SAT, downscaled using a digital terrain map and estimated <span class="hlt">temperature</span> lapse rates. The resulting polar-grid FT records for a selected study year (2004) show mean annual spatial classification accuracies of 90.1% (84.2%) and 93.1% (85.8%) for respective PM (AM) 3.125km and 6-km Tb retrievals relative to in situ SAT measurement based FT estimates from regional weather stations. Areas with enhanced FT accuracy include water-land boundaries and mountainous terrain. Differences in FT patterns and relative accuracy obtained from the enhanced grid Tb records were attributed to several factors, including different noise contributions from underlying Tb processing and spatial mismatches between Tb</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>Surface-induced <span class="hlt">brightness</span> <span class="hlt">temperature</span> 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> <span class="hlt">temperature</span> 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 surface 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> <span class="hlt">temperature</span> 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 surface 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 surface emissivity with wavelength should be taken into account in algorithms for retrieving surface <span class="hlt">temperatures</span> and low-level atmospheric <span class="hlt">temperature</span> and water vapor profiles.</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 <span class="hlt">temperature</span> 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://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> </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/2017AGUFM.A13H2205M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A13H2205M"><span>The Rapid Intensification of Typhoon Soudelor (2015) Explored through Next-Generation 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>Munsell, E.; Braun, S. A.; Zhang, F.</p> <p>2017-12-01</p> <p>The dynamics that govern the intensification of tropical cyclones (TC) are dominated by rapidly evolving moist convective processes in the inner-core region. Remotely sensed satellite <span class="hlt">observations</span> are typically available but in the past have lacked the necessary resolution to sufficiently examine TC intensification processes. However, as a result of the recent launch of next-generation high-resolution satellites (JMA's Himawari-8 and NOAA/NASA's GOES-16), the spatial and temporal frequency of remotely-sensed <span class="hlt">observations</span> of TCs have increased significantly. This study utilizes <span class="hlt">brightness</span> <span class="hlt">temperatures</span> <span class="hlt">observed</span> by the Advanced Himawari Imager to examine the structure of Typhoon Soudelor (2015) throughout its rapid intensification (RI) from a tropical storm to a super typhoon. Wavenumber decompositions are performed on <span class="hlt">brightness</span> <span class="hlt">temperature</span> fields that correspond to channels sensitive to upper-, mid-, and lower-level water vapor, and IR longwave radiation, to study wave features associated with the inner-core region. A scale-separation is also performed to assess the degree to which the intensification processes are dominated by phenomenon of various wavelengths. Higher-order wavenumbers reveal asymmetric features that propagate outwards from the storm on short time scales ( 1-2 h). The identification of these waves and their contribution to intensification is ongoing. A deterministic forecast of Typhoon Soudelor performed using a convection-permitting WRF simulation coupled to an Ensemble Kalman Filter that assimilates <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, accurately captures the TCs RI event. The Community Radiative Transfer Model (CRTM) is used to produce simulated <span class="hlt">brightness</span> <span class="hlt">temperature</span> fields for the applicable channels. The model demonstrates the ability to reproduce the <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperatures</span> in great detail, including smaller-scale features such as primary rainbands and the eye; however, a uniform warm bias is present. It is hypothesized that this likely results</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040171655&hterms=sensitivity+scale&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dsensitivity%2Bscale','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040171655&hterms=sensitivity+scale&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dsensitivity%2Bscale"><span>Comparison of Local Scale Measured and Modeled <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> and Snow Parameters from the CLPX 2003 by Means of a Dense Medium Radiative Transfer Theory Model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tedescol, Marco; Kim, Edward J.; Cline, Don; Graf, Tobias; Koike, Toshio; Armstrong, Richard; Brodzik, Mary J.; Hardy, Janet</p> <p>2004-01-01</p> <p>Microwave remote sensing offers distinct advantages for <span class="hlt">observing</span> the cryosphere. Solar illumination is not required, and spatial and temporal coverage are excellent from polar-orbiting satellites. Passive microwave measurements are sensitive to the two most useful physical quantities for many hydrological applications: physical <span class="hlt">temperature</span> and water content/state. Sensitivity to the latter is a direct result of the microwave sensitivity to the dielectric properties of natural media, including snow, ice, soil (frozen or thawed), and vegetation. These considerations are factors motivating the development of future cryospheric satellite remote sensing missions, continuing and improving on a 26-year microwave measurement legacy. Perhaps the biggest issues regarding the use of such satellite measurements involve how to relate parameter values at spatial scales as small as a hectare to <span class="hlt">observations</span> with sensor footprints that may be up to 25 x 25 km. The NASA Cold-land Processes Field Experiment (CLPX) generated a dataset designed to enhance understanding of such scaling issues. CLPX <span class="hlt">observations</span> were made in February (dry snow) and March (wet snow), 2003 in Colorado, USA, at scales ranging from plot scale to 25 x 25 km satellite footprints. Of interest here are passive microwave <span class="hlt">observations</span> from ground-based, airborne, and satellite sensors, as well as meteorological and snowpack measurements that will enable studies of the effects of spatial heterogeneity of surface conditions on the <span class="hlt">observations</span>. Prior to performing such scaling studies, an evaluation of snowpack forward modelling at the plot scale (least heterogeneous scale) is in order. This is the focus of this paper. Many forward models of snow signatures (<span class="hlt">brightness</span> <span class="hlt">temperatures</span>) have been developed over the years. It is now recognized that a dense medium radiative transfer (DMRT) treatment represents a high degree of physical fidelity for snow modeling, yet dense medium models are particularly sensitive to</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('https://ntrs.nasa.gov/search.jsp?R=19760035678&hterms=mcdonald&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmcdonald','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760035678&hterms=mcdonald&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmcdonald"><span>The night sky <span class="hlt">brightness</span> at McDonald Observatory</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kalinowski, J. K.; Roosen, R. G.; Brandt, J. C.</p> <p>1975-01-01</p> <p>Baseline <span class="hlt">observations</span> of the night sky <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> are found. Possible causes for these variations are discussed. The largest variation in sky <span class="hlt">brightness</span> found during a single night is approximately a factor of two, a value which corresponds to a factor-of-four variation in airglow <span class="hlt">brightness</span>. The data are used to comment on the accuracy of previously published surface photometry of M 81.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A13M..04G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A13M..04G"><span>Analyzing nearly four decades of historical radiosonde <span class="hlt">observations</span> of tropical tropopause layer and cold-point <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gilford, D.; Randel, W. J.</p> <p>2017-12-01</p> <p>An understanding of historical trends and variability in the thermal structure of the tropical tropopause layer (TTL) is important for assessing climate and investigating TTL processes. In particular, the cold-point tropopause (CPT) plays an important role in stratospheric dehydration, the potential intensities of tropical cyclones, and other forms of stratospheric-tropospheric coupling. Uncertainties and biases of in-situ <span class="hlt">observations</span>, however, make long-term estimation of TTL <span class="hlt">temperatures</span> challenging, especially in the early decades of the satellite era. The goal of this study is to construct and analyze a long-term record of radiosondes <span class="hlt">temperatures</span> with minimal biases. <span class="hlt">Temperature</span> <span class="hlt">observations</span> from 1979-present are drawn from the Integrated Global Radiosonde Archive version 2 (IGRA2). Vertically integrated radiosonde <span class="hlt">temperatures</span> are compared with <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from the Microwave Sounding Units (MSU) Lower Stratosphere channel to identify the radiosonde stations with the smallest temporal discontinuities. Insights from this comparison highlight the importance of independent measurements when evaluating TTL <span class="hlt">temperatures</span>. The 38-year dataset constructed from IGRA2 stations with the smallest biases spans the tropics and has high vertical resolution, permitting reasonable estimates of the CPT <span class="hlt">temperature</span>. Radiosonde <span class="hlt">temperatures</span> show good agreement with GPS radio occultation measurements over the past decade. A multivariate regression model incorporating the Quasi-Biennial Oscillation and the El Nino Southern Oscillation is fit to the deseasonalized data to evaluate the spatial and temporal structures in its variability. Long-term trends in CPT <span class="hlt">temperatures</span> are considered in the context of historical estimates from climate models. Correlations with TTL water vapor concentrations from the Stratospheric Water and OzOne Satellite Homogenized (SWOOSH) data set suggest a strong relationship between the historically <span class="hlt">observed</span> CPT <span class="hlt">temperatures</span> and dehydration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020091934','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020091934"><span>Retrieval of Precipitation Profiles from Multiresolution, Multifrequency, Active and Passive 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>Grecu, Mircea; Anagnostou, Emmanouil N.; Olson, William S.; Starr, David OC. (Technical Monitor)</p> <p>2002-01-01</p> <p>In this study, a technique for estimating vertical profiles of precipitation from multifrequency, multiresolution active and passive microwave <span class="hlt">observations</span> is investigated using both simulated and airborne data. The technique is applicable to the Tropical Rainfall Measuring Mission (TRMM) satellite multi-frequency active and passive <span class="hlt">observations</span>. These <span class="hlt">observations</span> are characterized by various spatial and sampling resolutions. This makes the retrieval problem mathematically more difficult and ill-determined because the quality of information decreases with decreasing resolution. A model that, given reflectivity profiles and a small set of parameters (including the cloud water content, the intercept drop size distribution, and a variable describing the frozen hydrometeor properties), simulates high-resolution <span class="hlt">brightness</span> <span class="hlt">temperatures</span> is used. The high-resolution simulated <span class="hlt">brightness</span> <span class="hlt">temperatures</span> are convolved at the real sensor resolution. An optimal estimation procedure is used to minimize the differences between simulated and <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperatures</span>. The retrieval technique is investigated using cloud model synthetic and airborne data from the Fourth Convection And Moisture Experiment. Simulated high-resolution <span class="hlt">brightness</span> <span class="hlt">temperatures</span> and reflectivities and airborne <span class="hlt">observation</span> strong are convolved at the resolution of the TRMM instruments and retrievals are performed and analyzed relative to the reference data used in <span class="hlt">observations</span> synthesis. An illustration of the possible use of the technique in satellite rainfall estimation is presented through an application to TRMM data. The study suggests improvements in combined active and passive retrievals even when the instruments resolutions are significantly different. Future work needs to better quantify the retrievals performance, especially in connection with satellite applications, and the uncertainty of the models used in retrieval.</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('https://ntrs.nasa.gov/search.jsp?R=19950031639&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D60%26Ntt%3Dwater','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950031639&hterms=water&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D60%26Ntt%3Dwater"><span>Simulations of the effects of water vapor, cloud liquid water, and ice on AMSU moisture channel <span class="hlt">brightness</span> <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Muller, Bradley M.; Fuelberg, Henry E.; Xiang, Xuwu</p> <p>1994-01-01</p> <p>Radiative transfer simulations are performed to determine how water vapor and nonprecipitating cloud liquid water and ice particles within typical midlatitude atmospheres affect <span class="hlt">brightness</span> <span class="hlt">temperatures</span> T(sub B)'s of moisture sounding channels used in the Advanced Microwave Sounding Unit (AMSU) and AMSU-like instruments. The purpose is to promote a general understanding of passive top-of-atmosphere T(sub B)'s for window frequencies at 23.8, 89.0, and 157.0 GHz, and water vapor frequencies at 176.31, 180.31, and 182.31 GHz by documenting specific examples. This is accomplished through detailed analyses of T(sub B)'s for idealized atmospheres, mostly representing temperate conditions over land. Cloud effects are considered in terms of five basic properties: droplet size distribution, phase, liquid or ice water content, altitude, and thickness. Effects on T(sub B) of changing surface emissivity also are addressed. The <span class="hlt">brightness</span> <span class="hlt">temperature</span> contribution functions are presented as an aid to physically interpreting AMSU T(sub B)'s. Both liquid and ice clouds impact the T(sub B)'s in a variety of ways. The T(sub B)'s at 23.8 and 89 GHz are more strongly affected by altostratus liquid clouds than by cirrus clouds for equivalent water paths. In contrast, channels near 157 and 183 GHz are more strongly affected by ice clouds. Higher clouds have a greater impact on 157- and 183-GHz T(sub B)'s than do lower clouds. Clouds depress T(sub B)'s of the higher-frequency channels by suppressing, but not necessarily obscuring, radiance contributions from below. Thus, T(sub B)'s are less closely associated with cloud-top <span class="hlt">temperatures</span> than are IR radiometric <span class="hlt">temperatures</span>. Water vapor alone accounts for up to 89% of the total attenuation by a midtropospheric liquid cloud for channels near 183 GHz. The Rayleigh approximation is found to be adequate for typical droplet size distributions; however, Mie scattering effects from liquid droplets become important for droplet size distribution</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..122.8593F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..122.8593F"><span>Assessment of COSMIC radio occultation and AIRS hyperspectral IR sounder <span class="hlt">temperature</span> products in the stratosphere using <span class="hlt">observed</span> radiances</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Feltz, M. L.; Knuteson, R. O.; Revercomb, H. E.</p> <p>2017-08-01</p> <p>Upper air <span class="hlt">temperature</span> is defined as an essential climate variable by the World Meteorological Organization. Two remote sensing technologies being promoted for monitoring stratospheric <span class="hlt">temperatures</span> are GPS radio occultation (RO) and spectrally resolved IR radiances. This study assesses RO and hyperspectral IR sounder derived <span class="hlt">temperature</span> products within the stratosphere by comparing IR spectra calculated from GPS RO and IR sounder products to coincident IR <span class="hlt">observed</span> radiances, which are used as a reference standard. RO dry <span class="hlt">temperatures</span> from the University Corporation for Atmospheric Research (UCAR) Constellation <span class="hlt">Observing</span> System for Meteorology, Ionosphere, and Climate (COSMIC) mission are compared to NASA Atmospheric Infrared Sounder (AIRS) retrievals using a previously developed profile-to-profile collocation method and vertical <span class="hlt">temperature</span> averaging kernels. <span class="hlt">Brightness</span> <span class="hlt">temperatures</span> (BTs) are calculated for both COSMIC and AIRS <span class="hlt">temperature</span> products and are then compared to coincident AIRS measurements. The COSMIC calculated minus AIRS measured BTs exceed the estimated 0.5 K measurement uncertainty for the winter time extratropics around 35 hPa. These differences are attributed to seasonal UCAR COSMIC biases. Unphysical vertical oscillations are seen in the AIRS L2 <span class="hlt">temperature</span> product in austral winter Antarctic regions, and results imply a small AIRS tropical warm bias around 35 hPa in the middle stratosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016A%26A...589A..46S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016A%26A...589A..46S"><span>Are solar <span class="hlt">brightness</span> variations faculae- or spot-dominated?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shapiro, A. I.; Solanki, S. K.; Krivova, N. A.; Yeo, K. L.; Schmutz, W. K.</p> <p>2016-05-01</p> <p>Context. Regular spaceborne measurements have revealed that solar <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> variability. We investigate how the solar <span class="hlt">brightness</span> variability and its facular and spot contributions depend on the wavelength, timescale of variability, and position of the <span class="hlt">observer</span> relative to the ecliptic plane. Methods: We performed calculations with the SATIRE model, which returns solar <span class="hlt">brightness</span> with daily cadence from solar disc area coverages of various magnetic features. We took coverages as seen by an Earth-based <span class="hlt">observer</span> 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 <span class="hlt">observer</span> 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 <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> variations are faculae-dominated for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JApSp..84..657K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JApSp..84..657K"><span>Energy and Emission Characteristics of a Short-Arc Xenon Flash Lamp Under "Saturated" Optical <span class="hlt">Brightness</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>Kamrukov, A. S.; Kireev, S. G.; Kozlov, N. P.; Shashkovskii, S. G.</p> <p>2017-09-01</p> <p>We present the results of a study of the electrical, energy, and spectral <span class="hlt">brightness</span> characteristics of an experimental three-electrode high-pressure xenon flash lamp under conditions ensuring close to maximum possible spectral <span class="hlt">brightness</span> for the xenon emission. We show that under saturated optical <span class="hlt">brightness</span> conditions (<span class="hlt">brightness</span> <span class="hlt">temperature</span> in the visible region of the spectrum 30,000 K), emission of a pulsed discharge in xenon is quite different from the emission from an ideal blackbody: the maximum <span class="hlt">brightness</span> <span class="hlt">temperatures</span> are 24,000 K in the short-wavelength UV region and 19,000 K in the near IR range. The relative fraction of UV radiation in the emission spectrum of the lamp is >50%, which lets us consider such lamps as promising broadband sources of radiation with high spectral <span class="hlt">brightness</span> for many important practical applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7481410','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7481410"><span>Circadian rhythm adaptation to simulated night shift work: effect of nocturnal <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>Eastman, C I; Liu, L; Fogg, L F</p> <p>1995-07-01</p> <p>We compared <span class="hlt">bright</span>-light durations of 6, 3 and 0 hours (i.e. dim light) during simulated night shifts for phase shifting the circadian rectal <span class="hlt">temperature</span> rhythm to align with a 12-hour shift of the sleep schedule. After 10 baseline days there were 8 consecutive night-work, day-sleep days, with 8-hour sleep (dark) periods. The <span class="hlt">bright</span> light (about 5,000 lux, around the baseline <span class="hlt">temperature</span> minimum) was used during all 8 night shifts, and dim light was < 500 lux. This was a field study in which subjects (n = 46) went outside after the night shifts and slept at home. Substantial circadian adaptation (i.e. a large cumulative <span class="hlt">temperature</span> rhythm phase shift) was produced in many subjects in the <span class="hlt">bright</span> light groups, but not in the dim light group. Six and 3 hours of <span class="hlt">bright</span> light were each significantly better than dim light for phase shifting the <span class="hlt">temperature</span> rhythm, but there was no significant difference between 6 and 3 hours. Thus, durations > 3 hours are probably not necessary in similar shift-work situations. Larger <span class="hlt">temperature</span> rhythm phase shifts were associated with better subjective daytime sleep, less subjective fatigue and better overall mood.</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://pubs.er.usgs.gov/publication/70027286','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027286"><span><span class="hlt">Observations</span> and <span class="hlt">temperatures</span> of Io's Pele Patera from Cassini and Galileo spacecraft images</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Radebaugh, J.; McEwen, A.S.; Milazzo, M.P.; Keszthelyi, L.P.; Davies, A.G.; Turtle, E.P.; Dawson, D.D.</p> <p>2004-01-01</p> <p>Pele has been the most intense high-<span class="hlt">temperature</span> hotspot on Io to be continuously active during the Galileo monitoring from 1996-2001. A suite of characteristics suggests that Pele is an active lava lake inside a volcanic depression. In 2000-2001, Pele was <span class="hlt">observed</span> by two spacecraft, Cassini and Galileo. The Cassini <span class="hlt">observations</span> revealed that Pele is variable in activity over timescales of minutes, typical of active lava lakes in Hawaii and Ethiopia. These <span class="hlt">observations</span> also revealed that the short-wavelength thermal emission from Pele decreases with rotation of Io by a factor significantly greater than the cosine of the emission angle, and that the color <span class="hlt">temperature</span> becomes more variable and hotter at high emission angles. This behavior suggests that a significant portion of the visible thermal emission from Pele comes from lava fountains within a topographically confined lava body. High spatial resolution, nightside images from a Galileo flyby in October 2001 revealed a large, relatively cool (< 800 K) region, ringed by <span class="hlt">bright</span> hotspots, and a central region of high thermal emission, which is hypothesized to be due to fountaining and convection in the lava lake. Images taken through different filters revealed color <span class="hlt">temperatures</span> of 1500 ?? 80 K from Cassini ISS data and 1605 ?? 220 and 1420 ?? 100 K from small portions of Galileo SSI data. Such <span class="hlt">temperatures</span> are near the upper limit for basaltic compositions. Given the limitations of deriving lava eruption <span class="hlt">temperature</span> in the absence of in situ measurement, it is possible that Pele has lavas with ultramafic compositions. The long-lived, vigorous activity of what is most likely an actively overturning lava lake in Pele Patera indicates that there is a strong connection to a large, stable magma source region. ?? 2003 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14564893','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14564893"><span>The effect of a change in sleep-wakefulness timing, <span class="hlt">bright</span> light and physical exercise interventions on 24-hour patterns of performance, mood and body <span class="hlt">temperature</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Iskra-Golec, I; Fafrowicz, M; Marek, T; Costa, G; Folkard, S; Foret, J; Kundi, M; Smith, L</p> <p>2001-12-01</p> <p>Experiments consisting of baseline, <span class="hlt">bright</span> light and physical exercise studies were carried out to compare the effect of a 9-hour delay in sleep-wakefulness timing, and the effects of <span class="hlt">bright</span> light and physical exercise interventions on 24-hour patterns of performance, mood and body <span class="hlt">temperature</span> were examined. Each study comprised a 24-hour constant routine at the beginning followed by 3 night shifts and 24-hour constant routine at the end. Performance on tasks differing in cognitive load, mood and body <span class="hlt">temperature</span> was measured during each constant routine and the interventions were applied during the night shifts. The 24-hour pattern of alertness and performance on the tasks with low cognitive load in post-treatment conditions followed the change in sleep-wakefulness timing while more cognitively loaded tasks tended to show a reverse trend when compared to pre-treatment conditions. There was a phase delay around 4 hours in circadian rhythms of body <span class="hlt">temperature</span> in post-treatment conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930041423&hterms=pig&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpig','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930041423&hterms=pig&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpig"><span>PIG (partially ionized globule) anatomy - Density and <span class="hlt">temperature</span> structure of the <span class="hlt">bright</span>-rimmed globule IC 1396E</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Serabyn, E.; Guesten, R.; Mundy, L.</p> <p>1993-01-01</p> <p>The density and <span class="hlt">temperature</span> structure of the <span class="hlt">bright</span>-rimmed cometary globule IC 1396E is estimated, and the possibility that recent internal star formation was triggered by the ionization front in its southern surface is assessed. On the basis of NH3 data, gas <span class="hlt">temperatures</span> in the globule are found to increase outward from the center, from a minimum of 17 K in its tail to a maximum of 26 K on the surface most directly facing the stars ionizing IC 1396. On the basis of a microturbulent radiative transfer code to model the radial dependence of the CS line intensities, and also the intensities of the optically thin 2-1 and 5-4 lines toward the cloud center, a radial density dependence of r exp -1.55 to r exp -1.75 is found.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AtmRe.120..268M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AtmRe.120..268M"><span>Cloud episode propagation over the Indonesian Maritime Continent from 10 years of infrared <span class="hlt">brightness</span> <span class="hlt">temperature</span> <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>Marzuki; Hashiguchi, Hiroyuki; Yamamoto, Masayuki K.; Yamamoto, Mamoru; Mori, Shuichi; Yamanaka, Manabu D.; Carbone, Richard E.; Tuttle, John D.</p> <p>2013-02-01</p> <p>The cloud-top <span class="hlt">brightness</span> <span class="hlt">temperature</span> data from 2001 to 2010 are used to derive a climatology of deep convection duration, span, and propagation speed over the Indonesian Maritime Continent (10°S-10°N, 80°E-160°E). The full domain of study is divided into northern (0°-10°N) and southern (0°-10°S) regions to investigate the seasonal and latitudinal variabilities of cloud streaks. The ratio of westward- to eastward-propagating cloud streaks is found to be approximately 3:1. Westward-moving streaks generally have longer spans and faster speeds than eastward-moving systems. Coherent episodes of westward- (eastward-) propagating systems have 9.5 (7.5) h durations and 519 (378) km spans on average; most episodes have zonal phase speeds of 6-30 m s- 1. Median zonal phase speeds of 14.2 (westward) and 13.5 m s- 1 (eastward) are found for events with > 1000 km spans and > 20 h durations. The recurrence frequency, which is categorized from 1 event per day to 1 event per month, is also discussed. The latitudinal and seasonal dependences of statistical properties are strongly influenced by the Inter-Tropical Convergence Zone annual cycle. The number of westward-migrating systems is significant every month, while eastward-migrating systems strongly vary by season and latitude. Eastward migration is less frequent in the southern region during June, July and August (JJA) and in the northern region during December, January and February (DJF). In the northern region, the westward-propagating events' mean span is much longer during JJA, September, October, and November (SON) than the other periods; this effect is partially due to the favorable environmental shear conditions necessary to sustain a long-lived system. Eastward- and westward-propagating events are found during the shortwave heating and dissipation modes of diurnal cycle phase. Thus, thermal forcing, which is associated with the elevated terrain found over the islands and the land-sea interface, is dominant on a</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>Surface and Atmospheric Contributions to Passive Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> 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> <span class="hlt">temperatures</span> (TBs) and the physical state of the underlying atmosphere and surface. These relationships are nonlinear, such that inversions are ill ]posed especially over variable land surfaces. 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 surface, 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 surface, 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 surfaces.</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> <span class="hlt">temperature</span> ({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> </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/2018SPIE10514E..0FK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10514E..0FK"><span>Diode lasers optimized in <span class="hlt">brightness</span> for fiber laser pumping</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kelemen, M.; Gilly, J.; Friedmann, P.; Hilzensauer, S.; Ogrodowski, L.; Kissel, H.; Biesenbach, J.</p> <p>2018-02-01</p> <p>In diode laser applications for fiber laser pumping and fiber-coupled direct diode laser systems high <span class="hlt">brightness</span> becomes essential in the last years. Fiber coupled modules benefit from continuous improvements of high-power diode lasers on chip level regarding output power, efficiency and beam characteristics resulting in record highbrightness values and increased pump power. To gain high <span class="hlt">brightness</span> not only output power must be increased, but also near field widths and far field angles have to be below a certain value for higher power levels because <span class="hlt">brightness</span> is proportional to output power divided by beam quality. While fast axis far fields typically show a current independent behaviour, for broadarea lasers far-fields in the slow axis suffer from a strong current and <span class="hlt">temperature</span> dependence, limiting the <span class="hlt">brightness</span> and therefore their use in fibre coupled modules. These limitations can be overcome by carefully optimizing chip <span class="hlt">temperature</span>, thermal lensing and lateral mode structure by epitaxial and lateral resonator designs and processing. We present our latest results for InGaAs/AlGaAs broad-area single emitters with resonator lengths of 4mm emitting at 976nm and illustrate the improvements in beam quality over the last years. By optimizing the diode laser design a record value of the <span class="hlt">brightness</span> for broad-area lasers with 4mm resonator length of 126 MW/cm2sr has been demonstrated with a maximum wall-plug efficiency of more than 70%. From these design also pump modules based on 9 mini-bars consisting of 5 emitters each have been realized with 360W pump power.</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 <span class="hlt">temperature</span>, 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('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/2016cosp...41E.169B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E.169B"><span>Comparison Between AQUARIUS and SMOS <span class="hlt">brightness</span> <span class="hlt">temperatures</span> for Heterogeneous Land Areas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benlloch, Amparo; Lopez-Baeza, Ernesto; Tenjo, Carolina; Navarro, Enrique</p> <p>2016-07-01</p> <p>Intercomparison between Aquarius and SMOS <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TBs) over land surfaces is more challenging than over oceans because land footprints are more heterogeneous. In this work we are comparing Aquarius and SMOS TBs under coherente conditions obtained both by considering similar areas, according to land uses and by stratifying by means of TVDI (<span class="hlt">Temperature</span> Vegetation Dryness Index) that accounts for the dynamics of the vegetation instead of assuming static characteristics as in the previous approches. The area of study was chosen in central Spain where we could get a significant number of matches between both instruments. The study period corresponded to 2012-2014. SMOS level-3 data were obtained from the Centre Aval de Traitement des Données SMOS (CATDS) and Aquarius' from the Physical Oceanography Distributed Active Archive Center (PODAAC). Land uses were obtained from the Spanish SIOSE facility (Sistema de Informacion de Ocupacion del Suelo en España) that uses a scale of 1:25.000 and polygon geometrical structure layer. SIOSE is based on panchromatic and multispectral 2.5 m resolution SPOT-5 images together with Landsat-5 images and orthophotos from the Spanish Nacional Plan of Aerial Orthophotography (PNOA). TVDI values were obtained from MODIS operational products of land surface <span class="hlt">temperature</span> and NDVI. SMOS ascending TBs were compared to inner-beam Aquarius descending half-orbit TBs coinciding over the study area at 06:00 h. The Aquarius inner beam has an incidence angle of 28,7º and SMOS data were considered for the 27,5º incidence angle. The SMOS products corresponded to version 2.6x (data before 31st Oct 2013) and version 2.7x (data after 1st Jan 2014). Intersections between both footprints were analysed under conditions of similar areas, land uses and TVDI values. For the latter (land uses/TVDI), a linear combination of SMOS land uses/TVDI was obtained to match the larger Aquarius footprint. A more physical approach is also under way</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> <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://hdl.handle.net/2060/20110015365','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110015365"><span>Seasonal Changes in Titan's Surface <span class="hlt">Temperatures</span></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 surface <span class="hlt">brightness</span> <span class="hlt">temperatures</span> have been <span class="hlt">observed</span> by Cassini in the thermal infrared. The Composite Infrared Spectrometer (CIRS) measured surface 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> <span class="hlt">temperatures</span> corrected for effects of the atmosphere. Between late northern winter and northern spring equinox a shift occurred in the <span class="hlt">temperature</span> 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 surface <span class="hlt">temperatures</span> 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 surface <span class="hlt">temperature</span> distribution, consistent with the post-equinox results of Voyager 1 as well as with predictions from general circulation modeling. A slightly elevated <span class="hlt">temperature</span> 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://adsabs.harvard.edu/abs/2017ESSD....9..293B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ESSD....9..293B"><span>The global SMOS Level 3 daily soil moisture and <span class="hlt">brightness</span> <span class="hlt">temperature</span> maps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bitar, Ahmad Al; Mialon, Arnaud; Kerr, Yann H.; Cabot, François; Richaume, Philippe; Jacquette, Elsa; Quesney, Arnaud; Mahmoodi, Ali; Tarot, Stéphane; Parrens, Marie; Al-Yaari, Amen; Pellarin, Thierry; Rodriguez-Fernandez, Nemesio; Wigneron, Jean-Pierre</p> <p>2017-06-01</p> <p>The objective of this paper is to present the multi-orbit (MO) surface soil moisture (SM) and angle-binned <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) products for the SMOS (Soil Moisture and Ocean Salinity) mission based on a new multi-orbit algorithm. The Level 3 algorithm at CATDS (Centre Aval de Traitement des Données SMOS) makes use of MO retrieval to enhance the robustness and quality of SM retrievals. The motivation of the approach is to make use of the longer temporal autocorrelation length of the vegetation optical depth (VOD) compared to the corresponding SM autocorrelation in order to enhance the retrievals when an acquisition occurs at the border of the swath. The retrieval algorithm is implemented in a unique operational processor delivering multiple parameters (e.g. SM and VOD) using multi-angular dual-polarisation TB from MO. A subsidiary angle-binned TB product is provided. In this study the Level 3 TB V310 product is showcased and compared to SMAP (Soil Moisture Active Passive) TB. The Level 3 SM V300 product is compared to the single-orbit (SO) retrievals from the Level 2 SM processor from ESA with aligned configuration. The advantages and drawbacks of the Level 3 SM product (L3SM) are discussed. The comparison is done on a global scale between the two datasets and on the local scale with respect to in situ data from AMMA-CATCH and USDA ARS Watershed networks. The results obtained from the global analysis show that the MO implementation enhances the number of retrievals: up to 9 % over certain areas. The comparison with the in situ data shows that the increase in the number of retrievals does not come with a decrease in quality, but rather at the expense of an increased time lag in product availability from 6 h to 3.5 days, which can be a limiting factor for applications like flood forecast but reasonable for drought monitoring and climate change studies. The SMOS L3 soil moisture and L3 <span class="hlt">brightness</span> <span class="hlt">temperature</span> products are delivered using an open licence and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008ApJ...687..936K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008ApJ...687..936K"><span>Extracting Galaxy Cluster Gas Inhomogeneity from X-Ray Surface <span class="hlt">Brightness</span>: A Statistical Approach and Application to Abell 3667</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kawahara, Hajime; Reese, Erik D.; Kitayama, Tetsu; Sasaki, Shin; Suto, Yasushi</p> <p>2008-11-01</p> <p>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 <span class="hlt">observations</span> 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 <span class="hlt">brightness</span>. We first create a set of synthetic clusters with lognormal fluctuations around their mean profile given by spherical isothermal β-models, later considering polytropic <span class="hlt">temperature</span> profiles as well. Performing mock <span class="hlt">observations</span> of these synthetic clusters, we find that the resulting X-ray surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span>. We analyze Chandra <span class="hlt">observations</span> of the galaxy cluster Abell 3667, and find that its X-ray surface <span class="hlt">brightness</span> fluctuations follow the lognormal distribution. While the lognormal model was originally motivated by cosmological hydrodynamic simulations, this is the first <span class="hlt">observational</span> 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 <span class="hlt">brightness</span> alone. Our current methodology combined with existing <span class="hlt">observational</span> data is useful in describing and inferring the</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 surface, a relationship of <span class="hlt">brightness</span> <span class="hlt">temperature</span> 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> <span class="hlt">temperature</span> may occur. This study examines the extent to which the relationships of thickness with <span class="hlt">brightness</span> <span class="hlt">temperature</span> (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> <span class="hlt">temperature</span> data from an over-flying aircraft. The results show that the <span class="hlt">brightness</span> <span class="hlt">temperature</span> 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 surface from a field experiment, which qualitatively support the idea that changes in the near-surface brine characteristics contribute to the <span class="hlt">observed</span> thickness-<span class="hlt">brightness</span> <span class="hlt">temperature</span>/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/2016AGUFM.P24A..02B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.P24A..02B"><span>Preliminary Analysis of Chang'E-2 Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Maps of the Moon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Blewett, D. T.; Zheng, Y. C.; Chan, K. L.; Neish, C.; Tsang, K. T.; Zhu, Y. C.; Jozwiak, L.</p> <p>2016-12-01</p> <p>China's Chang'E-2 (CE-2) lunar orbiter carried a microwave radiometer (MRM) that conducted passive remote sensing of the Moon at 3, 7.8, 19.35 and 37 GHz during 2010-2011. Earlier, the Chang'E-1 MRM obtained lower spatial resolution microwave data from a 200-km orbit, higher than CE-2's 100-km orbit. The MRM datasets represent a unique set of measurements of a type that have not been conducted by any previous lunar missions. Thermal emission of the lunar surface was measured and calibrated to <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB). Spherical harmonics fits were then used to model the TB variation as functions of local time and latitude for each of the four channels. Using the spherical harmonics fits, the day- and nighttime TB maps measured at various local times were normalized to noon-time and midnight conditions. The resulting eight MRM TB maps provide key information on the surface and near-subsurface structure and thermophysical properties of the lunar regolith; this information is complementary to that derived from LRO Diviner <span class="hlt">observations</span> in the infrared. We have <span class="hlt">observed</span> many thermal anomalies on the Moon, i.e., hot regions in the daytime map and cold spots in the nighttime map. We find that the high-Ti maria are heated in the day and cool in the night much more quickly than the other maria, attributable to the greater abundance of ilmenite (which has higher dielectric loss tangent than silicate minerals) in the high-Ti basalts. We note interesting contrasts in thermal behavior among high-reflectance, rayed craters. For example, the high-reflectance rays of Tycho are cooler than the surroundings in the 3 GHz daytime and nighttime maps, while the prominent rays of some other craters like Giordano Bruno are not distinctive in the 3 GHz maps. These differences can be understood in terms of variations in composition, structure, and thermophysical properties of the ray materials.</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('http://adsabs.harvard.edu/abs/2014AGUFMGC51D0442G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMGC51D0442G"><span>Inter-annual variation of the surface <span class="hlt">temperature</span> 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 surface <span class="hlt">temperatures</span> (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 <span class="hlt">temperatures</span> 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> <span class="hlt">temperatures</span> 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/2018BSRSL..87..365M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018BSRSL..87..365M"><span>Giant Low Surface <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>Mishra, Alka; Kantharia, Nimisha G.; Das, Mousumi</p> <p>2018-04-01</p> <p>In this paper, we present radio <span class="hlt">observations</span> of the giant low surface <span class="hlt">brightness</span> (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 <span class="hlt">brightness</span> 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.</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('https://ntrs.nasa.gov/search.jsp?R=20160013720&hterms=Physical&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DPhysical','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160013720&hterms=Physical&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DPhysical"><span>Physical Models of Layered Polar Firn <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> from 0.5 to 2 GHz</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tan, Shurun; Aksoy, Mustafa; Brogioni, Marco; Macelloni, Giovanni; Durand, Michael; Jezek, Kenneth C.; Wang, Tian-Lin; Tsang, Leung; Johnson, Joel T.; Drinkwater, Mark R.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20160013720'); toggleEditAbsImage('author_20160013720_show'); toggleEditAbsImage('author_20160013720_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20160013720_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20160013720_hide"></p> <p>2015-01-01</p> <p>We investigate physical effects influencing 0.5-2 GHz <span class="hlt">brightness</span> <span class="hlt">temperatures</span> of layered polar firn to support the Ultra Wide Band Software Defined Radiometer (UWBRAD) experiment to be conducted in Greenland and in Antarctica. We find that because ice particle grain sizes are very small compared to the 0.5-2 GHz wavelengths, volume scattering effects are small. Variations in firn density over cm- to m-length scales, however, cause significant effects. Both incoherent and coherent models are used to examine these effects. Incoherent models include a 'cloud model' that neglects any reflections internal to the ice sheet, and the DMRT-ML and MEMLS radiative transfer codes that are publicly available. The coherent model is based on the layered medium implementation of the fluctuation dissipation theorem for thermal microwave radiation from a medium having a nonuniform <span class="hlt">temperature</span>. Density profiles are modeled using a stochastic approach, and model predictions are averaged over a large number of realizations to take into account an averaging over the radiometer footprint. Density profiles are described by combining a smooth average density profile with a spatially correlated random process to model density fluctuations. It is shown that coherent model results after ensemble averaging depend on the correlation lengths of the vertical density fluctuations. If the correlation length is moderate or long compared with the wavelength (approximately 0.6x longer or greater for Gaussian correlation function without regard for layer thinning due to compaction), coherent and incoherent model results are similar (within approximately 1 K). However, when the correlation length is short compared to the wavelength, coherent model results are significantly different from the incoherent model by several tens of kelvins. For a 10-cm correlation length, the differences are significant between 0.5 and 1.1 GHz, and less for 1.1-2 GHz. Model results are shown to be able to match the v</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA00427&hterms=Eurasia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DEurasia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA00427&hterms=Eurasia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DEurasia"><span>Global Average <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> for April 2003</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2003-01-01</p> <p><p/> [figure removed for brevity, see original site] Figure 1 <p/> This image shows average <span class="hlt">temperatures</span> in April, 2003, <span class="hlt">observed</span> by AIRS at an infrared wavelength that senses either the Earth's surface or any intervening cloud. Similar to a photograph of the planet taken with the camera shutter held open for a month, stationary features are captured while those obscured by moving clouds are blurred. Many continental features stand out boldly, such as our planet's vast deserts, and India, now at the end of its long, clear dry season. Also obvious are the high, cold Tibetan plateau to the north of India, and the mountains of North America. The band of yellow encircling the planet's equator is the Intertropical Convergence Zone (ITCZ), a region of persistent thunderstorms and associated high, cold clouds. The ITCZ merges with the monsoon systems of Africa and South America. Higher latitudes are increasingly obscured by clouds, though some features like the Great Lakes, the British Isles and Korea are apparent. The highest latitudes of Europe and Eurasia are completely obscured by clouds, while Antarctica stands out cold and clear at the bottom of the image. <p/> The Atmospheric Infrared Sounder Experiment, with its visible, infrared, and microwave detectors, provides a three-dimensional look at Earth's weather. Working in tandem, the three instruments can make simultaneous <span class="hlt">observations</span> all the way down to the Earth's surface, even in the presence of heavy clouds. With more than 2,000 channels sensing different regions of the atmosphere, the system creates a global, 3-D map of atmospheric <span class="hlt">temperature</span> and humidity and provides information on clouds, greenhouse gases, and many other atmospheric phenomena. The AIRS Infrared Sounder Experiment flies onboard NASA's Aqua spacecraft and is managed by NASA's Jet Propulsion Laboratory, Pasadena, Calif., under contract to NASA. JPL is a division of the California Institute of Technology in Pasadena.</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://adsabs.harvard.edu/abs/2017EGUGA..1914320G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1914320G"><span>Modelling of the L-band <span class="hlt">brightness</span> <span class="hlt">temperatures</span> measured with ELBARA III radiometer on Bubnow wetland</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gluba, Lukasz; Sagan, Joanna; Lukowski, Mateusz; Szlazak, Radoslaw; Usowicz, Boguslaw</p> <p>2017-04-01</p> <p>Microwave radiometry has become the main tool for investigating soil moisture (SM) with remote sensing methods. ESA - SMOS (Soil Moisture and Ocean Salinity) satellite operating at L-band provides global distribution of soil moisture. An integral part of SMOS mission are calibration and validation activities involving measurements with ELBARA III which is an L-band microwave passive radiometer. It is done in order to improve soil moisture retrievals - make them more time-effective and accurate. The instrument is located at Bubnow test-site, on the border of cultivated field, fallow, meadow and natural wetland being a part of Polesie National Park (Poland). We obtain both temporal and spatial dependences of <span class="hlt">brightness</span> <span class="hlt">temperatures</span> for varied types of land covers with the ELBARA III directed at different azimuths. Soil moisture is retrieved from <span class="hlt">brightness</span> <span class="hlt">temperature</span> using L-band Microwave Emission of the Biosphere (L-MEB) model, the same as currently used radiative transfer model for SMOS. Parametrization of L-MEB, as well as input values are still under debate. We discuss the results of SM retrievals basing on data obtained during first year of the radiometer's operation. We analyze temporal dependences of retrieved SM for one-parameter (SM), two-parameter (SM, τ - optical depth) and three-parameter (SM, τ, Hr - roughness parameter) retrievals, as well as spatial dependences for specific dates. Special case of Simplified Roughness Parametrization, combining the roughness parameter and optical depth, is considered. L-MEB processing is supported by the continuous measurements of soil moisture and <span class="hlt">temperature</span> obtained from nearby agrometeorological station, as well as studies on the soil granulometric composition of the Bubnow test-site area. Furthermore, for better estimation of optical depth, the satellite-derived Normalized Difference Vegetation Index (NDVI) was employed, supported by measured in situ vegetation parameters (such as Leaf Area Index and Vegetation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24528115','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24528115"><span>The seasonal variation in skin hydration, sebum, scaliness, <span class="hlt">brightness</span> and elasticity in Korean females.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nam, G W; Baek, J H; Koh, J S; Hwang, J-K</p> <p>2015-02-01</p> <p>Age, gender, regional, and ethnic differences influence skin conditions. The purpose of this study was to <span class="hlt">observe</span> the effects of environments, especially the air <span class="hlt">temperature</span>, relative humidity, air pressure, duration of sunshine, and precipitation on skin, and the seasonal variation in skin hydration, sebum, scales, <span class="hlt">brightness</span>, and elasticity in Korean females. The study included 89 Korean subjects, aged 29.7 ± 6.2 years. The five skin biophysical parameters (skin hydration, sebum, scales, <span class="hlt">brightness</span>, and elasticity) were measured at six sites: forehead, under the eye, frontal cheek, crow's foot, lateral cheek, and inner forearm. Skin hydration was measured using the Corneometer® CM 825. Skin sebum was measured with Sebumeter® SM 815. Skin scaliness was measured with Visioscan® VC 98. Skin <span class="hlt">brightness</span> (L* value) was measured by using Spectrophotometer. A suction chamber device, Cutometer® MPA 580, was used to measure the skin elasticity. The measurements were performed every month for 13 months, from April 2007 to April 2008. There were significantly seasonal variations in environmental factors. The air <span class="hlt">temperature</span> was the lowest in January (-1.7°C), and the highest in August (26.5°C). The relative humidity was the lowest in February (46%), and the highest in July and August (75%). There was a negative correlation between skin scaliness and three environmental factors such as air <span class="hlt">temperature</span>, relative humidity, and highest precipitation. There was a positive correlation between skin scaliness and two environmental factors such as air pressure and duration of sunshine. Elasticity was correlated with air <span class="hlt">temperature</span> positively and with air pressure negatively. The correlations shown between the skin biophysical parameters and environmental factors demonstrate that the skin biophysical parameters are affected by environmental factors. © 2014 John Wiley & Sons A/S. Published by John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19770040941&hterms=Krieger&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DKrieger','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19770040941&hterms=Krieger&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DKrieger"><span><span class="hlt">Observation</span> of spatial and temporal variations in X-ray <span class="hlt">bright</span> point emergence patterns. [at solar surface</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Golub, L.; Krieger, A. S.; Vaiana, G. S.</p> <p>1976-01-01</p> <p><span class="hlt">Observations</span> of X-ray <span class="hlt">bright</span> 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 <span class="hlt">observed</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=316288','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=316288"><span>Inter-comparison of SMAP, Aquarius and SMOS L-band <span class="hlt">brightness</span> <span class="hlt">temperature</span> <span class="hlt">observations</span></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>Soil Moisture Active Passive (SMAP) mission is scheduled for launch on January 29, 2015. SMAP will make <span class="hlt">observations</span> with an L-band radar and radiometer using a shared 6 m rotating reflector antenna. SMAP is a fully polarimetric radiometer with the center frequency of 1.41 GHz. The target accuracy o...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997Icar..128..189K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997Icar..128..189K"><span><span class="hlt">Temperatures</span> and Altitudes of Jupiter's Ultraviolet Aurora Inferred from GHRS <span class="hlt">Observations</span> with the Hubble Space Telescope</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Y. H.; Fox, J. L.; Caldwell, John J.</p> <p>1997-07-01</p> <p>We <span class="hlt">observed</span> the jovian UV auroral regions with the Goddard high resolution spectrograph (GHRS) on board the Hubble Space Telescope (HST) on Apr. 29, May 2, and June 10, 1995. <span class="hlt">Observations</span> of target areas were made in pairs in the two wavelength ranges 1257-1293 Å and 1587-1621 Å. Spectra in the long wavelength range are dominated by emissions of the H2Lyman band system and show well separated rotational features, which we have used to determine the <span class="hlt">temperatures</span> of the auroral emission regions. Spectra in the short wavelength range are mostly due to emission in the H2Lyman and Werner band systems, but their intensities are reduced by hydrocarbon absorption. The brightest spectral pair was <span class="hlt">observed</span> toward an area with longitude 155° and jovicentric latitude 58° when the central meridian longitudes (CMLs) were 191° and 203°. This area was found to be <span class="hlt">bright</span> in our previous HST <span class="hlt">observations</span> in 1993 and in HST faint object camera images. Assuming that electron impact excitation is the major source of the jovian aurora, we estimate total emission rates in the Lyman band system of about 270 and 46 kR for the long and short wavelength spectra of the pair, respectively. The attenuation of emission rate in the short wavelength spectrum implies a methane column density of about 3 × 1016cm-2, and a <span class="hlt">temperature</span> of about 450 K is inferred from the long wavelength spectrum of the brightest pair. For all six pairs of <span class="hlt">observed</span> spectra, we estimate methane column densities in the range (1-7) × 1016cm-2, which, when compared to a standard mid-latitude model, corresponds to a pressure range from a few μbar to a few tens of μbar. The <span class="hlt">temperatures</span> derived are in the range 400-850 K with a possible tendency toward lower <span class="hlt">temperatures</span> for higher methane column densities. This tendency and the uncertainty in the <span class="hlt">temperatures</span> derived may indicate that the <span class="hlt">temperatures</span> increases rapidly with altitude around the methane homopause in the auroral regions.</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 Surface <span class="hlt">Temperatures</span> 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 surface 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> <span class="hlt">temperature</span> of the surface. 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 surface <span class="hlt">temperatures</span>. 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 surface <span class="hlt">brightness</span> <span class="hlt">temperature</span> to be 93.7+/-0.6 K, the same as the <span class="hlt">temperature</span> measured at the Huygens landing site. The surface <span class="hlt">brightness</span> <span class="hlt">temperature</span> decreases by 2 K toward the south pole and by 3 K toward the north pole. The drop in surface <span class="hlt">temperature</span> 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 surface and atmospheric <span class="hlt">temperatures</span> is emerging from CIRS measurements over the Cassini mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Natur.553..189B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Natur.553..189B"><span><span class="hlt">Bright</span> triplet excitons in caesium lead halide perovskites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Becker, Michael A.; Vaxenburg, Roman; Nedelcu, Georgian; Sercel, Peter C.; Shabaev, Andrew; Mehl, Michael J.; Michopoulos, John G.; Lambrakos, Samuel G.; Bernstein, Noam; Lyons, John L.; Stöferle, Thilo; Mahrt, Rainer F.; Kovalenko, Maksym V.; Norris, David J.; Rainò, Gabriele; Efros, Alexander L.</p> <p>2018-01-01</p> <p>Nanostructured semiconductors emit light from electronic states known as excitons. For organic materials, Hund’s rules state that the lowest-energy exciton is a poorly emitting triplet state. For inorganic semiconductors, similar rules predict an analogue of this triplet state known as the ‘dark exciton’. Because dark excitons release photons slowly, hindering emission from inorganic nanostructures, materials that disobey these rules have been sought. However, despite considerable experimental and theoretical efforts, no inorganic semiconductors have been identified in which the lowest exciton is <span class="hlt">bright</span>. Here we show that the lowest exciton in caesium lead halide perovskites (CsPbX3, with X = Cl, Br or I) involves a highly emissive triplet state. We first use an effective-mass model and group theory to demonstrate the possibility of such a state existing, which can occur when the strong spin-orbit coupling in the conduction band of a perovskite is combined with the Rashba effect. We then apply our model to CsPbX3 nanocrystals, and measure size- and composition-dependent fluorescence at the single-nanocrystal level. The <span class="hlt">bright</span> triplet character of the lowest exciton explains the anomalous photon-emission rates of these materials, which emit about 20 and 1,000 times faster than any other semiconductor nanocrystal at room and cryogenic <span class="hlt">temperatures</span>, respectively. The existence of this <span class="hlt">bright</span> triplet exciton is further confirmed by analysis of the fine structure in low-<span class="hlt">temperature</span> fluorescence spectra. For semiconductor nanocrystals, which are already used in lighting, lasers and displays, these excitons could lead to materials with brighter emission. More generally, our results provide criteria for identifying other semiconductors that exhibit <span class="hlt">bright</span> excitons, with potential implications for optoelectronic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ApJ...796...73H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ApJ...796...73H"><span>Coronal <span class="hlt">Bright</span> Points Associated with Minifilament Eruptions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hong, Junchao; Jiang, Yunchun; Yang, Jiayan; Bi, Yi; Li, Haidong; Yang, Bo; Yang, Dan</p> <p>2014-12-01</p> <p>Coronal <span class="hlt">bright</span> points (CBPs) are small-scale, long-lived coronal brightenings that always correspond to photospheric network magnetic features of opposite polarity. In this paper, we subjectively adopt 30 CBPs in a coronal hole to study their eruptive behavior using data from the Atmospheric Imaging Assembly (AIA) and the Helioseismic and Magnetic Imager (HMI) on board the Solar Dynamics Observatory. About one-quarter to one-third of the CBPs in the coronal hole go through one or more minifilament eruption(s) (MFE(s)) throughout their lifetimes. The MFEs occur in temporal association with the <span class="hlt">brightness</span> maxima of CBPs and possibly result from the convergence and cancellation of underlying magnetic dipoles. Two examples of CBPs with MFEs are analyzed in detail, where minifilaments appear as dark features of a cool channel that divide the CBPs along the neutral lines of the dipoles beneath. The MFEs show the typical rising movements of filaments and mass ejections with brightenings at CBPs, similar to large-scale filament eruptions. Via differential emission measure analysis, it is found that CBPs are heated dramatically by their MFEs and the ejected plasmas in the MFEs have average <span class="hlt">temperatures</span> close to the pre-eruption BP plasmas and electron densities typically near 109 cm-3. These new <span class="hlt">observational</span> results indicate that CBPs are more complex in dynamical evolution and magnetic structure than previously thought.</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/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/2016AGUFM.C13D0859R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C13D0859R"><span>Glacier Melt Detection in Complex Terrain Using New AMSR-E Calibrated Enhanced Daily EASE-Grid 2.0 <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> (CETB) Earth System Data Record</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramage, J. M.; Brodzik, M. J.; Hardman, M.</p> <p>2016-12-01</p> <p>Passive microwave (PM) 18 GHz and 36 GHz horizontally- and vertically-polarized <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (Tb) channels from the Advanced Microwave Scanning Radiometer for EOS (AMSR-E) have been important sources of information about snow melt status in glacial environments, particularly at high latitudes. PM data are sensitive to the changes in near-surface liquid water that accompany melt onset, melt intensification, and refreezing. Overpasses are frequent enough that in most areas multiple (2-8) <span class="hlt">observations</span> per day are possible, yielding the potential for determining the dynamic state of the snow pack during transition seasons. AMSR-E Tb data have been used effectively to determine melt onset and melt intensification using daily Tb and diurnal amplitude variation (DAV) thresholds. Due to mixed pixels in historically coarse spatial resolution Tb data, melt analysis has been impractical in ice-marginal zones where pixels may be only fractionally snow/ice covered, and in areas where the glacier is near large bodies of water: even small regions of open water in a pixel severely impact the microwave signal. We use the new enhanced-resolution Calibrated Passive Microwave Daily EASE-Grid 2.0 <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> (CETB) Earth System Data Record product's twice daily obserations to test and update existing snow melt algorithms by determining appropriate melt thresholds for both Tb and DAV for the CETB 18 and 36 GHz channels. We use the enhanced resolution data to evaluate melt characteristics along glacier margins and melt transition zones during the melt seasons in locations spanning a wide range of melt scenarios, including the Patagonian Andes, the Alaskan Coast Range, and the Russian High Arctic icecaps. We quantify how improvement of spatial resolution from the original 12.5 - 25 km-scale pixels to the enhanced resolution of 3.125 - 6.25 km improves the ability to evaluate melt timing across boundaries and transition zones in diverse glacial environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...854L..29K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...854L..29K"><span>Solar Radio Burst Associated with the Falling <span class="hlt">Bright</span> EUV Blob</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Karlický, Marian; Zemanová, Alena; Dudík, Jaroslav; Radziszewski, Krzysztof</p> <p>2018-02-01</p> <p>At the beginning of the 2015 November 4 flare, in the 1300–2000 MHz frequency range, we <span class="hlt">observed</span> a very rare slow positively drifting burst. We searched for associated phenomena in simultaneous EUV <span class="hlt">observations</span> made by IRIS, SDO/AIA, and Hinode/XRT, as well as in H α <span class="hlt">observations</span>. We found that this radio burst was accompanied with the <span class="hlt">bright</span> blob, visible at transition region, coronal, and flare <span class="hlt">temperatures</span>, falling down to the chromosphere along the dark loop with a velocity of about 280 km s‑1. The dark loop was visible in H α but disappeared afterward. Furthermore, we found that the falling blob interacted with the chromosphere as expressed by a sudden change of the H α spectra at the location of this interaction. Considering different possibilities, we propose that the <span class="hlt">observed</span> slow positively drifting burst is generated by the thermal conduction front formed in front of the falling hot EUV blob.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9910E..1SK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9910E..1SK"><span>Through thick and thin: quantitative classification of photometric <span class="hlt">observing</span> conditions on Paranal</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kerber, Florian; Querel, Richard R.; Neureiter, Bianca; Hanuschik, Reinhard</p> <p>2016-07-01</p> <p>A Low Humidity and <span class="hlt">Temperature</span> Profiling (LHATPRO) microwave radiometer is used to monitor sky conditions over ESO's Paranal observatory. It provides measurements of precipitable water vapour (PWV) at 183 GHz, which are being used in Service Mode for scheduling <span class="hlt">observations</span> that can take advantage of favourable conditions for infrared (IR) <span class="hlt">observations</span>. The instrument also contains an IR camera measuring sky <span class="hlt">brightness</span> <span class="hlt">temperature</span> at 10.5 μm. It is capable of detecting cold and thin, even sub-visual, cirrus clouds. We present a diagnostic diagram that, based on a sophisticated time series analysis of these IR sky <span class="hlt">brightness</span> data, allows for the automatic and quantitative classification of photometric <span class="hlt">observing</span> conditions over Paranal. The method is highly sensitive to the presence of even very thin clouds but robust against other causes of sky <span class="hlt">brightness</span> variations. The diagram has been validated across the complete range of conditions that occur over Paranal and we find that the automated process provides correct classification at the 95% level. We plan to develop our method into an operational tool for routine use in support of ESO Science Operations.</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://hdl.handle.net/2060/20160002400','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160002400"><span><span class="hlt">Bright</span> Stuff on Ceres = Sulfates and Carbonates on CI Chondrites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zolensky, Michael; Chan, Queenie H. S.; Gounelle, Matthieu; Fries, Marc</p> <p>2016-01-01</p> <p>Recent reports of the DAWN spacecraft's <span class="hlt">observations</span> of the surface of Ceres indicate that there are <span class="hlt">bright</span> 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 <span class="hlt">bright</span> 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 <span class="hlt">observations</span> of chondritic materials in the lab shed light on the nature of the <span class="hlt">bright</span> spots on Ceres</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('http://www.ars.usda.gov/research/publications/publication/?seqNo115=252606','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=252606"><span>Passive Polarimetric Microwave Signatures <span class="hlt">Observed</span> Over Antarctica</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>WindSat satellite-based fully polarimetric passive microwave <span class="hlt">observations</span>, expressed in the form of the Stokes vector, were analyzed over the Antarctic ice sheet. The vertically and horizontally polarized <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (first two Stokes components) from WindSat are shown to be consistent w...</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('https://www.ncbi.nlm.nih.gov/pubmed/16171276','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16171276"><span>The treatment of early-morning awakening insomnia with 2 evenings of <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>Lack, Leon; Wright, Helen; Kemp, Kristyn; Gibbon, Samantha</p> <p>2005-05-01</p> <p>To assess the effectiveness of brief <span class="hlt">bright</span>-light therapy for the treatment of early-morning awakening insomnia. Twenty-four healthy adults with early-morning awakening insomnia were assigned to either the <span class="hlt">bright</span>-light condition (2,500-lux white light) or the control (dim red light) condition. The circadian phase of rectal <span class="hlt">temperature</span> and urinary melatonin rhythms were assessed with 26-hour constant routines before and after 2 evenings of light therapy. Sleep and daytime functioning were monitored using sleep diaries, activity monitors, and mood scales before light therapy and for 4 weeks during the follow-up period. While there were no significant circadian phase changes in the dim-light control group, the <span class="hlt">bright</span>-light group had significant 2-hour phase delays of circadian <span class="hlt">temperature</span> and melatonin rhythm. Compared to pretreatment measures, over the 4-week follow-up period, the <span class="hlt">bright</span>-light group had a greater reduction of time awake after sleep onset, showed a trend toward waking later, and had a greater increase of total sleep time. Participants in the <span class="hlt">bright</span>-light condition also tended to report greater reductions of negative daytime symptoms, including significantly fewer days of feeling depressed at the 4-week follow-up, as compared with the control group. Two evenings of <span class="hlt">bright</span>-light exposure phase delayed the circadian rhythms of early-morning awakening insomniacs. It also improved diary and actigraphy sleep measures and improved some indexes of daytime functioning for up to 1 month after light exposure. The study suggests that a brief course of evening <span class="hlt">bright</span>-light therapy can be an effective treatment for early-morning awakening insomniacs who have relatively phase advanced circadian rhythms.</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://hdl.handle.net/2060/19740007902','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19740007902"><span>Detection of moisture and moisture related phenomena from Skylab. [correlation of <span class="hlt">brightness</span> and antenna <span class="hlt">temperature</span> with soil moisture for Texas and Kansas test sites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Eagleman, J. R.; Pogge, E. C.; Moore, R. K. (Principal Investigator); Hardy, N.; Lin, W.; League, L.</p> <p>1974-01-01</p> <p>The author has identified the following significant results. Skylab 2 data for June 5, 1973 (Texas site) relates favorably with previously calculated aircraft data when correlating <span class="hlt">brightness</span> <span class="hlt">temperature</span> to soil moisture. However, more detailed work is needed to determine the corrected surface <span class="hlt">temperature</span>. In addition, correlations between the S194 antenna <span class="hlt">temperature</span> and soil moisture have been obtained for five sets of Skylab data. The best correlations were obtained for the surface to one inch depth in four cases and for surface to two inches depth for the fifth case. Correlation coefficients for the surface to one inch depth were -0.98, -0.95, -0.90, -0.82, and -0.80.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1913703L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1913703L"><span>Microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> and thermal inertia - towards synergistic method of high-resolution soil moisture retrieval</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lukowski, Mateusz; Usowicz, Boguslaw; Sagan, Joanna; Szlazak, Radoslaw; Gluba, Lukasz; Rojek, Edyta</p> <p>2017-04-01</p> <p>Soil moisture is an important parameter in many environmental studies, as it influences the exchange of water and energy at the interface between the land surface and the atmosphere. Accurate assessment of the soil moisture spatial and temporal variations is crucial for numerous studies; starting from a small scale of single field, then catchment, mesoscale basin, ocean conglomeration, finally ending at the global water cycle. Despite numerous advantages, such as fine accuracy (undisturbed by clouds or daytime conditions) and good temporal resolution, passive microwave remote sensing of soil moisture, e.g. SMOS and SMAP, are not applicable to a small scale - simply because of too coarse spatial resolution. On the contrary, thermal infrared-based methods of soil moisture retrieval have a good spatial resolution, but are often disturbed by clouds and vegetation interferences or night effects. The methods that base on point measurements, collected in situ by monitoring stations or during field campaigns, are sometimes called "ground truth" and may serve as a reference for remote sensing, of course after some up-scaling and approximation procedures that are, unfortunately, potential source of error. Presented research concern attempt to synergistic approach that join two remote sensing methods: passive microwave and thermal infrared, supported by in situ measurements. Microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> of soil was measured by ELBARA, the radiometer at 1.4 GHz frequency, installed at 6 meters high tower at Bubnow test site in Poland. Thermal inertia around the tower was modelled using the statistical-physical model whose inputs were: soil physical properties, its water content, albedo and surface <span class="hlt">temperatures</span> measured by an infrared pyrometer, directed at the same footprint as ELBARA. The results coming from this method were compared to in situ data obtained during several field campaigns and by the stationary agrometeorological stations. The approach seems to be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20120010218&hterms=Butterfly&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DButterfly','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20120010218&hterms=Butterfly&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DButterfly"><span>Behavior of Solar Cycles 23 and 24 Revealed by 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>Gopalswamy, N.; Yashiro, S.; Maekelae, P.; Michalek, G.; Shibasaki, K.; Hathaway, D. H.</p> <p>2012-01-01</p> <p>Using magnetic and microwave butterfly diagrams, we compare the behavior of solar polar regions to show that (1) the polar magnetic field and the microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> during solar minimum substantially diminished during the cycle 23/24 minimum compared to the 22/23 minimum. (2) The polar microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> (Tb) seems to be a good proxy for the underlying magnetic field strength (B). The analysis indicates a relationship, B = 0.0067Tb - 70, where B is in G and Tb in K. (3) Both the <span class="hlt">brightness</span> <span class="hlt">temperature</span> and the magnetic field strength show north-south asymmetry most of the time except for a short period during the maximum phase. (4) The rush-to-the-pole phenomenon <span class="hlt">observed</span> in the prominence eruption (PE) activity seems to be complete in the northern hemisphere as of 2012 March. (5) The decline of the microwave <span class="hlt">brightness</span> <span class="hlt">temperature</span> in the north polar region to the quiet-Sun levels and the sustained PE activity poleward of 60degN suggest that solar maximum conditions have arrived at the northern hemisphere. The southern hemisphere continues to exhibit conditions corresponding to the rise phase of solar cycle 24. Key words: Sun: chromosphere Sun: coronal mass ejections (CMEs) Sun: filaments, prominences Sun: photosphere Sun: radio radiation Sun: surface magnetism</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080034522&hterms=poster&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dposter','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080034522&hterms=poster&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dposter"><span>Ion <span class="hlt">Temperature</span> Control of the Io Plasma Torus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Delamere, P. A.; Schneider, N. M.; Steffl, A. J.; Robbins, S. J.</p> <p>2005-01-01</p> <p>We report on <span class="hlt">observational</span> and theoretical studies of ion <span class="hlt">temperature</span> in the Io plasma torus. Ion <span class="hlt">temperature</span> is a critical factor for two reasons. First, ions are a major supplier of energy to the torus electrons which power the intense EUV emissions. Second, ion <span class="hlt">temperature</span> determines the vertical extent of plasma along field lines. Higher <span class="hlt">temperatures</span> spread plasma out, lowers the density and slows reaction rates. The combined effects can play a controlling role in torus energetics and chemistry. An unexpected tool for the study of ion <span class="hlt">temperature</span> is the longitudinal structure in the plasma torus which often manifests itself as periodic <span class="hlt">brightness</span> variations. Opposite sides of the torus (especially magnetic longitudes 20 and 200 degrees) have been <span class="hlt">observed</span> on numerous occasions to have dramatically different <span class="hlt">brightness</span>, density, composition, ionization state, electron <span class="hlt">temperature</span> and ion <span class="hlt">temperature</span>. These asymmetries must ultimately be driven by different energy flows on the opposite sides, presenting an opportunity to <span class="hlt">observe</span> key torus processes operating under different conditions. The most comprehensive dataset for the study of longitudinal variations was obtained by the Cassini UVIS instrument during its Jupiter flyby. Steffl (Ph.D. thesis, 2005) identified longitudinal variations in all the quantities listed above wit the exception of ion <span class="hlt">temperature</span>. We extend his work by undertaking the first search for such variation in the UVIS dataset. We also report on a 'square centimeter' model of the torus which extend the traditional 'cubic centimeter' models by including the controlling effects of ion <span class="hlt">temperature</span> more completely.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140012069','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140012069"><span>Optimization of a Radiative Transfer Forward Operator for Simulating SMOS <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> over the Upper Mississippi Basin, USA</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lievens, H.; Verhoest, N. E. C.; Martens, B.; VanDenBerg, M. J.; Bitar, A. Al; Tomer, S. Kumar; Merlin, O.; Cabot, F.; Kerr, Y.; DeLannoy, G. J. M.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20140012069'); toggleEditAbsImage('author_20140012069_show'); toggleEditAbsImage('author_20140012069_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20140012069_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20140012069_hide"></p> <p>2014-01-01</p> <p>The Soil Moisture and Ocean Salinity (SMOS) satellite mission is routinely providing global multi-angular <span class="hlt">observations</span> of <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) at both horizontal and vertical polarization with a 3-day repeat period. The assimilation of such data into a land surface model (LSM) may improve the skill of operational flood forecasts through an improved estimation of soil moisture (SM). To accommodate for the direct assimilation of the SMOS TB data, the LSM needs to be coupled with a radiative transfer model (RTM), serving as a forward operator for the simulation of multi-angular and multi-polarization top of atmosphere TBs. This study investigates the use of the Variable Infiltration Capacity (VIC) LSM coupled with the Community Microwave Emission Modelling platform (CMEM) for simulating SMOS TB <span class="hlt">observations</span> over the Upper Mississippi basin, USA. For a period of 2 years (2010-2011), a comparison between SMOS TBs and simulations with literature-based RTM parameters reveals a basin averaged bias of 30K. Therefore, time series of SMOS TB <span class="hlt">observations</span> are used to investigate ways for mitigating these large biases. Specifically, the study demonstrates the impact of the LSM soil moisture climatology in the magnitude of TB biases. After CDF matching the SM climatology of the LSM to SMOS retrievals, the average bias decreases from 30K to less than 5K. Further improvements can be made through calibration of RTM parameters related to the modeling of surface roughness and vegetation. Consequently, it can be concluded that SM rescaling and RTM optimization are efficient means for mitigating biases and form a necessary preparatory step for data assimilation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840018164','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840018164"><span>Satellite <span class="hlt">observations</span> of a monsoon depression</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Warner, C.</p> <p>1984-01-01</p> <p>The exploration of a monsoon depression over Burma and the Bay of Bengal is discussed. Aircraft and satellite data were examined, with an emphasis on the Microwave Sounding Unit (MSU) aboard TIROS-N and the Scanning Multichannel Microwave Radiometer (SMMR) aboard Nimbus-7. The structure of the monsoon depression was found to be dominated by cumulus convection. The only systematic large scale behavior discerned was a propagation of the depression westward, and diurnal migration of contours of <span class="hlt">brightness</span> <span class="hlt">temperature</span>. These contours in the middle troposphere showed a gradient toward the north with the patterns migrating northward at night. From SMMR and dropwindsonde data, water vapor contents were found to be near 65 mm, increasing to more than 70 mm in the northeast Bay of Bengal. Cloud water contents reached about three mm. Rainfall rates exceeding 5.7 mm/h occurred over a small part of the storm area, while mean rainfall rates in areas of order 20,000 sq km reached approximately 0.5 mm/h. Measured MSU <span class="hlt">brightness</span> <span class="hlt">temperatures</span> were reconciled very well with dropwindsonde data and with airborne in situ <span class="hlt">observations</span> of clouds (by photography) and hydrometeors (by radar). Diffuse scattering was determined to be important in computing <span class="hlt">brightness</span> <span class="hlt">temperature</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19790050227&hterms=Krieger&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DKrieger','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19790050227&hterms=Krieger&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DKrieger"><span>Anticorrelation of X-ray <span class="hlt">bright</span> points with sunspot number, 1970-1978</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Golub, L.; Davis, J. M.; Krieger, A. S.</p> <p>1979-01-01</p> <p>Soft X-ray <span class="hlt">observations</span> of the solar corona over the period 1970-1978 show that the number of small short-lived bipolar magnetic features (X-ray <span class="hlt">bright</span> 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 <span class="hlt">bright</span> points. In 1970, near the peak of solar cycle 20, the contributions from <span class="hlt">bright</span> points and from active regions appear to be approximately equal. These <span class="hlt">observations</span> 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.</p> </li> <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('https://www.osti.gov/biblio/22370183-coronal-bright-points-associated-minifilament-eruptions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22370183-coronal-bright-points-associated-minifilament-eruptions"><span>Coronal <span class="hlt">bright</span> points associated with minifilament eruptions</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>Hong, Junchao; Jiang, Yunchun; Yang, Jiayan</p> <p>2014-12-01</p> <p>Coronal <span class="hlt">bright</span> points (CBPs) are small-scale, long-lived coronal brightenings that always correspond to photospheric network magnetic features of opposite polarity. In this paper, we subjectively adopt 30 CBPs in a coronal hole to study their eruptive behavior using data from the Atmospheric Imaging Assembly (AIA) and the Helioseismic and Magnetic Imager (HMI) on board the Solar Dynamics Observatory. About one-quarter to one-third of the CBPs in the coronal hole go through one or more minifilament eruption(s) (MFE(s)) throughout their lifetimes. The MFEs occur in temporal association with the <span class="hlt">brightness</span> maxima of CBPs and possibly result from the convergence and cancellationmore » of underlying magnetic dipoles. Two examples of CBPs with MFEs are analyzed in detail, where minifilaments appear as dark features of a cool channel that divide the CBPs along the neutral lines of the dipoles beneath. The MFEs show the typical rising movements of filaments and mass ejections with brightenings at CBPs, similar to large-scale filament eruptions. Via differential emission measure analysis, it is found that CBPs are heated dramatically by their MFEs and the ejected plasmas in the MFEs have average <span class="hlt">temperatures</span> close to the pre-eruption BP plasmas and electron densities typically near 10{sup 9} cm{sup –3}. These new <span class="hlt">observational</span> results indicate that CBPs are more complex in dynamical evolution and magnetic structure than previously thought.« less</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('http://adsabs.harvard.edu/abs/2011ApJ...732L..25C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApJ...732L..25C"><span><span class="hlt">Observing</span> Flux Rope Formation During the Impulsive Phase of a Solar Eruption</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.; Zhang, J.; Liu, Y.; Ding, M. D.</p> <p>2011-05-01</p> <p>Magnetic flux ropes are believed to be an important structural component of coronal mass ejections (CMEs). While there exists much <span class="hlt">observational</span> evidence of flux ropes after the eruption, e.g., as seen in remote-sensing coronagraph images or in situ solar wind data, the direct <span class="hlt">observation</span> of flux ropes during CME impulsive phase has been rare. In this Letter, we present an unambiguous <span class="hlt">observation</span> of a flux rope still in the formation phase in the low corona. The CME of interest occurred above the east limb on 2010 November 3 with footpoints partially blocked. The flux rope was seen as a <span class="hlt">bright</span> blob of hot plasma in the Atmospheric Imaging Assembly (AIA) 131 Å passband (peak <span class="hlt">temperature</span> ~11 MK) rising from the core of the source active region, rapidly moving outward and stretching the surrounding background magnetic field upward. The stretched magnetic field seemed to curve-in behind the core, similar to the classical magnetic reconnection scenario in eruptive flares. On the other hand, the flux rope appeared as a dark cavity in the AIA 211 Å passband (2.0 MK) and 171 Å passband (0.6 MK) in these relatively cool <span class="hlt">temperature</span> bands, a <span class="hlt">bright</span> rim clearly enclosed the dark cavity. The <span class="hlt">bright</span> rim likely represents the pileup of the surrounding coronal plasma compressed by the expanding flux rope. The composite structure seen in AIA multiple <span class="hlt">temperature</span> bands is very similar to that in the corresponding coronagraph images, which consists of a <span class="hlt">bright</span> leading edge and a dark cavity, commonly believed to be a flux rope.</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://hdl.handle.net/2060/19800016766','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800016766"><span>The Barnes-Evans color-surface <span class="hlt">brightness</span> relation: A preliminary theoretical interpretation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shipman, H. L.</p> <p>1980-01-01</p> <p>Model atmosphere calculations are used to assess whether an empirically derived relation between V-R and surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> 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 <span class="hlt">temperature</span> cooler than 12,000 K. The error in white dwarf radii determined from a main sequence color surface <span class="hlt">brightness</span> relation is roughly 10%.</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://www.osti.gov/pages/biblio/1336139-laser-ion-source-high-brightness-heavy-ion-beam','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1336139-laser-ion-source-high-brightness-heavy-ion-beam"><span>Laser ion source for high <span class="hlt">brightness</span> heavy ion beam</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Okamura, M.</p> <p>2016-09-01</p> <p>A laser ion source is known as a high current high charge state heavy ion source. But, we place great emphasis on the capability to realize a high <span class="hlt">brightness</span> ion source. A laser ion source has a pinpoint small volume where materials are ionized and can achieve quite uniform low <span class="hlt">temperature</span> ion beam. Those features may enable us to realize very small emittance beams. Furthermore, a low charge state high <span class="hlt">brightness</span> laser ion source was successfully commissioned in Brookhaven National Laboratory in 2014. Now most of all the solid based heavy ions are being provided from the laser ion sourcemore » for regular operation.« less</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('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> <span class="hlt">Temperature</span> and Ocean Surface 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 surface emission is affected by: a) Sea surface <span class="hlt">temperature</span>. 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 surface 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('http://adsabs.harvard.edu/abs/2017NatAs...1..612S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatAs...1..612S"><span>The nature of solar <span class="hlt">brightness</span> variations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shapiro, A. I.; Solanki, S. K.; Krivova, N. A.; Cameron, R. H.; Yeo, K. L.; Schmutz, W. K.</p> <p>2017-09-01</p> <p>Determining the sources of solar <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> variability using high-cadence8,9 <span class="hlt">observations</span> from the Solar Dynamics Observatory (SDO) and the Spectral And Total Irradiance REconstruction (SATIRE)10,11 model. The <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> variations similar to that of the Sun, but with a much wider variety of patterns15. Our finding that solar <span class="hlt">brightness</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990044011','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990044011"><span>35 GHz Measurements of CO2 Crystals for Simulating <span class="hlt">Observations</span> of the Martian Polar Caps</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Foster, J. L.; Chang, A. T. C.; Hall, D. K.; Tait, A. B.; Barton, J. S.</p> <p>1998-01-01</p> <p>In order to learn more about the Martian polar caps, it is important to compare and contrast the behavior of both frozen H2O and CO2 in different parts of the electromagnetic spectrum. Relatively little attention has been given, thus far, to <span class="hlt">observing</span> the thermal microwave part of the spectrum. In this experiment, passive microwave radiation emanating from within a 33 cm snowpack was measured with a 35 GHz hand-held radiometer, and in addition to the natural snow measurements, the radiometer was used to measure the microwave emission and scattering from layers of manufactured CO2 (dry ice). A 1 m x 2 m plate of aluminum sheet metal was positioned beneath the natural snow so that microwave emissions from the underlying soil layers would be minimized. Compared to the natural snow crystals, results for the dry ice layers exhibit lower' microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> for similar thicknesses, regardless of the incidence angle of the radiometer. For example, at 50 degree H (horizontal polarization) and with a covering of 21 cm of snow and 18 cm of dry ice, the <span class="hlt">brightness</span> <span class="hlt">temperatures</span> were 150 K and 76 K, respectively. When the snow depth was 33 cm, the <span class="hlt">brightness</span> <span class="hlt">temperature</span> was 144 K, and when the total thickness of the dry ice was 27 cm, the <span class="hlt">brightness</span> <span class="hlt">temperature</span> was 86 K. The lower <span class="hlt">brightness</span> <span class="hlt">temperatures</span> are due to a combination of the lower physical <span class="hlt">temperature</span> and the larger crystal sizes of the commercial CO2 Crystals compared to the snow crystals. As the crystal size approaches the size of the microwave wavelength, it scatters microwave radiation more effectively, thus lowering the <span class="hlt">brightness</span> <span class="hlt">temperature</span>. The dry ice crystals in this experiment were about an order of magnitude larger than the snow crystals and three orders of magnitude larger than the CO2 Crystals produced in the cold stage of a scanning electron microscope. Spreading soil, approximately 2 mm in thickness, on the dry ice appeared to have no effect on the <span class="hlt">brightness</span> <span class="hlt">temperatures</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_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('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/2017NJPh...19g3018L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NJPh...19g3018L"><span>Alloying effect on <span class="hlt">bright</span>-dark exciton states in ternary monolayer Mo x W1-x Se2</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Yanping; Tom, Kyle; Zhang, Xiaowei; Lou, Shuai; Liu, Yin; Yao, Jie</p> <p>2017-07-01</p> <p>Binary transition metal dichalcogenides (TMDCs) in the class MX2 (M = Mo, W; X = S, Se) have been widely investigated for potential applications in optoelectronics and nanoelectronics. Recently, alloy-based monolayers of TMDCs have provided a stable and versatile technique to tune the physical properties and optimize them for potential applications. Here, we present experimental evidence for the existence of an intermediate alloy state between the MoSe2-like and the WSe2-like behavior of the neutral exciton (X 0) using <span class="hlt">temperature</span>-dependent photoluminescence (PL) of the monolayer Mo x W1-x Se2 alloy. The existence of a maximum PL intensity around 120 K can be explained by the competition between the thermally activated <span class="hlt">bright</span> states and the non-radiative quenching of the <span class="hlt">bright</span> states. Moreover, we also measured localized exciton (XB ) PL peak in the alloy and the <span class="hlt">observed</span> behavior agrees well with a model previously proposed for the 3D case, which indicates the theory also applies to 2D systems. Our results not only shed light on <span class="hlt">bright</span>-dark states and localized exciton physics of 2D semiconductors, but also offer a new route toward the control of the <span class="hlt">bright</span>-dark transition and tailoring optical properties of 2D semiconductors through defect engineering.</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-surface 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://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/2060/19760013983','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('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/2018A%26A...614A..54V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018A%26A...614A..54V"><span>LOFAR <span class="hlt">observations</span> of the quiet solar corona</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vocks, C.; Mann, G.; Breitling, F.; Bisi, M. M.; Dąbrowski, B.; Fallows, R.; Gallagher, P. T.; Krankowski, A.; Magdalenić, J.; Marqué, C.; Morosan, D.; Rucker, H.</p> <p>2018-06-01</p> <p>Context. The quiet solar corona emits meter-wave thermal bremsstrahlung. Coronal radio emission can only propagate above that radius, Rω, where the local plasma frequency equals the <span class="hlt">observing</span> frequency. The radio interferometer LOw Frequency ARray (LOFAR) <span class="hlt">observes</span> in its low band (10-90 MHz) solar radio emission originating from the middle and upper corona. Aims: We present the first solar aperture synthesis imaging <span class="hlt">observations</span> in the low band of LOFAR in 12 frequencies each separated by 5 MHz. From each of these radio maps we infer Rω, and a scale height <span class="hlt">temperature</span>, T. These results can be combined into coronal density and <span class="hlt">temperature</span> profiles. Methods: We derived radial intensity profiles from the radio images. We focus on polar directions with simpler, radial magnetic field structure. Intensity profiles were modeled by ray-tracing simulations, following wave paths through the refractive solar corona, and including free-free emission and absorption. We fitted model profiles to <span class="hlt">observations</span> with Rω and T as fitting parameters. Results: In the low corona, Rω < 1.5 solar radii, we find high scale height <span class="hlt">temperatures</span> up to 2.2 × 106 K, much more than the <span class="hlt">brightness</span> <span class="hlt">temperatures</span> usually found there. But if all Rω values are combined into a density profile, this profile can be fitted by a hydrostatic model with the same <span class="hlt">temperature</span>, thereby confirming this with two independent methods. The density profile deviates from the hydrostatic model above 1.5 solar radii, indicating the transition into the solar wind. Conclusions: These results demonstrate what information can be gleaned from solar low-frequency radio images. The scale height <span class="hlt">temperatures</span> we find are not only higher than <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, but also than <span class="hlt">temperatures</span> derived from coronograph or extreme ultraviolet (EUV) data. Future <span class="hlt">observations</span> will provide continuous frequency coverage. This continuous coverage eliminates the need for local hydrostatic density models in the data analysis and</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('https://www.osti.gov/pages/biblio/1375405-microwave-signatures-ice-hydrometeors-from-ground-based-observations-above-summit-greenland','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1375405-microwave-signatures-ice-hydrometeors-from-ground-based-observations-above-summit-greenland"><span>Microwave signatures of ice hydrometeors from ground-based <span class="hlt">observations</span> above Summit, Greenland</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Pettersen, Claire; Bennartz, Ralf; Kulie, Mark S.; ...</p> <p>2016-04-15</p> <p>Multi-instrument, ground-based measurements provide unique and comprehensive data sets of the atmosphere for a specific location over long periods of time and resulting data compliment past and existing global satellite <span class="hlt">observations</span>. Our paper explores the effect of ice hydrometeors on ground-based, high-frequency passive microwave measurements and attempts to isolate an ice signature for summer seasons at Summit, Greenland, from 2010 to 2013. Furthermore, data from a combination of passive microwave, cloud radar, radiosonde, and ceilometer were examined to isolate the ice signature at microwave wavelengths. By limiting the study to a cloud liquid water path of 40 g m -2more » or less, the cloud radar can identify cases where the precipitation was dominated by ice. These cases were examined using liquid water and gas microwave absorption models, and <span class="hlt">brightness</span> <span class="hlt">temperatures</span> were calculated for the high-frequency microwave channels: 90, 150, and 225GHz. By comparing the measured <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from the microwave radiometers and the calculated <span class="hlt">brightness</span> <span class="hlt">temperature</span> using only gas and liquid contributions, any residual <span class="hlt">brightness</span> <span class="hlt">temperature</span> difference is due to emission and scattering of microwave radiation from the ice hydrometeors in the column. The ice signature in the 90, 150, and 225 GHz channels for the Summit Station summer months was isolated. Then, this measured ice signature was compared to an equivalent <span class="hlt">brightness</span> <span class="hlt">temperature</span> difference calculated with a radiative transfer model including microwave single-scattering properties for several ice habits. Furthermore, initial model results compare well against the 4 years of summer season isolated ice signature in the high-frequency microwave channels.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990028291&hterms=Cluster+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DCluster%2Banalysis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990028291&hterms=Cluster+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DCluster%2Banalysis"><span>Analysis of Mass Profiles and Cooling Flows of <span class="hlt">Bright</span>, Early-Type Galaxies AO2, AO3 and Surface <span class="hlt">Brightness</span> Profiles and Energetics of Intracluster Gas in Cool Galaxy Clusters AO3</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>1998-01-01</p> <p>This final report uses ROSAT <span class="hlt">observations</span> to analyze two different studies. These studies are: Analysis of Mass Profiles and Cooling Flows of <span class="hlt">Bright</span>, Early-Type Galaxies; and Surface <span class="hlt">Brightness</span> Profiles and Energetics of Intracluster Gas in Cool Galaxy Clusters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770011045','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770011045"><span>Skylab experiment SO73: Gegenschein/zodiacal light. [electrophotometry of surface <span class="hlt">brightness</span> and polarization</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.</p> <p>1976-01-01</p> <p>A 10 color photoelectric polarimeter was used to measure the surface <span class="hlt">brightness</span> and polarization associated with zodiacal light, background starlight, and spacecraft corona during each of the Skylab missions. Fixed position and sky scanning <span class="hlt">observations</span> 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 <span class="hlt">brightness</span> of the zodiacal light. Included among the fixed position regions that were <span class="hlt">observed</span> 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 <span class="hlt">brightness</span> of the zodiacal light was found to have the color of the sun at each of these positions. Because previous <span class="hlt">observations</span> found the total <span class="hlt">brightness</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890066466&hterms=sonic+temperature&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsonic%2Btemperature','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890066466&hterms=sonic+temperature&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dsonic%2Btemperature"><span>Low-<span class="hlt">temperature</span> transonic cooling flows in galaxy clusters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sulkanen, Martin E.; Burns, Jack O.; Norman, Michael L.</p> <p>1989-01-01</p> <p>Calculations are presented which demonstrate that cooling flow models with large sonic radii may be consistent with <span class="hlt">observed</span> cluster gas properties. It is found that plausible cluster parameters and cooling flow mass accretion rates can produce sonic radii of 10-20 kpc for sonic point <span class="hlt">temperatures</span> of 1-3 x 10 to the 6th K. The numerical calculations match these cooling flows to hydrostatic atmosphere solutions for the cluster gas beyond the cooling flow region. The cooling flows produce no appreciable 'holes' in the surface <span class="hlt">brightness</span> toward the cluster center, and the model can be made to match the <span class="hlt">observed</span> X-ray surface <span class="hlt">brightness</span> of three clusters in which cooling flows had been believed to be absent. It is suggested that clusters with low velocity dispersion may be the natural location for such 'cool' cooling flows, and fits of these models to the X-ray surface <span class="hlt">brightness</span> profiles for three clusters are presented.</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('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 surface 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 surface thermal radiation. Higher energy <span class="hlt">observations</span> may be effected by poorly established effects including magnetized atmospheres, chemical compositions, <span class="hlt">temperature</span> 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('https://ntrs.nasa.gov/search.jsp?R=20150008580&hterms=tb&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dtb','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150008580&hterms=tb&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dtb"><span>Azimuthal Signature of Coincidental <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> and Normalized Radar Cross-Section Obtained Using Airborne PALS Instrument</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Colliander, Andreas; Kim, Seungbum; Yueh, Simon; Cosh, Mike; Jackson, Tom; Njoku, Eni</p> <p>2010-01-01</p> <p>Coincidental airborne <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) and normalized radar-cross section (NRCS) measurements were carried out with the PALS (Passive and Active L- and S-band) instrument in the SMAPVEX08 (SMAP Validation Experiment 2008) field campaign. This paper describes results obtained from a set of flights which measured a field in 45(sup o) steps over the azimuth angle. The field contained mature soy beans with distinct row structure. The measurement shows that both TB and NRCS experience modulation effects over the azimuth as expected based on the theory. The result is useful in development and validation of land surface parameter forward models and retrieval algorithms, such as the soil moisture algorithm for NASA's SMAP (Soil Moisture Active and Passive) mission. Although the footprint of the SMAP will not be sensitive to the small resolution scale effects as the one presented in this paper, it is nevertheless important to understand the effects at smaller scale.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22167295-chromospheric-solar-millimeter-wave-cavity-originates-temperature-minimum-region','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22167295-chromospheric-solar-millimeter-wave-cavity-originates-temperature-minimum-region"><span>THE CHROMOSPHERIC SOLAR MILLIMETER-WAVE CAVITY ORIGINATES IN THE <span class="hlt">TEMPERATURE</span> MINIMUM REGION</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>De la Luz, Victor; Raulin, Jean-Pierre; Lara, Alejandro</p> <p>2013-01-10</p> <p>We present a detailed theoretical analysis of the local radio emission at the lower part of the solar atmosphere. To accomplish this, we have used a numerical code to simulate the emission and transport of high-frequency electromagnetic waves from 2 GHz up to 10 THz. As initial conditions, we used VALC, SEL05, and C7 solar chromospheric models. In this way, the generated synthetic spectra allow us to study the local emission and absorption processes with high resolution in both altitude and frequency. Associated with the <span class="hlt">temperature</span> minimum predicted by these models, we found that the local optical depth at millimetermore » wavelengths remains constant, producing an optically thin layer that is surrounded by two layers of high local emission. We call this structure the Chromospheric Solar Millimeter-wave Cavity (CSMC). The <span class="hlt">temperature</span> profile, which features <span class="hlt">temperature</span> minimum layers and a subsequent <span class="hlt">temperature</span> rise, produces the CSMC phenomenon. The CSMC shows the complexity of the relation between the theoretical <span class="hlt">temperature</span> profile and the <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperature</span> and may help us to understand the dispersion of the <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperature</span> in the millimeter wavelength range.« less</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/2018AMT....11..611B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AMT....11..611B"><span>Retrieval of an ice water path over the ocean from ISMAR and MARSS millimeter and submillimeter <span class="hlt">brightness</span> <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brath, Manfred; Fox, Stuart; Eriksson, Patrick; Chawn Harlow, R.; Burgdorf, Martin; Buehler, Stefan A.</p> <p>2018-02-01</p> <p>A neural-network-based retrieval method to determine the snow ice water path (SIWP), liquid water path (LWP), and integrated water vapor (IWV) from millimeter and submillimeter <span class="hlt">brightness</span> <span class="hlt">temperatures</span>, measured by using airborne radiometers (ISMAR and MARSS), is presented. The neural networks were trained by using atmospheric profiles from the ICON numerical weather prediction (NWP) model and by radiative transfer simulations using the Atmospheric Radiative Transfer Simulator (ARTS). The basic performance of the retrieval method was analyzed in terms of offset (bias) and the median fractional error (MFE), and the benefit of using submillimeter channels was studied in comparison to pure microwave retrievals. The retrieval is offset-free for SIWP > 0.01 kg m-2, LWP > 0.1 kg m-2, and IWV > 3 kg m-2. The MFE of SIWP decreases from 100 % at SIWP = 0.01 kg m-2 to 20 % at SIWP = 1 kg m-2 and the MFE of LWP from 100 % at LWP = 0.05 kg m-2 to 30 % at LWP = 1 kg m-2. The MFE of IWV for IWV > 3 kg m-2 is 5 to 8 %. The SIWP retrieval strongly benefits from submillimeter channels, which reduce the MFE by a factor of 2, compared to pure microwave retrievals. The IWV and the LWP retrievals also benefit from submillimeter channels, albeit to a lesser degree. The retrieval was applied to ISMAR and MARSS <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from FAAM flight B897 on 18 March 2015 of a precipitating frontal system west of the coast of Iceland. Considering the given uncertainties, the retrieval is in reasonable agreement with the SIWP, LWP, and IWV values simulated by the ICON NWP model for that flight. A comparison of the retrieved IWV with IWV from 12 dropsonde measurements shows an offset of 0.5 kg m-2 and an RMS difference of 0.8 kg m-2, showing that the retrieval of IWV is highly effective even under cloudy conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16210535','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16210535"><span>A 5-micron-<span class="hlt">bright</span> spot on Titan: evidence for surface diversity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>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</p> <p>2005-10-07</p> <p><span class="hlt">Observations</span> from the Cassini Visual and Infrared Mapping Spectrometer show an anomalously <span class="hlt">bright</span> spot on Titan located at 80 degrees W and 20 degrees S. This area is <span class="hlt">bright</span> in reflected light at all <span class="hlt">observed</span> 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.</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('https://www.ncbi.nlm.nih.gov/pubmed/28366816','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28366816"><span>The impact of morning light intensity and environmental <span class="hlt">temperature</span> on body <span class="hlt">temperatures</span> and alertness.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Te Kulve, Marije; Schlangen, Luc J M; Schellen, Lisje; Frijns, Arjan J H; van Marken Lichtenbelt, Wouter D</p> <p>2017-06-01</p> <p>Indoor <span class="hlt">temperature</span> and light exposure are known to affect body <span class="hlt">temperature</span>, productivity and alertness of building occupants. However, not much is known about the interaction between light and <span class="hlt">temperature</span> exposure and the relationship between morning light induced alertness and its effect on body <span class="hlt">temperature</span>. Light intensity and room <span class="hlt">temperature</span> during morning office hours were investigated under strictly controlled conditions. In a randomized crossover study, two white light conditions (4000K, either <span class="hlt">bright</span> 1200lx or dim 5lx) under three different room <span class="hlt">temperatures</span> (26, 29 and 32°C) were investigated. A lower room <span class="hlt">temperature</span> increased the core body <span class="hlt">temperature</span> (CBT) and lowered skin <span class="hlt">temperature</span> and the distal-proximal <span class="hlt">temperature</span> gradient (DPG). Moreover, a lower room <span class="hlt">temperature</span> reduced the subjective sleepiness and reaction time on an auditory psychomotor vigilance task (PVT), irrespective of the light condition. Interestingly, the morning <span class="hlt">bright</span> light exposure did affect thermophysiological parameters, i.e. it decreased plasma cortisol, CBT and proximal skin <span class="hlt">temperature</span> and increased the DPG, irrespective of the room <span class="hlt">temperature</span>. During the <span class="hlt">bright</span> light session, subjective sleepiness decreased irrespective of the room <span class="hlt">temperature</span>. However, the change in sleepiness due to the light exposure was not related to these physiological changes. Copyright © 2017 Elsevier Inc. All rights reserved.</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://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' Surface <span class="hlt">Temperature</span>.</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 surface <span class="hlt">temperatures</span> 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 surface <span class="hlt">temperatures</span> below 180 K. This restricts <span class="hlt">temperature</span> 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 surface <span class="hlt">temperatures</span> 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 surface. The diurnal <span class="hlt">temperature</span> profiles were compared to the KRC thermal model [2, 3], which has been extensively used for Mars [e.g. 4, 5]. Variations in <span class="hlt">temperature</span> as a function of local time are <span class="hlt">observed</span> and are compared to predictions from the KRC model. The model <span class="hlt">temperatures</span> are converted to radiance (Jy/Steradian) and are corrected for near-surface thermal gradients and limb effects for comparison to <span class="hlt">observations</span>. Initial analysis is consistent with the presence of near-surface 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('http://hdl.handle.net/2060/20130014383','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130014383"><span>Surface <span class="hlt">Temperature</span> 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 surface <span class="hlt">temperature</span> 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) surface <span class="hlt">temperatures</span> based on luminescence decay of a thermographic phosphor could be obtained from the surface 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 surface <span class="hlt">temperature</span> 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://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://hdl.handle.net/2060/20170005686','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20170005686"><span>Evaluating Soil Moisture Retrievals from ESA's SMOS and NASA's SMAP <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Datasets</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Al-Yaari, A.; Wigernon, J.-P.; Kerr, Y.; Rodriguez-Fernandez, N.; O'Neill, P. E.; Jackson, T. J.; De Lannoy, G. J. M.; Al Bitar, A.; Mialon, A.; Richaume, P.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170005686'); toggleEditAbsImage('author_20170005686_show'); toggleEditAbsImage('author_20170005686_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170005686_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170005686_hide"></p> <p>2017-01-01</p> <p>Two satellites are currently monitoring surface soil moisture (SM) using L-band <span class="hlt">observations</span>: SMOS (Soil Moisture and Ocean Salinity), a joint ESA (European Space Agency), CNES (Centre national d'tudes spatiales), and CDTI (the Spanish government agency with responsibility for space) satellite launched on November 2, 2009 and SMAP (Soil Moisture Active Passive), a National Aeronautics and Space Administration (NASA) satellite successfully launched in January 2015. In this study, we used a multilinear regression approach to retrieve SM from SMAP data to create a global dataset of SM, which is consistent with SM data retrieved from SMOS. This was achieved by calibrating coefficients of the regression model using the CATDS (Centre Aval de Traitement des Donnes) SMOS Level 3 SM and the horizontally and vertically polarized <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TB) at 40 deg incidence angle, over the 2013 - 2014 period. Next, this model was applied to SMAP L3 TB data from Apr 2015 to Jul 2016. The retrieved SM from SMAP (referred to here as SMAP_Reg) was compared to: (i) the operational SMAP L3 SM (SMAP_SCA), retrieved using the baseline Single Channel retrieval Algorithm (SCA); and (ii) the operational SMOSL3 SM, derived from the multiangular inversion of the L-MEB model (L-MEB algorithm) (SMOSL3). This inter-comparison was made against in situ soil moisture measurements from more than 400 sites spread over the globe, which are used here as a reference soil moisture dataset. The in situ <span class="hlt">observations</span> were obtained from the International Soil Moisture Network (ISMN; https:ismn.geo.tuwien.ac.at) in North of America (PBO_H2O, SCAN, SNOTEL, iRON, and USCRN), in Australia (Oznet), Africa (DAHRA), and in Europe (REMEDHUS, SMOSMANIA, FMI, and RSMN). The agreement was analyzed in terms of four classical statistical criteria: Root Mean Squared Error (RMSE),Bias, Unbiased RMSE (UnbRMSE), and correlation coefficient (R). Results of the comparison of these various products with in situ</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29743730','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29743730"><span>Evaluating soil moisture retrievals from ESA's SMOS and NASA's SMAP <span class="hlt">brightness</span> <span class="hlt">temperature</span> datasets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Al-Yaari, A; Wigneron, J-P; Kerr, Y; Rodriguez-Fernandez, N; O'Neill, P E; Jackson, T J; De Lannoy, G J M; Al Bitar, A; Mialon, A; Richaume, P; Walker, J P; Mahmoodi, A; Yueh, S</p> <p>2017-05-01</p> <p>Two satellites are currently monitoring surface soil moisture (SM) using L-band <span class="hlt">observations</span>: SMOS (Soil Moisture and Ocean Salinity), a joint ESA (European Space Agency), CNES (Centre national d'études spatiales), and CDTI (the Spanish government agency with responsibility for space) satellite launched on November 2, 2009 and SMAP (Soil Moisture Active Passive), a National Aeronautics and Space Administration (NASA) satellite successfully launched in January 2015. In this study, we used a multilinear regression approach to retrieve SM from SMAP data to create a global dataset of SM, which is consistent with SM data retrieved from SMOS. This was achieved by calibrating coefficients of the regression model using the CATDS (Centre Aval de Traitement des Données) SMOS Level 3 SM and the horizontally and vertically polarized <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TB) at 40° incidence angle, over the 2013 - 2014 period. Next, this model was applied to SMAP L3 TB data from Apr 2015 to Jul 2016. The retrieved SM from SMAP (referred to here as SMAP_Reg) was compared to: (i) the operational SMAP L3 SM (SMAP_SCA), retrieved using the baseline Single Channel retrieval Algorithm (SCA); and (ii) the operational SMOSL3 SM, derived from the multiangular inversion of the L-MEB model (L-MEB algorithm) (SMOSL3). This inter-comparison was made against in situ soil moisture measurements from more than 400 sites spread over the globe, which are used here as a reference soil moisture dataset. The in situ <span class="hlt">observations</span> were obtained from the International Soil Moisture Network (ISMN; https://ismn.geo.tuwien.ac.at/) in North of America (PBO_H2O, SCAN, SNOTEL, iRON, and USCRN), in Australia (Oznet), Africa (DAHRA), and in Europe (REMEDHUS, SMOSMANIA, FMI, and RSMN). The agreement was analyzed in terms of four classical statistical criteria: Root Mean Squared Error (RMSE), Bias, Unbiased RMSE (UnbRMSE), and correlation coefficient (R). Results of the comparison of these various products with in situ</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/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://hdl.handle.net/2060/20080023294','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080023294"><span>A Physical Model to Estimate Snowfall over Land using AMSU-B <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>Kim, Min-Jeong; Weinman, J. A.; Olson, W. S.; Chang, D.-E.; Skofronick-Jackson, G.; Wang, J. R.</p> <p>2008-01-01</p> <p>In this study, we present an improved physical model to retrieve snowfall rate over land using <span class="hlt">brightness</span> <span class="hlt">temperature</span> <span class="hlt">observations</span> from the National Oceanic and Atmospheric Administration's (NOAA) Advanced Microwave Sounder Unit-B (AMSU-B) at 89 GHz, 150 GHz, 183.3 +/- 1 GHz, 183.3 +/- 3 GHz, and 183.3 +/- 7 GHz. The retrieval model is applied to the New England blizzard of March 5, 2001 which deposited about 75 cm of snow over much of Vermont, New Hampshire, and northern New York. In this improved physical model, prior retrieval assumptions about snowflake shape, particle size distributions, environmental conditions, and optimization methodology have been updated. Here, single scattering parameters for snow particles are calculated with the Discrete-Dipole Approximation (DDA) method instead of assuming spherical shapes. Five different snow particle models (hexagonal columns, hexagonal plates, and three different kinds of aggregates) are considered. Snow particle size distributions are assumed to vary with air <span class="hlt">temperature</span> and to follow aircraft measurements described by previous studies. <span class="hlt">Brightness</span> <span class="hlt">temperatures</span> at AMSU-B frequencies for the New England blizzard are calculated using these DDA calculated single scattering parameters and particle size distributions. The vertical profiles of pressure, <span class="hlt">temperature</span>, relative humidity and hydrometeors are provided by MM5 model simulations. These profiles are treated as the a priori data base in the Bayesian retrieval algorithm. In algorithm applications to the blizzard data, calculated <span class="hlt">brightness</span> <span class="hlt">temperatures</span> associated with selected database profiles agree with AMSU-B <span class="hlt">observations</span> to within about +/- 5 K at all five frequencies. Retrieved snowfall rates compare favorably with the near-concurrent National Weather Service (NWS) radar reflectivity measurements. The relationships between the NWS radar measured reflectivities Z(sub e) and retrieved snowfall rate R for a given snow particle model are derived by a histogram</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA00427.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA00427.html"><span>Global Average <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> for April 2003</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2003-06-02</p> <p>This image shows average <span class="hlt">temperatures</span> in April, 2003, <span class="hlt">observed</span> by AIRS at an infrared wavelength that senses either the Earth's surface or any intervening cloud. Similar to a photograph of the planet taken with the camera shutter held open for a month, stationary features are captured while those obscured by moving clouds are blurred. Many continental features stand out boldly, such as our planet's vast deserts, and India, now at the end of its long, clear dry season. Also obvious are the high, cold Tibetan plateau to the north of India, and the mountains of North America. The band of yellow encircling the planet's equator is the Intertropical Convergence Zone (ITCZ), a region of persistent thunderstorms and associated high, cold clouds. The ITCZ merges with the monsoon systems of Africa and South America. Higher latitudes are increasingly obscured by clouds, though some features like the Great Lakes, the British Isles and Korea are apparent. The highest latitudes of Europe and Eurasia are completely obscured by clouds, while Antarctica stands out cold and clear at the bottom of the image. http://photojournal.jpl.nasa.gov/catalog/PIA00427</p> </li> <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=19820055687&hterms=oceans+puerto+rico&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doceans%2Bpuerto%2Brico','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820055687&hterms=oceans+puerto+rico&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Doceans%2Bpuerto%2Brico"><span>Passive microwave measurements of <span class="hlt">temperature</span> and salinity in coastal zones</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Blume, H.-J. C.; Kendall, B. M.</p> <p>1982-01-01</p> <p>Experimental methods and results from the maritime remote sensing (MARSEN) experiments using dual frequency microwave radiometer detecting systems on board aircraft are described. The radiometers were operated at 1.43 and 2.65 GHz and flown above U.S. Atlantic coastal areas, Chesapeake Bay, around Puerto Rico, and over the German Bight. The advanced switched radiometers used were configured to be independent of gain variations and errors originating from front-end losses and determined the absolute <span class="hlt">brightness</span> <span class="hlt">temperatures</span> to within a few tenths Kelvin. Corrections to the <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperature</span> of the ocean are analytically defined, including accounts made for roughness, the cosmic background radiation, and the solar radio source. The coastal flight data for salinity gradients and surface <span class="hlt">temperatures</span> were compared with sea truth measured from ships and found to be accurate to within 1 C and 1 pph.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12502234','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12502234"><span>Effects of electromagnetic radiation (<span class="hlt">bright</span> light, extremely low-frequency magnetic fields, infrared radiation) on the circadian rhythm of melatonin synthesis, rectal <span class="hlt">temperature</span>, and heart rate.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Griefahn, Barbara; Künemund, Christa; Blaszkewicz, Meinolf; Lerchl, Alexander; Degen, Gisela H</p> <p>2002-10-01</p> <p>Electromagnetic spectra reduce melatonin production and delay the nadirs of rectal <span class="hlt">temperature</span> and heart rate. Seven healthy men (16-22 yrs) completed 4 permuted sessions. The control session consisted of a 24-hours bedrest at < 30 lux, 18 degrees C, and < 50 dBA. In the experimental sessions, either light (1500 lux), magnetic field (16.7 Hz, 0.2 mT), or infrared radiation (65 degrees C) was applied from 5 pm to 1 am. Salivary melatonin level was determined hourly, rectal <span class="hlt">temperature</span> and heart rate were continuously recorded. Melatonin synthesis was completely suppressed by light but resumed thereafter. The nadirs of rectal <span class="hlt">temperature</span> and heart rate were delayed. The magnetic field had no effect. Infrared radiation elevated rectal <span class="hlt">temperature</span> and heart rate. Only <span class="hlt">bright</span> light affected the circadian rhythms of melatonin synthesis, rectal <span class="hlt">temperature</span>, and heart rate, however, differently thus causing a dissociation, which might enhance the adverse effects of shiftwork in the long run.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26191921','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26191921"><span>Fabrication of <span class="hlt">bright</span> and thin Zn₂SiO₄ luminescent film for electron beam excitation-assisted optical microscope.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Furukawa, Taichi; Kanamori, Satoshi; Fukuta, Masahiro; Nawa, Yasunori; Kominami, Hiroko; Nakanishi, Yoichiro; Sugita, Atsushi; Inami, Wataru; Kawata, Yoshimasa</p> <p>2015-07-13</p> <p>We fabricated a <span class="hlt">bright</span> and thin Zn₂SiO₄ luminescent film to serve as a nanometric light source for high-spatial-resolution optical microscopy based on electron beam excitation. The Zn₂SiO₄ luminescent thin film was fabricated by annealing a ZnO film on a Si₃N₄ substrate at 1000 °C in N₂. The annealed film emitted <span class="hlt">bright</span> cathodoluminescence compared with the as-deposited film. The film is promising for nano-imaging with electron beam excitation-assisted optical microscopy. We evaluated the spatial resolution of a microscope developed using this Zn₂SiO₄ luminescent thin film. This is the first report of the investigation and application of ZnO/Si₃N₄ annealed at a high <span class="hlt">temperature</span> (1000 °C). The fabricated Zn₂SiO₄ film is expected to enable high-frame-rate dynamic <span class="hlt">observation</span> with ultra-high resolution using our electron beam excitation-assisted optical microscopy.</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://pubs.er.usgs.gov/publication/70033421','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033421"><span>Summer season variability of the north residual cap of Mars as <span class="hlt">observed</span> by the Mars Global Surveyor Thermal Emission Spectrometer (MGS-TES)</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Calvin, W.M.; Titus, T.N.</p> <p>2008-01-01</p> <p>Previous <span class="hlt">observations</span> have noted the change in albedo in a number of North Pole <span class="hlt">bright</span> outliers and in the distribution of <span class="hlt">bright</span> ice deposits between Mariner 9, Viking, and Mars Global Surveyor (MGS) data sets. Changes over the summer season as well as between regions at the same season (Ls) in different years have been <span class="hlt">observed</span>. We used the bolometric albedo and <span class="hlt">brightness</span> <span class="hlt">temperature</span> channels of the Thermal Emission Spectrometer (TES) on the MGS spacecraft to monitor north polar residual ice cap variations between Mars years and within the summer season for three northern Martian summers between July 1999 and April 2003. Large-scale <span class="hlt">brightness</span> variations are <span class="hlt">observed</span> in four general areas: (1) the patchy outlying frost deposits from 90 to 270??E, 75 to 80??N; (2) the large "tail" below the Chasma Boreale and its associated plateau from 315 to 45??E, 80 to 85??N, that we call the "Boreale Tongue" and in Hyperboreae Undae; (3) the troughed terrain in the region from 0 to 120??E longitude (the lower right on a polar stereographic projection) we have called "Shackleton's Grooves" and (4) the unit mapped as residual ice in Olympia Planitia. We also note two areas which seem to persist as cool and <span class="hlt">bright</span> throughout the summer and between Mars years. One is at the "source" of Chasma Boreale (???15??E, 85??N) dubbed "McMurdo", and the "Cool and <span class="hlt">Bright</span> Anomaly (CABA)" noted by Kieffer and Titus 2001. TES Mapping of Mars' north seasonal cap. Icarus 154, 162-180] at ???330??E, 87??N called here "Vostok". Overall defrosting occurs early in the summer as the <span class="hlt">temperatures</span> rise and then after the peak <span class="hlt">temperatures</span> are reached (Ls???110) higher elevations and outlier <span class="hlt">bright</span> deposits cold trap and re-accumulate new frost. Persistent <span class="hlt">bright</span> areas are associated with either higher elevations or higher background albedos suggesting complex feedback mechanisms including cold-trapping of frost due to albedo and elevation effects, as well as influence of mesoscale atmospheric dynamics</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('http://adsabs.harvard.edu/abs/2017ApJ...847...37C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...847...37C"><span>Do Low Surface <span class="hlt">Brightness</span> Galaxies Host Stellar Bars?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cervantes Sodi, Bernardo; Sánchez García, Osbaldo</p> <p>2017-09-01</p> <p>With the aim of assessing if low surface <span class="hlt">brightness</span> galaxies host stellar bars and by studying the dependence of the occurrence of bars as a function of surface <span class="hlt">brightness</span>, 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 <span class="hlt">brightness</span> in terms of their central surface <span class="hlt">brightness</span>. We find that the fraction of low surface <span class="hlt">brightness</span> galaxies hosting strong bars is systematically lower than that found for high surface <span class="hlt">brightness</span> galaxies. The dependence of the bar fraction on the central surface <span class="hlt">brightness</span> is mostly driven by a correlation of the surface <span class="hlt">brightness</span> with the spin and the gas richness of the galaxies, showing only a minor dependence on the surface <span class="hlt">brightness</span>. We also find that the length of the bars is strongly dependent on the surface <span class="hlt">brightness</span>, and although some of this dependence is attributed to the gas content, even at a fixed gas-to-stellar mass ratio, high surface <span class="hlt">brightness</span> galaxies host longer bars than their low surface <span class="hlt">brightness</span> counterparts, which we attribute to an anticorrelation of the surface <span class="hlt">brightness</span> with the spin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22679825-do-low-surface-brightness-galaxies-host-stellar-bars','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22679825-do-low-surface-brightness-galaxies-host-stellar-bars"><span>Do Low Surface <span class="hlt">Brightness</span> Galaxies Host Stellar Bars?</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>Cervantes Sodi, Bernardo; Sánchez García, Osbaldo, E-mail: b.cervantes@irya.unam.mx, E-mail: o.sanchez@irya.unam.mx</p> <p></p> <p>With the aim of assessing if low surface <span class="hlt">brightness</span> galaxies host stellar bars and by studying the dependence of the occurrence of bars as a function of surface <span class="hlt">brightness</span>, 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 <span class="hlt">brightness</span> in terms of their central surface <span class="hlt">brightness</span>. We find that the fraction of low surface <span class="hlt">brightness</span> galaxies hosting strong bars is systematically lower than that found for high surface <span class="hlt">brightness</span> galaxies. The dependence of the bar fraction on the central surface <span class="hlt">brightness</span> ismore » mostly driven by a correlation of the surface <span class="hlt">brightness</span> with the spin and the gas richness of the galaxies, showing only a minor dependence on the surface <span class="hlt">brightness</span>. We also find that the length of the bars is strongly dependent on the surface <span class="hlt">brightness</span>, and although some of this dependence is attributed to the gas content, even at a fixed gas-to-stellar mass ratio, high surface <span class="hlt">brightness</span> galaxies host longer bars than their low surface <span class="hlt">brightness</span> counterparts, which we attribute to an anticorrelation of the surface <span class="hlt">brightness</span> with the spin.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SSEle.121...20K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SSEle.121...20K"><span>Measurement of <span class="hlt">brightness</span> <span class="hlt">temperature</span> of two-dimensional electron gas in channel of a high electron mobility transistor at ultralow dissipation power</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Korolev, A. M.; Shulga, V. M.; Turutanov, O. G.; Shnyrkov, V. I.</p> <p>2016-07-01</p> <p>A technically simple and physically clear method is suggested for direct measurement of the <span class="hlt">brightness</span> <span class="hlt">temperature</span> of two-dimensional electron gas (2DEG) in the channel of a high electron mobility transistor (HEMT). The usage of the method was demonstrated with the pseudomorphic HEMT as a specimen. The optimal HEMT dc regime, from the point of view of the "back action" problem, was found to belong to the unsaturated area of the static characteristics possibly corresponding to the ballistic electron transport mode. The proposed method is believed to be a convenient tool to explore the ballistic transport, electron diffusion, 2DEG properties and other electrophysical processes in heterostructures.</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://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 <span class="hlt">temperatures</span> 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 surface <span class="hlt">brightness</span> measurements made in the ultraviolet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1983A%26A...119..319G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1983A%26A...119..319G"><span>Detection of a late B star companion of the <span class="hlt">bright</span> cluster giant C PUP equals HD 63032</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Groote, D.; Reimers, D.</p> <p>1983-03-01</p> <p>IUE <span class="hlt">observations</span> show that c Pup, the central <span class="hlt">bright</span> K giant in the open cluster NGC 2451, has a blue companion. A fit of theoretical line blanketed model atmosphere fluxes to the <span class="hlt">observed</span> energy distribution yields reddening E(B-V) = 0.15 (from λ2200 Å feature), an effective <span class="hlt">temperature</span> Te = 10,200K, and an angular diameter θ = 0.060. If the companion is a main-sequence star, c Pup and its companion are located at a distance of 310 ± 50 pc which lends additional support to membership of c Pup in NGC 2451. The evolutionary status of c Pup is briefly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMSH13B4105B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSH13B4105B"><span>Harnessing AIA Diffraction Patterns to Determine Flare Footpoint <span class="hlt">Temperatures</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bain, H. M.; Schwartz, R. A.; Torre, G.; Krucker, S.; Raftery, C. L.</p> <p>2014-12-01</p> <p>In the "Standard Flare Model" energy from accelerated electrons is deposited at the footpoints of newly reconnected flare loops, heating the surrounding plasma. Understanding the relation between the multi-thermal nature of the footpoints and the energy flux from accelerated electrons is therefore fundamental to flare physics. Extreme ultraviolet (EUV) images of <span class="hlt">bright</span> flare kernels, obtained from the Atmospheric Imaging Assembly (AIA) onboard the Solar Dynamics Observatory, are often saturated despite the implementation of automatic exposure control. These kernels produce diffraction patterns often seen in AIA images during the most energetic flares. We implement an automated image reconstruction procedure, which utilizes diffraction pattern artifacts, to de-saturate AIA images and reconstruct the flare <span class="hlt">brightness</span> in saturated pixels. Applying this technique to recover the footpoint <span class="hlt">brightness</span> in each of the AIA EUV passbands, we investigate the footpoint <span class="hlt">temperature</span> distribution. Using <span class="hlt">observations</span> from the Ramaty High Energy Solar Spectroscopic Imager (RHESSI), we will characterize the footpoint accelerated electron distribution of the flare. By combining these techniques, we investigate the relation between the nonthermal electron energy flux and the <span class="hlt">temperature</span> response of the flare footpoints.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840016854','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840016854"><span>Calculations of atmospheric transmittance in the 11 micrometer window for estimating skin <span class="hlt">temperature</span> from VISSR infrared <span class="hlt">brightness</span> <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chesters, D.</p> <p>1984-01-01</p> <p>An algorithm for calculating the atmospheric transmittance in the 10 to 20 micro m spectral band from a known <span class="hlt">temperature</span> and dewpoint profile, and then using this transmittance to estimate the surface (skin) <span class="hlt">temperature</span> from a VISSR <span class="hlt">observation</span> in the 11 micro m window is presented. Parameterizations are drawn from the literature for computing the molecular absorption due to the water vapor continuum, water vapor lines, and carbon dioxide lines. The FORTRAN code is documented for this application, and the sensitivity of the derived skin <span class="hlt">temperature</span> to variations in the model's parameters is calculated. The VISSR calibration uncertainties are identified as the largest potential source of error.</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('http://adsabs.harvard.edu/abs/2017AGUFMGC54C..01M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMGC54C..01M"><span>Global Sea Surface <span class="hlt">Temperature</span>: A Harmonized Multi-sensor Time-series from 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>Merchant, C. J.</p> <p>2017-12-01</p> <p>This paper presents the methods used to obtain a new global sea surface <span class="hlt">temperature</span> (SST) dataset spanning the early 1980s to the present, intended for use as a climate data record (CDR). The dataset provides skin SST (the fundamental measurement) and an estimate of the daily mean SST at depths compatible with drifting buoys (adjusting for skin and diurnal variability). The depth SST provided enables the CDR to be used with in situ records and centennial-scale SST reconstructions. The new SST timeseries is as independent as possible from in situ <span class="hlt">observations</span>, and from 1995 onwards is harmonized to an independent satellite reference (namely, SSTs from the Advanced Along Track Scanning Radiometer (Advanced ATSR)). This maximizes the utility of our new estimates of variability and long-term trends in interrogating previous datasets tied to in situ <span class="hlt">observations</span>. The new SSTs include full resolution (swath, level 2) data, single-sensor gridded data (level 3, 0.05 degree latitude-longitude grid) and a multi-sensor optimal analysis (level 4, same grid). All product levels are consistent. All SSTs have validated uncertainty estimates attached. The sensors used include all Advanced Very High Resolution Radiometers from NOAA-6 onwards and the ATSR series. AVHRR <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (BTs) are calculated from counts using a new in-flight re-calibration for each sensor, ultimately linked through to the AATSR BT calibration by a new harmonization technique. Artefacts in AVHRR BTs linked to varying instrument <span class="hlt">temperature</span>, orbital regime and solar contamination are significantly reduced. These improvements in the AVHRR BTs (level 1) translate into improved cloud detection and SST (level 2). For cloud detection, we use a Bayesian approach for all sensors. For the ATSRs, SSTs are derived with sufficient accuracy and sensitivity using dual-view coefficients. This is not the case for single-view AVHRR <span class="hlt">observations</span>, for which a physically based retrieval is employed, using a hybrid</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/992944','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/992944"><span>Thermometric- and Acoustic-Based Beam Power Monitor for Ultra-<span class="hlt">Bright</span> X-Rays</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>Bentsen, Gregory; /Rochester U. /SLAC</p> <p>2010-08-25</p> <p>A design for an average beam power monitor for ultra-<span class="hlt">bright</span> X-ray sources is proposed that makes simultaneous use of calorimetry and radiation acoustics. Radiation incident on a solid target will induce heating and ultrasonic vibrations, both of which may be measured to give a fairly precise value of the beam power. The monitor is intended for measuring ultra-<span class="hlt">bright</span> Free-Electron Laser (FEL) X-ray beams, for which traditional monitoring technologies such as photo-diodes or scintillators are unsuitable. The monitor consists of a Boron Carbide (B{sub 4}C) target designed to absorb most of the incident beam's energy. Resistance <span class="hlt">temperature</span> detectors (RTD) and piezoelectricmore » actuators are mounted on the outward faces of the target to measure the <span class="hlt">temperature</span> changes and ultrasonic vibrations induced by the incident beam. The design was tested using an optical pulsed beam (780 nm, 120 and 360 Hz) from a Ti:sapphire oscillator at several energies between 0.8 and 2.6 mJ. The RTDs measured an increase in <span class="hlt">temperature</span> of about 10 K over a period of several minutes. The piezoelectric sensors recorded ringing acoustic oscillations at 580 {+-} 40 kHz. Most importantly, the amplitude of the acoustic signals was <span class="hlt">observed</span> to scale linearly with beam power up to 2 mJ of pulse energy. Above this pulse energy, the vibrational signals became nonlinear. Several causes for this nonlinearity are discussed, including amplifier saturation and piezoelectric saturation. Despite this nonlinearity, these measurements demonstrate the feasibility of such a beam power measurement device. The advantage of two distinct measurements (acoustic and thermometric) provides a useful method of calibration that is unavailable to current LCLS diagnostics tools.« less</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('http://adsabs.harvard.edu/abs/2014AGUFM.A51I3147H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A51I3147H"><span>An Inter-calibrated Passive Microwave <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> Data Record and Ocean Products</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hilburn, K. A.; Wentz, F. J.</p> <p>2014-12-01</p> <p>Inter-calibration of passive microwave sensors has been the subject of on-going activity at Remote Sensing Systems (RSS) since 1974. RSS has produced a <span class="hlt">brightness</span> <span class="hlt">temperature</span> TB data record that spans the last 28 years (1987-2014) from inter-calibrated passive microwave sensors on 14 satellites: AMSR-E, AMSR2, GMI, SSMI F08-F15, SSMIS F16-F18, TMI, WindSat. Accompanying the TB record are a suite of ocean products derived from the TBs that provide a 28-year record of wind speed, water vapor, cloud liquid, and rain rate; and 18 years (1997-2014) of sea surface <span class="hlt">temperatures</span>, corresponding to the period for which 6 and/or 10 GHz measurements are available. Crucial to the inter-calibration and ocean product retrieval are a highly accurate radiative transfer model RTM. The RSS RTM has been continually refined for over 30 years and is arguably the most accurate model in the 1-100 GHz spectrum. The current generation of TB and ocean products, produced using the latest version of the RTM, is called Version-7. The accuracy of the Version-7 inter-calibration is estimated to be 0.1 K, based on inter-satellite comparisons and validation of the ocean products against in situ measurements. The data record produced by RSS has had a significant scientific impact. Over just the last 14 years (2000-2013) RSS data have been used in 743 peer-reviewed journal articles. This is an average of 4.5 peer-reviewed papers published every month made possible with RSS data. Some of the most important scientific contributions made by RSS data have been to the study of the climate. The AR5 Report "Climate Change 2013: The Physical Science Basis" by the Intergovernmental Panel on Climate Change (IPCC), the internationally accepted authority on climate change, references 20 peer-reviewed journal papers from RSS scientists. The report makes direct use of RSS water vapor data, RSS atmospheric <span class="hlt">temperatures</span> from MSU/AMSU, and 9 other datasets that are derived from RSS data. The RSS TB data record is</p> </li> <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('https://ntrs.nasa.gov/search.jsp?R=19920046809&hterms=sos&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dsos','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920046809&hterms=sos&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dsos"><span>A catalog of low surface <span class="hlt">brightness</span> galaxies - List II</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schombert, James M.; Bothun, Gregory D.; Schneider, Stephen E.; Mcgaugh, Stacy S.</p> <p>1992-01-01</p> <p>A list of galaxies characterized by low surface <span class="hlt">brightness</span> (LSB) is presented which facilitates the recognition of galaxies with <span class="hlt">brightnesses</span> 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-<span class="hlt">brightness</span> 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 <span class="hlt">observational</span> 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.</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://adsabs.harvard.edu/abs/2011SPD....42.2315C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011SPD....42.2315C"><span><span class="hlt">Observing</span> Flux Rope Formation During the Impulsive Phase of a Solar Eruption</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cheng, Xin; Zhang, J.; Yang, L.; Ding, M.</p> <p>2011-05-01</p> <p>Magnetic flux rope is believed to be an important structural component of coronal mass ejections (CMEs). While there exist much <span class="hlt">observational</span> evidence of the flux rope post the eruption, e.g., as seen in remote-sensing coronagraph images or in-situ solar wind data, the direct <span class="hlt">observation</span> of flux ropes during CME impulsive phase has been rare or non-exist. In this Letter, we present an unambiguous <span class="hlt">observation</span> of a flux rope still in the formation phase in the low corona. The CME of interest occurred above the east limb on 2010 November 03 with footpoints partially blocked. The flux rope was seen as a blob of hot plasma in AIA 131 A passband (peak <span class="hlt">temperature</span> 11 MK) rising from the core of the source active region, rapidly moving outward and stretching upward the surrounding background magnetic field. The stretched magnetic field seemed to curve-in, similar to the classical magnetic reconnection scenario in eruptive flares. The flux rope was also seen as a dark cavity in AIA 211 A passpand (2.0 MK) and 171 A passband (0.6 MK); in these relatively cool <span class="hlt">temperature</span> bands, a <span class="hlt">bright</span> rim clearly enclosed the dark cavity. The <span class="hlt">bright</span> rim likely represents the pile-up of the surrounding coronal plasma compressed by the expanding flux rope. The composite structure seen in AIA multiple <span class="hlt">temperature</span> bands is very similar to that in the corresponding coronagraph images, which consists of a <span class="hlt">bright</span> leading edge and a dark cavity, commonly believed to be a flux rope.</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> </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/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('https://images.nasa.gov/#/details-PIA18300.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA18300.html"><span>Little <span class="hlt">Bright</span> Spot</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2015-01-12</p> <p>A <span class="hlt">bright</span> 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 <span class="hlt">brightness</span>. Material that is newly excavated from below the moon's surface and tossed across the surface by a cratering event, appears <span class="hlt">bright</span>. 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170002034&hterms=algorithm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dalgorithm','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170002034&hterms=algorithm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dalgorithm"><span>A New Operational Snow Retrieval Algorithm Applied to Historical AMSR-E <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tedesco, Marco; Jeyaratnam, Jeyavinoth</p> <p>2016-01-01</p> <p>Snow is a key element of the water and energy cycles and the knowledge of spatio-temporal distribution of snow depth and snow water equivalent (SWE) is fundamental for hydrological and climatological applications. SWE and snow depth estimates can be obtained from spaceborne microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> at global scale and high temporal resolution (daily). In this regard, the data recorded by the Advanced Microwave Scanning Radiometer-Earth Orbiting System (EOS) (AMSR-E) onboard the National Aeronautics and Space Administration's (NASA) AQUA spacecraft have been used to generate operational estimates of SWE and snow depth, complementing estimates generated with other microwave sensors flying on other platforms. In this study, we report the results concerning the development and assessment of a new operational algorithm applied to historical AMSR-E data. The new algorithm here proposed makes use of climatological data, electromagnetic modeling and artificial neural networks for estimating snow depth as well as a spatio-temporal dynamic density scheme to convert snow depth to SWE. The outputs of the new algorithm are compared with those of the current AMSR-E operational algorithm as well as in-situ measurements and other operational snow products, specifically the Canadian Meteorological Center (CMC) and GlobSnow datasets. Our results show that the AMSR-E algorithm here proposed generally performs better than the operational one and addresses some major issues identified in the spatial distribution of snow depth fields associated with the evolution of effective grain size.</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('http://hdl.handle.net/2060/20110015546','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110015546"><span>Earth as an Extrasolar Planet: Earth Model Validation Using EPOXI Earth <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>Robinson, Tyler D.; Meadows, Victoria S.; Crisp, David; Deming, Drake; A'Hearn, Michael F.; Charbonneau, David; Livengood, Timothy A.; Seager, Sara; Barry, Richard; Hearty, Thomas; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20110015546'); toggleEditAbsImage('author_20110015546_show'); toggleEditAbsImage('author_20110015546_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20110015546_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20110015546_hide"></p> <p>2011-01-01</p> <p>The EPOXI Discovery Mission of Opportunity reused the Deep Impact flyby spacecraft to obtain spatially and temporally resolved visible photometric and moderate resolution near-infrared (NIR) spectroscopic <span class="hlt">observations</span> of Earth. These remote <span class="hlt">observations</span> provide a rigorous validation of whole disk Earth model simulations used to better under- stand remotely detectable extrasolar planet characteristics. We have used these data to upgrade, correct, and validate the NASA Astrobiology Institute s Virtual Planetary Laboratory three-dimensional line-by-line, multiple-scattering spectral Earth model (Tinetti et al., 2006a,b). This comprehensive model now includes specular reflectance from the ocean and explicitly includes atmospheric effects such as Rayleigh scattering, gas absorption, and <span class="hlt">temperature</span> structure. We have used this model to generate spatially and temporally resolved synthetic spectra and images of Earth for the dates of EPOXI <span class="hlt">observation</span>. Model parameters were varied to yield an optimum fit to the data. We found that a minimum spatial resolution of approx.100 pixels on the visible disk, and four categories of water clouds, which were defined using <span class="hlt">observed</span> cloud positions and optical thicknesses, were needed to yield acceptable fits. The validated model provides a simultaneous fit to the Earth s lightcurve, absolute <span class="hlt">brightness</span>, and spectral data, with a root-mean-square error of typically less than 3% for the multiwavelength lightcurves, and residuals of approx.10% for the absolute <span class="hlt">brightness</span> throughout the visible and NIR spectral range. We extend our validation into the mid-infrared by comparing the model to high spectral resolution <span class="hlt">observations</span> of Earth from the Atmospheric Infrared Sounder, obtaining a fit with residuals of approx.7%, and <span class="hlt">temperature</span> errors of less than 1K in the atmospheric window. For the purpose of understanding the <span class="hlt">observable</span> characteristics of the distant Earth at arbitrary viewing geometry and <span class="hlt">observing</span> cadence, our validated</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AsBio..11..393R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AsBio..11..393R"><span>Earth as an Extrasolar Planet: Earth Model Validation Using EPOXI Earth <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>Robinson, Tyler D.; Meadows, Victoria S.; Crisp, David; Deming, Drake; A'Hearn, Michael F.; Charbonneau, David; Livengood, Timothy A.; Seager, Sara; Barry, Richard K.; Hearty, Thomas; Hewagama, Tilak; Lisse, Carey M.; McFadden, Lucy A.; Wellnitz, Dennis D.</p> <p>2011-06-01</p> <p>The EPOXI Discovery Mission of Opportunity reused the Deep Impact flyby spacecraft to obtain spatially and temporally resolved visible photometric and moderate resolution near-infrared (NIR) spectroscopic <span class="hlt">observations</span> of Earth. These remote <span class="hlt">observations</span> provide a rigorous validation of whole-disk Earth model simulations used to better understand remotely detectable extrasolar planet characteristics. We have used these data to upgrade, correct, and validate the NASA Astrobiology Institute's Virtual Planetary Laboratory three-dimensional line-by-line, multiple-scattering spectral Earth model. This comprehensive model now includes specular reflectance from the ocean and explicitly includes atmospheric effects such as Rayleigh scattering, gas absorption, and <span class="hlt">temperature</span> structure. We have used this model to generate spatially and temporally resolved synthetic spectra and images of Earth for the dates of EPOXI <span class="hlt">observation</span>. Model parameters were varied to yield an optimum fit to the data. We found that a minimum spatial resolution of ∼100 pixels on the visible disk, and four categories of water clouds, which were defined by using <span class="hlt">observed</span> cloud positions and optical thicknesses, were needed to yield acceptable fits. The validated model provides a simultaneous fit to Earth's lightcurve, absolute <span class="hlt">brightness</span>, and spectral data, with a root-mean-square (RMS) error of typically less than 3% for the multiwavelength lightcurves and residuals of ∼10% for the absolute <span class="hlt">brightness</span> throughout the visible and NIR spectral range. We have extended our validation into the mid-infrared by comparing the model to high spectral resolution <span class="hlt">observations</span> of Earth from the Atmospheric Infrared Sounder, obtaining a fit with residuals of ∼7% and <span class="hlt">brightness</span> <span class="hlt">temperature</span> errors of less than 1 K in the atmospheric window. For the purpose of understanding the <span class="hlt">observable</span> characteristics of the distant Earth at arbitrary viewing geometry and <span class="hlt">observing</span> cadence, our validated forward model can be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21631250','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21631250"><span>Earth as an extrasolar planet: Earth model validation using EPOXI earth <span class="hlt">observations</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Robinson, Tyler D; Meadows, Victoria S; Crisp, David; Deming, Drake; A'hearn, Michael F; Charbonneau, David; Livengood, Timothy A; Seager, Sara; Barry, Richard K; Hearty, Thomas; Hewagama, Tilak; Lisse, Carey M; McFadden, Lucy A; Wellnitz, Dennis D</p> <p>2011-06-01</p> <p>The EPOXI Discovery Mission of Opportunity reused the Deep Impact flyby spacecraft to obtain spatially and temporally resolved visible photometric and moderate resolution near-infrared (NIR) spectroscopic <span class="hlt">observations</span> of Earth. These remote <span class="hlt">observations</span> provide a rigorous validation of whole-disk Earth model simulations used to better understand remotely detectable extrasolar planet characteristics. We have used these data to upgrade, correct, and validate the NASA Astrobiology Institute's Virtual Planetary Laboratory three-dimensional line-by-line, multiple-scattering spectral Earth model. This comprehensive model now includes specular reflectance from the ocean and explicitly includes atmospheric effects such as Rayleigh scattering, gas absorption, and <span class="hlt">temperature</span> structure. We have used this model to generate spatially and temporally resolved synthetic spectra and images of Earth for the dates of EPOXI <span class="hlt">observation</span>. Model parameters were varied to yield an optimum fit to the data. We found that a minimum spatial resolution of ∼100 pixels on the visible disk, and four categories of water clouds, which were defined by using <span class="hlt">observed</span> cloud positions and optical thicknesses, were needed to yield acceptable fits. The validated model provides a simultaneous fit to Earth's lightcurve, absolute <span class="hlt">brightness</span>, and spectral data, with a root-mean-square (RMS) error of typically less than 3% for the multiwavelength lightcurves and residuals of ∼10% for the absolute <span class="hlt">brightness</span> throughout the visible and NIR spectral range. We have extended our validation into the mid-infrared by comparing the model to high spectral resolution <span class="hlt">observations</span> of Earth from the Atmospheric Infrared Sounder, obtaining a fit with residuals of ∼7% and <span class="hlt">brightness</span> <span class="hlt">temperature</span> errors of less than 1 K in the atmospheric window. For the purpose of understanding the <span class="hlt">observable</span> characteristics of the distant Earth at arbitrary viewing geometry and <span class="hlt">observing</span> cadence, our validated forward model can be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3133830','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3133830"><span>Earth as an Extrasolar Planet: Earth Model Validation Using EPOXI Earth <span class="hlt">Observations</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Meadows, Victoria S.; Crisp, David; Deming, Drake; A'Hearn, Michael F.; Charbonneau, David; Livengood, Timothy A.; Seager, Sara; Barry, Richard K.; Hearty, Thomas; Hewagama, Tilak; Lisse, Carey M.; McFadden, Lucy A.; Wellnitz, Dennis D.</p> <p>2011-01-01</p> <p>Abstract The EPOXI Discovery Mission of Opportunity reused the Deep Impact flyby spacecraft to obtain spatially and temporally resolved visible photometric and moderate resolution near-infrared (NIR) spectroscopic <span class="hlt">observations</span> of Earth. These remote <span class="hlt">observations</span> provide a rigorous validation of whole-disk Earth model simulations used to better understand remotely detectable extrasolar planet characteristics. We have used these data to upgrade, correct, and validate the NASA Astrobiology Institute's Virtual Planetary Laboratory three-dimensional line-by-line, multiple-scattering spectral Earth model. This comprehensive model now includes specular reflectance from the ocean and explicitly includes atmospheric effects such as Rayleigh scattering, gas absorption, and <span class="hlt">temperature</span> structure. We have used this model to generate spatially and temporally resolved synthetic spectra and images of Earth for the dates of EPOXI <span class="hlt">observation</span>. Model parameters were varied to yield an optimum fit to the data. We found that a minimum spatial resolution of ∼100 pixels on the visible disk, and four categories of water clouds, which were defined by using <span class="hlt">observed</span> cloud positions and optical thicknesses, were needed to yield acceptable fits. The validated model provides a simultaneous fit to Earth's lightcurve, absolute <span class="hlt">brightness</span>, and spectral data, with a root-mean-square (RMS) error of typically less than 3% for the multiwavelength lightcurves and residuals of ∼10% for the absolute <span class="hlt">brightness</span> throughout the visible and NIR spectral range. We have extended our validation into the mid-infrared by comparing the model to high spectral resolution <span class="hlt">observations</span> of Earth from the Atmospheric Infrared Sounder, obtaining a fit with residuals of ∼7% and <span class="hlt">brightness</span> <span class="hlt">temperature</span> errors of less than 1 K in the atmospheric window. For the purpose of understanding the <span class="hlt">observable</span> characteristics of the distant Earth at arbitrary viewing geometry and <span class="hlt">observing</span> cadence, our validated forward</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/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://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> <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('https://pubs.er.usgs.gov/publication/70027628','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027628"><span>Planetary science: A 5-micron-<span class="hlt">bright</span> spot on Titan: Evidence for surface diversity</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>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.</p> <p>2005-01-01</p> <p><span class="hlt">Observations</span> from the Cassini Visual and Infrared Mapping Spectrometer show an anomalously <span class="hlt">bright</span> spot on Titan located at 80??W and 20??S. This area is <span class="hlt">bright</span> in reflected tight at all <span class="hlt">observed</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..557..740E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..557..740E"><span>Exploiting the synergy between SMAP and SMOS to improve <span class="hlt">brightness</span> <span class="hlt">temperature</span> simulations and soil moisture retrievals in arid regions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ebrahimi, Mohsen; Alavipanah, Seyed Kazem; Hamzeh, Saeid; Amiraslani, Farshad; Neysani Samany, Najmeh; Wigneron, Jean-Pierre</p> <p>2018-02-01</p> <p>The objective of this study was to exploit the synergy between SMOS and SMAP based on vegetation optical depth (VOD) to improve <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TB) simulations and land surface soil moisture (SM) retrievals in arid regions of the world. In the current operational algorithm of SMAP (level 2), vegetation water content (VWC) is considered as a proxy to compute VOD which is calculated by an empirical conversion function of NDVI. Avoiding the empirical estimation of VOD, the SMOS algorithm is used to retrieve simultaneously SM and VOD from TB <span class="hlt">observations</span>. The present study attempted to improve SMAP TB simulations and SM retrievals by benefiting from the advantages of the SMOS (L-MEB) algorithm. This was achieved by using a synergy method based on replacing the default value of SMAP VOD with the retrieved value of VOD from the SMOS multi angular and bi-polarization <span class="hlt">observations</span> of TB. The insitu SM measurements, used as reference SM in this study, were obtained from the International Soil Moisture Network (ISMN) over 180 stations located in arid regions of the world. Furthermore, four stations were randomly selected to analyze the temporal variations in VOD and SM. Results of the synergy method showed that the accuracy of the TB simulations and SM retrievals was respectively improved at 144 and 124 stations (out of a total of 180 stations) in terms of coefficient of determination (R2) and unbiased root mean squared error (UbRMSE). Analyzing the temporal variations in VOD showed that the SMOS VOD, conversely to the SMAP VOD, can better illustrate the presence of herbaceous plants and may be a better indicator of the seasonal changes in the vegetation density and biomass over the year.</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> </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('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('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 surface <span class="hlt">temperature</span>, with cooler water along the coast and warmer water offshore. The sunglint <span class="hlt">brightness</span> reveals water surface 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 <span class="hlt">temperature</span> is the cloud pattern. Advection within the atmosphere as a result of warming at the sea surface 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('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://hdl.handle.net/2060/20140008866','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140008866"><span>A Coronal Hole Jet <span class="hlt">Observed</span> with Hinode and the Solar Dynamics Observatory</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Young, Peter H.; Muglach, Karin</p> <p>2014-01-01</p> <p>A small blowout jet was <span class="hlt">observed</span> at the boundary of the south coronal hole on 2011 February 8 at around 21:00 UT. Images from the Atmospheric Imaging Assembly (AIA) on board the Solar Dynamics Observatory (SDO) revealed an expanding loop rising from one footpoint of a compact, bipolar <span class="hlt">bright</span> point. Magnetograms from the Helioseismic Magnetic Imager (HMI) on board SDO showed that the jet was triggered by the cancelation of a parasitic positive polarity feature near the negative pole of the <span class="hlt">bright</span> point. The jet emission was present for 25 mins and it extended 30 Mm from the <span class="hlt">bright</span> point. Spectra from the EUV Imaging Spectrometer on board Hinode yielded a <span class="hlt">temperature</span> and density of 1.6 MK and 0.9-1.7 × 10( exp 8) cu cm for the ejected plasma. Line-of-sight velocities reached up to 250 km/s. The density of the <span class="hlt">bright</span> point was 7.6 × 10(exp 8) cu cm, and the peak of the <span class="hlt">bright</span> point's emission measure occurred at 1.3 MK, with no plasma above 3 MK.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRD..123.1065B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRD..123.1065B"><span>Consistency Between Convection Allowing Model Output and Passive Microwave 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>Bytheway, J. L.; Kummerow, C. D.</p> <p>2018-01-01</p> <p><span class="hlt">Observations</span> from the Global Precipitation Measurement (GPM) core satellite were used along with precipitation forecasts from the High Resolution Rapid Refresh (HRRR) model to assess and interpret differences between <span class="hlt">observed</span> and modeled storms. Using a feature-based approach, precipitating objects were identified in both the National Centers for Environmental Prediction Stage IV multisensor precipitation product and HRRR forecast at lead times of 1, 2, and 3 h at valid times corresponding to GPM overpasses. Precipitating objects were selected for further study if (a) the <span class="hlt">observed</span> feature occurred entirely within the swath of the GPM Microwave Imager (GMI) and (b) the HRRR model predicted it at all three forecast lead times. Output from the HRRR model was used to simulate microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (Tbs), which were compared to those <span class="hlt">observed</span> by the GMI. Simulated Tbs were found to have biases at both the warm and cold ends of the distribution, corresponding to the stratiform/anvil and convective areas of the storms, respectively. Several experiments altered both the simulation microphysics and hydrometeor classification in order to evaluate potential shortcomings in the model's representation of precipitating clouds. In general, inconsistencies between <span class="hlt">observed</span> and simulated <span class="hlt">brightness</span> <span class="hlt">temperatures</span> were most improved when transferring snow water content to supercooled liquid hydrometeor classes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110023545','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110023545"><span>Thermal Properties of A Solar Coronal Cavity <span class="hlt">Observed</span> with the X-Ray Telescope on Hinode</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reeves, Katherine K.; Gibson, Sarah E.; Kucera, Theresa A.; Hudson, Hugh S.; Kano, Ryouhei</p> <p>2011-01-01</p> <p>Coronal cavities are voids in coronal emission often <span class="hlt">observed</span> above high latitude filament channels. Sometimes, these cavities have areas of <span class="hlt">bright</span> X-ray emission in their centers. In this study, we use data from the X-ray Telescope (XRT) on the Hinode satellite to examine the thermal emission properties of a cavity <span class="hlt">observed</span> during July 2008 that contains <span class="hlt">bright</span> X-ray emission in its center. Using ratios of XRT filters, we find evidence for elevated <span class="hlt">temperatures</span> in the cavity center. The area of elevated <span class="hlt">temperature</span> evolves from a ring-shaped structure at the beginning of the <span class="hlt">observation</span>, to an elongated structure two days later, finally appearing as a compact round source four days after the initial <span class="hlt">observation</span>. We use a morphological model to fit the cavity emission, and find that a uniform structure running through the cavity does not fit the <span class="hlt">observations</span> well. Instead, the <span class="hlt">observations</span> are reproduced by modeling several short cylindrical cavity "cores" with different parameters on different days. These changing core parameters may be due to some <span class="hlt">observed</span> activity heating different parts of the cavity core at different times. We find that core <span class="hlt">temperatures</span> of 1.75 MK, 1.7 MK and 2.0 MK (for July 19, July 21 and July 23, respectively) in the model lead to structures that are consistent with the data, and that line-of-sight effects serve to lower the effective <span class="hlt">temperature</span> derived from the filter ratio.</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('https://www.ncbi.nlm.nih.gov/pubmed/23798032','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23798032"><span><span class="hlt">Brightness</span> and transparency in the early visual cortex.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Salmela, Viljami R; Vanni, Simo</p> <p>2013-06-24</p> <p>Several psychophysical studies have shown that transparency can have drastic effects on <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> perception. While single cell recordings suggest that surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> and transparency. To dissociate the blood oxygen level-dependent (BOLD) response to <span class="hlt">brightness</span> from the response to local border contrast and mean luminance, we used variants of White's <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> modulation to the target. The target luminance was constant or modulated in counterphase to null <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> than for stimuli with constant <span class="hlt">brightness</span>. In addition, the responses were stronger for transparent than for opaque stimuli, but there was more individual variation. No interaction between <span class="hlt">brightness</span> and transparency was found. The results show that the early visual areas V1-V3 are sensitive to surface <span class="hlt">brightness</span> and transparency and suggest that <span class="hlt">brightness</span> and transparency are represented separately.</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/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 surface and near-surface <span class="hlt">temperatures</span>. These <span class="hlt">temperatures</span> 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 surface properties and can be utilized to distinguish rocky from porous surfaces. 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> <span class="hlt">temperatures</span> as a function of wavelength, and implications for <span class="hlt">temperature</span> and physical state of the surface and near-surface. [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/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> <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('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://www.ncbi.nlm.nih.gov/pubmed/12638695','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12638695"><span>Effects of dim or <span class="hlt">bright</span>-light exposure during the daytime on human gastrointestinal activity.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sone, Yoshiaki; Hyun, Ki-Ja; Nishimura, Shinya; Lee, Young-Ah; Tokura, Hiromi</p> <p>2003-01-01</p> <p>On the basis of our previous findings that <span class="hlt">bright</span>-light exposure during the daytime has profound influence on physiological parameters such as melatonin secretion and tympanic <span class="hlt">temperature</span> in humans, we proposed the hypothesis that <span class="hlt">bright</span> vs. dim light-exposure during the daytime has a different influence on the activity of the digestive system via the endocrine and/or autonomic nervous system. To examine this hypothesis, we conducted a series of counterbalanced experiments in which subjects stayed the daytime (7:00 to 15:00h) under either a dim (80 lux) or <span class="hlt">bright</span> (5,000 lux) light condition. We measured gastrointestinal activity using a breath hydrogen (indicative of carbohydrate malabsorption) and an electrogastrography (EGG, indicative of gastric myoelectric activity) test. The results showed the postprandial breath hydrogen excretion during the following nighttime period after daytime exposure to the dim-light condition was significantly higher than under the <span class="hlt">bright</span>-light condition (p < 0.05). In addition, the spectrum total power of the EGG recorded after taking the evening meal was significantly lower for the dim than <span class="hlt">bright</span>-light condition (p < 0.05). These results support our hypothesis and indicate that dim-light exposure during the daytime suppresses the digestion of the evening meal, resulting in malabsorption of dietary carbohydrates in it.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H21I..02A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H21I..02A"><span>Multi-scale assimilation of remotely sensed snow <span class="hlt">observations</span> for hydrologic estimation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Andreadis, K.; Lettenmaier, D.</p> <p>2008-12-01</p> <p>Data assimilation provides a framework for optimally merging model predictions and remote sensing <span class="hlt">observations</span> of snow properties (snow cover extent, water equivalent, grain size, melt state), ideally overcoming limitations of both. A synthetic twin experiment is used to evaluate a data assimilation system that would ingest remotely sensed <span class="hlt">observations</span> from passive microwave and visible wavelength sensors (<span class="hlt">brightness</span> <span class="hlt">temperature</span> and snow cover extent derived products, respectively) with the objective of estimating snow water equivalent. Two data assimilation techniques are used, the Ensemble Kalman filter and the Ensemble Multiscale Kalman filter (EnMKF). One of the challenges inherent in such a data assimilation system is the discrepancy in spatial scales between the different types of snow-related <span class="hlt">observations</span>. The EnMKF represents the sample model error covariance with a tree that relates the system state variables at different locations and scales through a set of parent-child relationships. This provides an attractive framework to efficiently assimilate <span class="hlt">observations</span> at different spatial scales. This study provides a first assessment of the feasibility of a system that would assimilate <span class="hlt">observations</span> from multiple sensors (MODIS snow cover and AMSR-E <span class="hlt">brightness</span> <span class="hlt">temperatures</span>) and at different spatial scales for snow water equivalent estimation. The relative value of the different types of <span class="hlt">observations</span> is examined. Additionally, the error characteristics of both model and <span class="hlt">observations</span> are discussed.</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('https://ntrs.nasa.gov/search.jsp?R=19900061879&hterms=solar+two&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsolar%2Btwo','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900061879&hterms=solar+two&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsolar%2Btwo"><span>Multifrequency <span class="hlt">observations</span> of a solar microwave burst with two-dimensional spatial resolution</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gary, Dale E.; Hurford, G. J.</p> <p>1990-01-01</p> <p>Frequency-agile interferometry <span class="hlt">observations</span> using three baselines and the technique of frequency synthesis were used to obtain two-dimensional positions of multiple microwave sources at several frequency ranges in a solar flare. Source size and <span class="hlt">brightness</span> <span class="hlt">temperature</span> spectra were obtained near the peak of the burst. The size spectrum shows that the source size decreases rapidly with increasing frequency, but the <span class="hlt">brightness</span> <span class="hlt">temperature</span> spectrum can be well-fitted by gyrosynchrotron emission from a nonthermal distribution of electrons with power-law index of 4.8. The spatial structure of the burst showed several characteristics in common with primary/secondary bursts discussed by Nakajima et al. (1985). A source of coherent plasma emission at low frequencies is found near the secondary gyrosynchrotron source, associated with the leader spots of the active region.</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('http://adsabs.harvard.edu/abs/2008PhDT.......108H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008PhDT.......108H"><span>Ground and satellite-based remote sensing of mineral dust using AERI spectra and MODIS thermal infrared window <span class="hlt">brightness</span> <span class="hlt">temperatures</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hansell, Richard Allen, Jr.</p> <p></p> <p>The radiative effects of dust aerosol on our climate system have yet to be fully understood and remain a topic of contemporary research. To investigate these effects, detection/retrieval methods for dust events over major dust outbreak and transport areas have been developed using satellite and ground-based approaches. To this end, both the shortwave and longwave surface radiative forcing of dust aerosol were investigated. The ground-based remote sensing approach uses the Atmospheric Emitted Radiance Interferometer <span class="hlt">brightness</span> <span class="hlt">temperature</span> spectra to detect mineral dust events and to retrieve their properties. Taking advantage of the high spectral resolution of the AERI instrument, absorptive differences in prescribed thermal IR window sub-band channels were exploited to differentiate dust from cirrus clouds. AERI data collected during the UAE2 at Al-Ain UAE was employed for dust retrieval. Assuming a specified dust composition model a priori and using the light scattering programs of T-matrix and the finite difference time domain methods for oblate spheroids and hexagonal plates, respectively, dust optical depths have been retrieved and compared to those inferred from a collocated and coincident AERONET sun-photometer dataset. The retrieved optical depths were then used to determine the dust longwave surface forcing during the UAE2. Likewise, dust shortwave surface forcing is investigated employing a differential technique from previous field studies. The satellite-based approach uses MODIS thermal infrared <span class="hlt">brightness</span> <span class="hlt">temperature</span> window data for the simultaneous detection/separation of mineral dust and cirrus clouds. Based on the spectral variability of dust emissivity at the 3.75, 8.6, 11 and 12 mum wavelengths, the D*-parameter, BTD-slope and BTD3-11 tests are combined to identify dust and cirrus. MODIS data for the three dust-laden scenes have been analyzed to demonstrate the effectiveness of this detection/separation method. Detected daytime dust and cloud</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('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('http://www.ars.usda.gov/research/publications/publication/?seqNo115=263655','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=263655"><span>Using SMOS <span class="hlt">observations</span> in the development of the SMAP level 4 surface and root-zone soil moisture project</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 and Ocean Salinity (SMOS; [1]) 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. Along with these <span class="hlt">brightness</span> <span class="hlt">temperature</span> <span class="hlt">observations</span>, ESA also disseminates retrievals of surface soil moisture that are derived ...</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://hdl.handle.net/2060/20020067408','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020067408"><span>Real Time Monitoring of Flooding from Microwave 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>Galantowicz, John F.; Frey, H. (Technical Monitor)</p> <p>2001-01-01</p> <p>In this report, we review the progress to date including results from data analyses and present a schedule of milestones for the remainder of the project. We discuss the processing of flood extent data and SSM/I <span class="hlt">brightness</span> <span class="hlt">temperature</span> data for the 1993 Midwest Flood. We present preliminary results from the derivation of open water fraction from <span class="hlt">brightness</span> <span class="hlt">temperatures</span>.</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 surface instead of a hyperboloid surface 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/2006PhDT.......189A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhDT.......189A"><span>Coronal Properties of X-ray <span class="hlt">bright</span> stars in young associations: abundances, <span class="hlt">temperatures</span> and variability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Argiroffi, Costanza</p> <p>2006-03-01</p> <p>In this work I have investigated open issues related to the X-ray radiation from young stars, including heating mechanisms of the emitting plasma, its chemical composition, and possible effects due to circumstellar accretion disks. To this aim, I have analyzed <span class="hlt">observations</span> of young nearby stars taken with the X-ray observatories XMM-Newton and Chandra. For a detailed study of the characteristics of the X-ray emitting plasma, I have selected two X-ray <span class="hlt">bright</span> young stars, TWA 5 and PZ Tel, for which high-resolution X-ray spectroscopy was achievable, and two regions of the young stellar association Upper Scorpius (USco), for which X-ray images and medium-resolution spectra of individual sources were obtained. TWA 5 is a 10 Myr old star in the TW Hydrae association, which is still accreting material from its circumstellar envelope, while PZ Tel is a ? 12 Myr star in the beta Pictoris moving group, which already dissipated its circumstellar disk. The different evolutionary stages of these two stars allow to probe whether X-ray emission is produced, or affected, by accretion processes. High-resolution X-ray spectra of TWA 5 and PZ Tel were gathered with the grating spectrometers on board XMM-Newton and Chandra, respectively. From the measurements of individual emission line fluxes in their X-ray spectra, I have derived emission measure distributions vs. <span class="hlt">temperature</span>, abundances, and electron densities of the X-ray emitting plasma. I have found that, in spite of their different evolutionary status, hot (T ? 10 MK) plasma is the main responsible for the <span class="hlt">observed</span> X-ray emission of both stars. The hot plasma on TWA 5 displays peculiar element abundances with respect to the solar photospheric composition with Ne/Fe ? 10(Ne/Fe), while the coronal plasma on PZ Tel shows Ne/Fe ? 3(Ne/Fe). To explain the strong Fe underabundance (? 0.1 Fe) and the extremely high Ne/Fe ratio of TWA 5 I have considered three different scenarios: (1) coronal plasma may be affected by selective</p> </li> <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) surface during the initial room-<span class="hlt">temperature</span> 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 <span class="hlt">temperature</span>. 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) surface 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/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> <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('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5937273','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5937273"><span>Evaluating soil moisture retrievals from ESA’s SMOS and NASA’s SMAP <span class="hlt">brightness</span> <span class="hlt">temperature</span> datasets</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Al-Yaari, A.; Wigneron, J.-P.; Kerr, Y.; Rodriguez-Fernandez, N.; O’Neill, P. E.; Jackson, T. J.; De Lannoy, G.J.M.; Al Bitar, A; Mialon, A.; Richaume, P.; Walker, JP; Mahmoodi, A.; Yueh, S.</p> <p>2018-01-01</p> <p>Two satellites are currently monitoring surface soil moisture (SM) using L-band <span class="hlt">observations</span>: SMOS (Soil Moisture and Ocean Salinity), a joint ESA (European Space Agency), CNES (Centre national d’études spatiales), and CDTI (the Spanish government agency with responsibility for space) satellite launched on November 2, 2009 and SMAP (Soil Moisture Active Passive), a National Aeronautics and Space Administration (NASA) satellite successfully launched in January 2015. In this study, we used a multilinear regression approach to retrieve SM from SMAP data to create a global dataset of SM, which is consistent with SM data retrieved from SMOS. This was achieved by calibrating coefficients of the regression model using the CATDS (Centre Aval de Traitement des Données) SMOS Level 3 SM and the horizontally and vertically polarized <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TB) at 40° incidence angle, over the 2013 – 2014 period. Next, this model was applied to SMAP L3 TB data from Apr 2015 to Jul 2016. The retrieved SM from SMAP (referred to here as SMAP_Reg) was compared to: (i) the operational SMAP L3 SM (SMAP_SCA), retrieved using the baseline Single Channel retrieval Algorithm (SCA); and (ii) the operational SMOSL3 SM, derived from the multiangular inversion of the L-MEB model (L-MEB algorithm) (SMOSL3). This inter-comparison was made against in situ soil moisture measurements from more than 400 sites spread over the globe, which are used here as a reference soil moisture dataset. The in situ <span class="hlt">observations</span> were obtained from the International Soil Moisture Network (ISMN; https://ismn.geo.tuwien.ac.at/) in North of America (PBO_H2O, SCAN, SNOTEL, iRON, and USCRN), in Australia (Oznet), Africa (DAHRA), and in Europe (REMEDHUS, SMOSMANIA, FMI, and RSMN). The agreement was analyzed in terms of four classical statistical criteria: Root Mean Squared Error (RMSE), Bias, Unbiased RMSE (UnbRMSE), and correlation coefficient (R). Results of the comparison of these various products with in</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/2060/19990099126','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990099126"><span>Stratospheric <span class="hlt">Temperature</span> Changes: <span class="hlt">Observations</span> and Model Simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ramaswamy, V.; Chanin, M.-L.; Angell, J.; Barnett, J.; Gaffen, D.; Gelman, M.; Keckhut, P.; Koshelkov, Y.; Labitzke, K.; Lin, J.-J. R.</p> <p>1999-01-01</p> <p>This paper reviews <span class="hlt">observations</span> of stratospheric <span class="hlt">temperatures</span> that have been made over a period of several decades. Those <span class="hlt">observed</span> <span class="hlt">temperatures</span> have been used to assess variations and trends in stratospheric <span class="hlt">temperatures</span>. A wide range of <span class="hlt">observation</span> datasets have been used, comprising measurements by radiosonde (1940s to the present), satellite (1979 - present), lidar (1979 - present) and rocketsonde (periods varying with location, but most terminating by about the mid-1990s). In addition, trends have also been assessed from meteorological analyses, based on radiosonde and/or satellite data, and products based on assimilating <span class="hlt">observations</span> into a general circulation model. Radiosonde and satellite data indicate a cooling trend of the annual-mean lower stratosphere since about 1980. Over the period 1979-1994, the trend is 0.6K/decade. For the period prior to 1980, the radiosonde data exhibit a substantially weaker long-term cooling trend. In the northern hemisphere, the cooling trend is about 0.75K/decade in the lower stratosphere, with a reduction in the cooling in mid-stratosphere (near 35 km), and increased cooling in the upper stratosphere (approximately 2 K per decade at 50 km). Model simulations indicate that the depletion of lower stratospheric ozone is the dominant factor in the <span class="hlt">observed</span> lower stratospheric cooling. In the middle and upper stratosphere both the well-mixed greenhouse gases (such as CO) and ozone changes contribute in an important manner to the cooling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JGRD..117.6207H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JGRD..117.6207H"><span>Estimating effective particle size of tropical deep convective clouds with a look-up table method using satellite measurements of <span class="hlt">brightness</span> <span class="hlt">temperature</span> differences</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hong, Gang; Minnis, Patrick; Doelling, David; Ayers, J. Kirk; Sun-Mack, Szedung</p> <p>2012-03-01</p> <p>A method for estimating effective ice particle radius Re at the tops of tropical deep convective clouds (DCC) is developed on the basis of precomputed look-up tables (LUTs) of <span class="hlt">brightness</span> <span class="hlt">temperature</span> differences (BTDs) between the 3.7 and 11.0 μm bands. A combination of discrete ordinates radiative transfer and correlated k distribution programs, which account for the multiple scattering and monochromatic molecular absorption in the atmosphere, is utilized to compute the LUTs as functions of solar zenith angle, satellite zenith angle, relative azimuth angle, Re, cloud top <span class="hlt">temperature</span> (CTT), and cloud visible optical thickness τ. The LUT-estimated DCC Re agrees well with the cloud retrievals of the Moderate Resolution Imaging Spectroradiometer (MODIS) for the NASA Clouds and Earth's Radiant Energy System with a correlation coefficient of 0.988 and differences of less than 10%. The LUTs are applied to 1 year of measurements taken from MODIS aboard Aqua in 2007 to estimate DCC Re and are compared to a similar quantity from CloudSat over the region bounded by 140°E, 180°E, 0°N, and 20°N in the Western Pacific Warm Pool. The estimated DCC Re values are mainly concentrated in the range of 25-45 μm and decrease with CTT. Matching the LUT-estimated Re with ice cloud Re retrieved by CloudSat, it is found that the ice cloud τ values from DCC top to the vertical location where LUT-estimated Re is located at the CloudSat-retrieved Re profile are mostly less than 2.5 with a mean value of about 1.3. Changes in the DCC τ can result in differences of less than 10% for Re estimated from LUTs. The LUTs of 0.65 μm bidirectional reflectance distribution function (BRDF) are built as functions of viewing geometry and column amount of ozone above upper troposphere. The 0.65 μm BRDF can eliminate some noncore portions of the DCCs detected using only 11 μm <span class="hlt">brightness</span> <span class="hlt">temperature</span> thresholds, which result in a mean difference of only 0.6 μm for DCC Re estimated from BTD LUTs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JAMES...8.1453B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JAMES...8.1453B"><span>Validation of a weather forecast model at radiance level against satellite <span class="hlt">observations</span> allowing quantification of <span class="hlt">temperature</span>, humidity, and cloud-related biases</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bani Shahabadi, Maziar; Huang, Yi; Garand, Louis; Heilliette, Sylvain; Yang, Ping</p> <p>2016-09-01</p> <p>An established radiative transfer model (RTM) is adapted for simulating all-sky infrared radiance spectra from the Canadian Global Environmental Multiscale (GEM) model in order to validate its forecasts at the radiance level against Atmospheric InfraRed Sounder (AIRS) <span class="hlt">observations</span>. Synthetic spectra are generated for 2 months from short-term (3-9 h) GEM forecasts. The RTM uses a monthly climatological land surface emissivity/reflectivity atlas. An updated ice particle optical property library was introduced for cloudy radiance calculations. Forward model <span class="hlt">brightness</span> <span class="hlt">temperature</span> (BT) biases are assessed to be of the order of ˜1 K for both clear-sky and overcast conditions. To quantify GEM forecast meteorological variables biases, spectral sensitivity kernels are generated and used to attribute radiance biases to surface and atmospheric <span class="hlt">temperatures</span>, atmospheric humidity, and clouds biases. The kernel method, supplemented with retrieved profiles based on AIRS <span class="hlt">observations</span> in collocation with a microwave sounder, achieves good closure in explaining clear-sky radiance biases, which are attributed mostly to surface <span class="hlt">temperature</span> and upper tropospheric water vapor biases. Cloudy-sky radiance biases are dominated by cloud-induced radiance biases. Prominent GEM biases are identified as: (1) too low surface <span class="hlt">temperature</span> over land, causing about -5 K bias in the atmospheric window region; (2) too high upper tropospheric water vapor, inducing about -3 K bias in the water vapor absorption band; (3) too few high clouds in the convective regions, generating about +10 K bias in window band and about +6 K bias in the water vapor band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950064055&hterms=Throughput+accounting&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DThroughput%2Baccounting','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950064055&hterms=Throughput+accounting&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DThroughput%2Baccounting"><span>The local metallicity-surface <span class="hlt">brightness</span> relationship in galactic disks</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ryder, Stuart D.</p> <p>1995-01-01</p> <p>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 <span class="hlt">observed</span> using a combination of multiaperture masks and conventional long-slit spectrophotometry. The oxygen abundances derived from the multiaperture mask <span class="hlt">observations</span> 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 <span class="hlt">observed</span> 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 <span class="hlt">brightness</span> at the radial distance of the H II region and confirm the existence of a local metallicity-surface <span class="hlt">brightness</span> 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 <span class="hlt">observational</span> 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</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> </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/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 surface. 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-surface 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=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=19820059157&hterms=k-12&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dk-12','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820059157&hterms=k-12&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dk-12"><span>On OMC-1 <span class="hlt">temperatures</span> determined from methyl cyanide <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>Hollis, J. M.</p> <p>1982-01-01</p> <p>An analysis is performed on the J(k) = 12(k)-11(k) and 13(k)-12(k) transitions of methyl cyanide detected by other investigators in the direction of OMC-1. The original interpretation of those <span class="hlt">observations</span> argues for the presence of two distinct <span class="hlt">temperature</span> regions or possibly a <span class="hlt">temperature</span> gradient within the cloud. The analysis presented here demonstrates that the <span class="hlt">observations</span> of these particular molecular transitions are consistent with a single methyl cyanide emission region with a source kinetic <span class="hlt">temperature</span> of 121.2 + or - 8.2 K and a molecular rotational <span class="hlt">temperature</span> of 16.6 + or - 1.8 K.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930005117','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930005117"><span><span class="hlt">Bright</span> crater outflows: Possible emplacement mechanisms</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chadwick, D. John; Schaber, Gerald G.; Strom, Robert G.; Duval, Darla M.</p> <p>1992-01-01</p> <p>Lobate features with a strong backscatter are associated with 43 percent of the impact craters cataloged in Magellan's cycle 1. Their apparent thinness and great lengths are consistent with a low-viscosity material. The longest outflow yet identified is about 600 km in length and flows from the 90-km-diameter crater Addams. There is strong evidence that the outflows are largely composed of impact melt, although the mechanisms of their emplacement are not clearly understood. High <span class="hlt">temperatures</span> and pressures of target rocks on Venus allow for more melt to be produced than on other terrestrial planets because lower shock pressures are required for melting. The percentage of impact craters with outflows increases with increasing crater diameter. The mean diameter of craters without outflows is 14.4 km, compared with 27.8 km for craters with outflows. No craters smaller than 3 km, 43 percent of craters in the 10- to 30-km-diameter range, and 90 percent in the 80- to 100-km-diameter range have associated <span class="hlt">bright</span> outflows. More melt is produced in the more energetic impact events that produce larger craters. However, three of the four largest craters have no outflows. We present four possible mechanisms for the emplacement of <span class="hlt">bright</span> outflows. We believe this 'shotgun' approach is justified because all four mechanisms may indeed have operated to some degree.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.474.4500W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.474.4500W"><span>ALMA and VLA <span class="hlt">observations</span> of the HD 141569 system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>White, Jacob Aaron; Boley, A. C.; MacGregor, M. A.; Hughes, A. M.; Wilner, D. J.</p> <p>2018-03-01</p> <p>We present VLA 9 mm (33 GHz) and archival ALMA 2.9 mm (103 GHz) <span class="hlt">observations</span> of the HD 141569 system. The VLA <span class="hlt">observations</span> achieve a resolution of 0.25 arcsec (˜28 au) and a sensitivity of 4.7 μJy beam- 1. We find (1) a 52 ± 5 μJy point source at the location of HD 141569A that shows potential variability, (2) the detected flux is contained within the SED-inferred central clearing of the disc meaning the spectral index of the dust disc is steeper than previously inferred, and (3) the M dwarf companions are also detected and variable. Previous lower resolution VLA <span class="hlt">observations</span> (semester 14A) found a higher flux density, interpreted as solely dust emission. When combined with ALMA <span class="hlt">observations</span>, the VLA 14A <span class="hlt">observations</span> suggested the spectral index, and grain size distribution of HD 141569's disc was shallow and an outlier among debris systems. Using archival ALMA <span class="hlt">observations</span> of HD 141569 at 0.87 and 2.9 mm, we find a dust spectral index of αmm = 1.81 ± 0.20. The VLA 16A flux corresponds to a <span class="hlt">brightness</span> <span class="hlt">temperature</span> of ˜5 × 106 K, suggesting strong non-disc emission is affecting the inferred grain properties. The VLA 16A flux density of the M2V companion HD 141569B is 149 ± 9 μJy, corresponding to a <span class="hlt">brightness</span> <span class="hlt">temperature</span> of ˜2 × 108 K and suggesting significant stellar variability when compared to the VLA14A <span class="hlt">observations</span>, which are smaller by a factor of ˜6.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1913997H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1913997H"><span>Integrating SMOS <span class="hlt">brightness</span> <span class="hlt">temperatures</span> with a new conceptual spatially distributed hydrological model for improving flood and drought predictions at large scale.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hostache, Renaud; Rains, Dominik; Chini, Marco; Lievens, Hans; Verhoest, Niko E. C.; Matgen, Patrick</p> <p>2017-04-01</p> <p>, SUPERFLEX is capable of predicting runoff, soil moisture, and SMOS-like <span class="hlt">brightness</span> <span class="hlt">temperature</span> time series. Such a model is traditionally calibrated using only discharge measurements. In this study we designed a multi-objective calibration procedure based on both discharge measurements and SMOS-derived <span class="hlt">brightness</span> <span class="hlt">temperature</span> <span class="hlt">observations</span> in order to evaluate the added value of remotely sensed soil moisture data in the calibration process. As a test case we set up the SUPERFLEX model for the large scale Murray-Darling catchment in Australia ( 1 Million km2). When compared to in situ soil moisture time series, model predictions show good agreement resulting in correlation coefficients exceeding 70 % and Root Mean Squared Errors below 1 %. When benchmarked with the physically based land surface model CLM, SUPERFLEX exhibits similar performance levels. By adapting the runoff routing function within the SUPERFLEX model, the predicted discharge results in a Nash Sutcliff Efficiency exceeding 0.7 over both the calibration and the validation periods.</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/1982AdSpR...2..161S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1982AdSpR...2..161S"><span>The use of stereoscopic satellite <span class="hlt">observation</span> in the determination of the emissivity of cirrus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Szejwach, G.; Sletten, T. N.; Hasler, A. F.</p> <p></p> <p>The feasibility of determining cirrus ``emissivity'' from combined stereoscopic and infrared satellite <span class="hlt">observations</span> in conjunction with radiosounding data is investigated for a particular case study. Simultaneous visible images obtained during SESAME-1979 from two geosynchronous GOES meteorological satellites were processed on the NASA/Goddard interactive system (AOIPS) and were used to determine the stereo cloud top height ZC as described by Hasler [1]. Iso-contours of radiances were outlined on the corresponding infrared image. Total <span class="hlt">brightness</span> <span class="hlt">temperature</span> TB and ground surface <span class="hlt">brightness</span> <span class="hlt">temperature</span> TS were inferred from the radiances. The special SESAME network of radiosoundings was used to determine the cloud top <span class="hlt">temperature</span> TCLD at the level defined by ZC. The ``effective cirrus emissivity'' NE where N is the fractional cirrus cloudiness and E is the emissivity in a GOES infrared picture element of about 10 km × 10 km is then computed from TB, TS and TCLD.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120014342','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120014342"><span><span class="hlt">Observations</span> of C-band <span class="hlt">Brightness</span> <span class="hlt">Temperature</span> from the Hurricane Imaging Radiometer (HIRAD) During GRIP</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.; Roberts, J. B.; Buckley, C. D.; Jones, W. L.; Biswas, S.; May, C.; Ruf, C. S.; Uhlhorn, E. W.; Atlas, R.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20120014342'); toggleEditAbsImage('author_20120014342_show'); toggleEditAbsImage('author_20120014342_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20120014342_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20120014342_hide"></p> <p>2012-01-01</p> <p>HIRAD is a new technology developed by NASA/MSFC, in partnership with NOAA and the Universities of Central Florida, Michigan, and Alabama-Huntsville. HIRAD is designed to measure wind speed and rain rate over a wide swath in heavy-rain, strong-wind conditions. HIRAD is expected to eventually fly routinely on unmanned aerial vehicles (UAVs) such as Global Hawk over hurricanes threatening the U.S. coast and other Atlantic basin areas, and possibly in the Western Pacific as well. HIRAD first flew on GRIP in 2010 and is planned to fly 2012-14 on the NASA Hurricane and Severe Storm Sentinel (HS3) missions on the Global Hawk, a high-altitude UAV. HIRAD technology will eventually be used on a satellite platform to extend the dynamical range of Ocean Surface Wind (OSV) <span class="hlt">observations</span> from space.</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/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 surface 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=20030020917&hterms=methods+quantitative&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dmethods%2Bquantitative','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030020917&hterms=methods+quantitative&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dmethods%2Bquantitative"><span>A Method to Retrieve Rainfall Rate over Land from TRMM <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>Prabhakara, C.; Iacovazzi, R., Jr.; Yoo, J.-M.</p> <p>2002-01-01</p> <p>Tropical Rainfall Measuring Mission (TRMM) Precipitation Radar (PR) <span class="hlt">observations</span> over mesoscale convective systems (MCSs) reveal that there are localized maxima in the rain rate with a scale of about 10 to 20 km that represent thunderstorms (Cbs). Some of these Cbs are developing or intense, while others are decaying or weak. These Cbs constitute only about 20 % of the rain area of a given MCS. Outside of Cbs, the average rain rate is much weaker than that within Cbs. From an analysis of the PR data, we find that the spatial distribution of rain and its character, convective or stratiform, is highly inhomogeneous. This complex nature of rain exists on a scale comparable to that of a Cb. The 85 GHz <span class="hlt">brightness</span> <span class="hlt">temperature</span>, T85, <span class="hlt">observations</span> of the TRMM Microwave Imager (TMI) radiometer taken over an MCS reflect closely the PR rain rate pattern over land. Local maxima in rain rate shown by PR are <span class="hlt">observed</span> as local minima in T85. Where there are no minima in T85, PR <span class="hlt">observations</span> indicate there is light rain. However, the TMI <span class="hlt">brightness</span> <span class="hlt">temperature</span> measurements (Tbs) have poor ability to discriminate convective rain from stratiform rain. For this reason, a TMI rain retrieval procedure that depends primarily on the magnitude of Tbs performs poorly. In order to retrieve rain rate from TMI data on land one has to include the spatial distribution information deduced from the T85 data in the retrieval method. Then, quantitative estimation of rain rate can be accomplished. A TMI rain retrieval method developed along these lines can yield estimates of rain rate and its frequency distribution which agree closely with that given by PR. We find the current TRMM project TMI (Version 5) rain retrieval algorithm on land could be improved with the retrieval scheme developed here. To support the conceptual frame work of the rain retrieval method developed here, a theoretical analysis of the TMI <span class="hlt">brightness</span> <span class="hlt">temperatures</span> in convective and stratiform regions is presented.</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('https://ntrs.nasa.gov/search.jsp?R=19920058934&hterms=fashion+models&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dfashion%2Bmodels','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920058934&hterms=fashion+models&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dfashion%2Bmodels"><span>Foundations for statistical-physical precipitation retrieval from passive microwave satellite measurements. I - <span class="hlt">Brightness-temperature</span> properties of a time-dependent cloud-radiation model</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smith, Eric A.; Mugnai, Alberto; Cooper, Harry J.; Tripoli, Gregory J.; Xiang, Xuwu</p> <p>1992-01-01</p> <p>The relationship between emerging microwave <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (T(B)s) and vertically distributed mixtures of liquid and frozen hydrometeors was investigated, using a cloud-radiation model, in order to establish the framework for a hybrid statistical-physical rainfall retrieval algorithm. Although strong relationships were found between the T(B) values and various rain parameters, these correlations are misleading in that the T(B)s are largely controlled by fluctuations in the ice-particle mixing ratios, which in turn are highly correlated to fluctuations in liquid-particle mixing ratios. However, the empirically based T(B)-rain-rate (T(B)-RR) algorithms can still be used as tools for estimating precipitation if the hydrometeor profiles used for T(B)-RR algorithms are not specified in an ad hoc fashion.</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> <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('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 surface 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/2017Icar..294...72H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Icar..294...72H"><span>Comparison and evaluation of the Chang'E microwave radiometer data based on theoretical computation of <span class="hlt">brightness</span> <span class="hlt">temperatures</span> at the Apollo 15 and 17 sites</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; Chan, Kwing L.; Zheng, Yong-Chun; Tsang, Kang T.; Xu, Ao-Ao</p> <p>2017-09-01</p> <p>There are significant differences (in the order of 3 to 20 K) between the lunar <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (TBs) as measured by the microwave radiometers (MRM) onboard Chang'E (CE)-1 and -2. To determine which set is more accurate, we have carried out a dataset comparison using theoretical calculations of the TBs (four frequency channels) versus local time at the Apollo 15 and 17 landing sites, where the thermal parameters are well-constrained by the in-situ measurements. Based on these parameters, we sought to constrain fits between theory and <span class="hlt">observation</span>, as uncertainties still exist in parameters involved in the microwave transfer computation. We found that: (i) CE-1/2 TBs have almost constant biases (negative, different for different channels) from the theoretical TBs. The averaged biases for each channel are smaller for CE-1; (ii) TBs of the high frequency channels (19.35/37 GHz) show a better fit with theory than the low frequency channels. The channel 4 (37 GHz) TBs from CE-1 are consistently shifted by about 1 K from the theoretical values. Adjustments in the order of 20 K are instead needed for the two CE-2 low frequency channels (3/7.8 GHz). Based on this comparison, we conclude that the CE-1 dataset to be more accurate than CE-2 one in terms of <span class="hlt">temperature</span> accuracy (not spatial resolution). We also offer a possible explanation for the significant TB differences between CE-1 and CE-2, and propose a possible recalibration method as a starting point towards the realignment of the two datasets.</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> </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('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/2015AdSpR..55.1234W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AdSpR..55.1234W"><span>Effect of terrestrial radiation on <span class="hlt">brightness</span> <span class="hlt">temperature</span> at lunar nearside: Based on theoretical calculation and data analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wei, Guangfei; Li, Xiongyao; Wang, Shijie</p> <p>2015-02-01</p> <p>Terrestrial radiation is another possible source of heat in lunar thermal environment at its nearside besides the solar illumination. On the basis of Clouds and the Earth's Radiant Energy System (CERES) data products, the effect of terrestrial radiation on the <span class="hlt">brightness</span> <span class="hlt">temperature</span> (TBe) of the lunar nearside has been theoretically calculated. It shows that the mafic lunar mare with high TBe is more sensitive to terrestrial radiation than the feldspathic highland with low TBe value. According to the synchronous rotation of the Moon, we extract TBe on lunar nearside using the microwave radiometer data from the first Chinese lunar probe Chang'E-1 (CE-1). Consistently, the average TBe at Mare Serenitatis is about 1.2 K while the highland around the Geber crater (19.4°S, 13.9°E) is relatively small at ∼0.4 K. Our results indicate that there is no significant effect of terrestrial radiation on TBe at the lunar nearside. However, to extract TBe accurately, effects of heat flow, rock abundance and subsurface rock fragments which are more significant should be considered in the future work.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995AJ....110..573M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995AJ....110..573M"><span>Galaxy Selection and the Surface <span class="hlt">Brightness</span> Distribution</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.; Bothun, Gregory D.; Schombert, James M.</p> <p>1995-08-01</p> <p>Optical surveys for galaxies are biased against the inclusion of low surface <span class="hlt">brightness</span> (LSB) galaxies. Disney [Nature, 263,573(1976)] suggested that the constancy of disk central surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span>. Using the physically reasonable assumption (which is motivated by the data) that central surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span>) 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 <span class="hlt">brightness</span> galaxies (those with central surface <span class="hlt">brightness</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.H21G1501C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.H21G1501C"><span><span class="hlt">Observed</span> and Projected Droughts Conditioned on <span class="hlt">Temperature</span> Change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chiang, F.; AghaKouchak, A.; Mazdiyasni, O.</p> <p>2016-12-01</p> <p>Droughts have had severe urban, agricultural and wildlife impacts in historical and recent years. In addition, during times of water scarcity, heat stress has been shown to produce compounding climatic and environmental effects. Understanding the overall conditions associated with drought intensities is important for mapping the anatomy of the climate in the changing world. For the study, we evaluated the relationship drought severity has exhibited with <span class="hlt">temperature</span> shifts between <span class="hlt">observed</span> periods and also between an ensemble of BCSD downscaled CMIP5 projected and historically modeled datasets. We compared <span class="hlt">temperatures</span> during different categories of drought severity on a monthly scale, and mapped areas displaying an escalation of <span class="hlt">temperature</span> with stricter definitions of drought. A historical shift of warmer <span class="hlt">temperatures</span> in more severe droughts was <span class="hlt">observed</span> most consistently in Southwestern and Eastern states between the later half of the 20th century and a reference period of the early half of the 20th century. Future projections from an ensemble of CMIP5 models also showed a shift to warmer <span class="hlt">temperatures</span> during more intense drought events in similar states. Preliminary statistics show that in many areas future droughts will be warmer that the average projected climate. These <span class="hlt">observed</span> and forecasted shifts in the heating intensity of severe drought events underscore the need to further research these patterns and relationships both spatially and temporally.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9149E..0MK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9149E..0MK"><span>Quantifying photometric <span class="hlt">observing</span> conditions on Paranal using an IR camera</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kerber, Florian; Querel, Richard R.; Hanuschik, Reinhard</p> <p>2014-08-01</p> <p>A Low Humidity and <span class="hlt">Temperature</span> Profiling (LHATPRO) microwave radiometer, manufactured by Radiometer Physics GmbH (RPG), is used to monitor sky conditions over ESO's Paranal observatory in support of VLT science operations. In addition to measuring precipitable water vapour (PWV) the instrument also contains an IR camera measuring sky <span class="hlt">brightness</span> <span class="hlt">temperature</span> at 10.5 μm. Due to its extended operating range down to -100 °C it is capable of detecting very cold and very thin, even sub-visual, cirrus clouds. We present a set of instrument flux calibration values as compared with a detrended fluctuation analysis (DFA) of the IR camera zenith-looking sky <span class="hlt">brightness</span> data measured above Paranal taken over the past two years. We show that it is possible to quantify photometric <span class="hlt">observing</span> conditions and that the method is highly sensitive to the presence of even very thin clouds but robust against variations of sky <span class="hlt">brightness</span> caused by effects other than clouds such as variations of precipitable water vapour. Hence it can be used to determine photometric conditions for science operations. About 60 % of nights are free of clouds on Paranal. More work will be required to classify the clouds using this technique. For the future this approach might become part of VLT science operations for evaluating nightly sky conditions.</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 surfaces for four test field heights can be described as four partially overlapping sections of a single larger surface. 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/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('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('http://adsabs.harvard.edu/abs/2015AGUFMGC43A1174R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMGC43A1174R"><span><span class="hlt">Observed</span> Decrease of North American Winter <span class="hlt">Temperature</span> Variability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rhines, A. N.; Tingley, M.; McKinnon, K. A.; Huybers, P. J.</p> <p>2015-12-01</p> <p>There is considerable interest in determining whether <span class="hlt">temperature</span> variability has changed in recent decades. Model ensembles project that extratropical land <span class="hlt">temperature</span> variance will detectably decrease by 2070. We use quantile regression of station <span class="hlt">observations</span> to show that decreasing variability is already robustly detectable for North American winter during 1979--2014. Pointwise trends from GHCND stations are mapped into a continuous spatial field using thin-plate spline regression, resolving small-scales while providing uncertainties accounting for spatial covariance and varying station density. We find that variability of daily <span class="hlt">temperatures</span>, as measured by the difference between the 95th and 5th percentiles, has decreased markedly in winter for both daily minima and maxima. Composites indicate that the reduced spread of winter <span class="hlt">temperatures</span> primarily results from Arctic amplification decreasing the meridional <span class="hlt">temperature</span> gradient. Greater <span class="hlt">observed</span> warming in the 5th relative to the 95th percentile stems from asymmetric effects of advection during cold versus warm days; cold air advection is generally from northerly regions that have experienced greater warming than western or southwestern regions that are generally sourced during warm days.</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('https://ntrs.nasa.gov/search.jsp?R=20170007359&hterms=temperature+classes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtemperature%2Bclasses','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170007359&hterms=temperature+classes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtemperature%2Bclasses"><span>Scale Dependence of Cirrus Horizontal Heterogeneity Effects on TOA Measurements. Part I; MODIS <span class="hlt">Brightness</span> <span class="hlt">Temperatures</span> in the Thermal Infrared</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fauchez, Thomas; Platnick, Steven; Meyer, Kerry; Cornet, Celine; Szczap, Frederic; Varnai, Tamas</p> <p>2017-01-01</p> <p>This paper presents a study on the impact of cirrus cloud heterogeneities on MODIS simulated thermal infrared (TIR) <span class="hlt">brightness</span> <span class="hlt">temperatures</span> (BTs) at the top of the atmosphere (TOA) as a function of spatial resolution from 50 meters to 10 kilometers. A realistic 3-D (three-dimensional) cirrus field is generated by the 3DCLOUD model (average optical thickness of 1.4, cloudtop and base altitudes at 10 and 12 kilometers, respectively, consisting of aggregate column crystals of D (sub eff) equals 20 microns), and 3-D thermal infrared radiative transfer (RT) is simulated with the 3DMCPOL (3-D Monte Carlo Polarized) code. According to previous studies, differences between 3-D BT computed from a heterogenous pixel and 1-D (one-dimensional) RT computed from a homogeneous pixel are considered dependent at nadir on two effects: (i) the optical thickness horizontal heterogeneity leading to the plane-parallel homogeneous bias (PPHB); and the (ii) horizontal radiative transport (HRT) leading to the independent pixel approximation error (IPAE). A single but realistic cirrus case is simulated and, as expected, the PPHB mainly impacts the low-spatial resolution results (above approximately 250 meters), with averaged values of up to 5-7 K (thousand), while the IPAE mainly impacts the high-spatial resolution results (below approximately 250 meters) with average values of up to 1-2 K (thousand). A sensitivity study has been performed in order to extend these results to various cirrus optical thicknesses and heterogeneities by sampling the cirrus in several ranges of parameters. For four optical thickness classes and four optical heterogeneity classes, we have found that, for nadir <span class="hlt">observations</span>, the spatial resolution at which the combination of PPHB and HRT effects is the smallest, falls between 100 and 250 meters. These spatial resolutions thus appear to be the best choice to retrieve cirrus optical properties with the smallest cloud heterogeneity-related total bias in the thermal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870046562&hterms=Pleiades&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DPleiades','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870046562&hterms=Pleiades&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DPleiades"><span>IRAS surface <span class="hlt">brightness</span> maps of reflection nebulae in the Pleiades</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Castelaz, Michael W.; Werner, M. W.; Sellgren, K.</p> <p>1987-01-01</p> <p>Surface <span class="hlt">brightness</span> 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 <span class="hlt">brightness</span> of the 12 and 25 micron emission and the absence of <span class="hlt">temperature</span> 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.</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 surface strongly influences surface'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/2018ApJ...853..113W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...853..113W"><span>Empirical <span class="hlt">Temperature</span> Measurement in Protoplanetary Disks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weaver, Erik; Isella, Andrea; Boehler, Yann</p> <p>2018-02-01</p> <p>The accurate measurement of <span class="hlt">temperature</span> in protoplanetary disks is critical to understanding many key features of disk evolution and planet formation, from disk chemistry and dynamics, to planetesimal formation. This paper explores the techniques available to determine <span class="hlt">temperatures</span> from <span class="hlt">observations</span> of single, optically thick molecular emission lines. Specific attention is given to issues such as the inclusion of optically thin emission, problems resulting from continuum subtraction, and complications of real <span class="hlt">observations</span>. Effort is also made to detail the exact nature and morphology of the region emitting a given line. To properly study and quantify these effects, this paper considers a range of disk models, from simple pedagogical models to very detailed models including full radiative transfer. Finally, we show how the use of the wrong methods can lead to potentially severe misinterpretations of data, leading to incorrect measurements of disk <span class="hlt">temperature</span> profiles. We show that the best way to estimate the <span class="hlt">temperature</span> of emitting gas is to analyze the line peak emission map without subtracting continuum emission. Continuum subtraction, which is commonly applied to <span class="hlt">observations</span> of line emission, systematically leads to underestimation of the gas <span class="hlt">temperature</span>. We further show that once <span class="hlt">observational</span> effects such as beam dilution and noise are accounted for, the line <span class="hlt">brightness</span> <span class="hlt">temperature</span> derived from the peak emission is reliably within 10%–15% of the physical <span class="hlt">temperature</span> of the emitting region, assuming optically thick emission. The methodology described in this paper will be applied in future works to constrain the <span class="hlt">temperature</span>, and related physical quantities, in protoplanetary disks <span class="hlt">observed</span> with ALMA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9768E..05V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9768E..05V"><span>Single-crystal phosphors for high-<span class="hlt">brightness</span> white LEDs/LDs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Víllora, Encarnación G.; Arjoca, Stelian; Inomata, Daisuke; Shimamura, Kiyoshi</p> <p>2016-03-01</p> <p>White light-emitting diodes (wLEDs) are the new environmental friendly sources for general lighting purposes. For applications requiring a high-<span class="hlt">brightness</span>, current wLEDs present overheating problems, which drastically decrease their emission efficiency, color quality and lifetime. This work gives an overview of the recent investigations on single-crystal phosphors (SCPs), which are proposed as novel alternative to conventional ceramic powder phosphors (CPPs). This totally new approach takes advantage of the superior properties of single-crystals in comparison with ceramic materials. SCPs exhibit an outstanding conversion efficiency and thermal stability up to 300°C. Furthermore, compared with encapsulated CPPs, SCPs possess a superior thermal conductivity, so that generated heat can be released efficiently. The conjunction of all these characteristics results in a low <span class="hlt">temperature</span> rise of SCPs even under high blue irradiances, where conventional CPPs are overheated or even burned. Therefore, SCPs represent the ideal, long-demanded all-inorganic phosphors for high-<span class="hlt">brightness</span> white light sources, especially those involving the use of high-density laser-diode beams.</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://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 surface soil moisture that are derived from SMOS <span class="hlt">brightness</span> <span class="hlt">temperature</span> <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> <span class="hlt">temperatures</span> for surface 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> </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/2000A%26A...358..749R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000A%26A...358..749R"><span>Radio <span class="hlt">observations</span> of a coronal mass ejection induced depletion in the outer solar corona</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramesh, R.; Sastry, Ch. V.</p> <p>2000-06-01</p> <p>We report the first low frequency radio <span class="hlt">observations</span> of a depletion that occurred in the outer solar corona in the aftermath of the CME event of 1986 June 5, with the large E-W one dimensional grating interferometer at the Gauribidanur radio observatory. We estimated the mass loss associated with the depletion and found that it agrees well with the value obtained through white light <span class="hlt">observations</span> of the event. The radio <span class="hlt">brightness</span> <span class="hlt">temperature</span> at the location of the depletion was less by a factor of ~ 7 compared to the ambient. The angular extent over which the decrease in <span class="hlt">brightness</span> took place was <= 3'. The electron density variation was found to be proportional to r-10. Since <span class="hlt">observations</span> at different wavelength bands have different physical origins, the radio method might be useful in independently estimating the characteristics of CME induced coronal depletions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMGC21E0980P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMGC21E0980P"><span>An '<span class="hlt">Observational</span> Large Ensemble' to compare <span class="hlt">observed</span> and modeled <span class="hlt">temperature</span> trend uncertainty due to internal variability.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Poppick, A. N.; McKinnon, K. A.; Dunn-Sigouin, E.; Deser, C.</p> <p>2017-12-01</p> <p>Initial condition climate model ensembles suggest that regional <span class="hlt">temperature</span> trends can be highly variable on decadal timescales due to characteristics of internal climate variability. Accounting for trend uncertainty due to internal variability is therefore necessary to contextualize recent <span class="hlt">observed</span> <span class="hlt">temperature</span> changes. However, while the variability of trends in a climate model ensemble can be evaluated directly (as the spread across ensemble members), internal variability simulated by a climate model may be inconsistent with <span class="hlt">observations</span>. <span class="hlt">Observation</span>-based methods for assessing the role of internal variability on trend uncertainty are therefore required. Here, we use a statistical resampling approach to assess trend uncertainty due to internal variability in historical 50-year (1966-2015) winter near-surface air <span class="hlt">temperature</span> trends over North America. We compare this estimate of trend uncertainty to simulated trend variability in the NCAR CESM1 Large Ensemble (LENS), finding that uncertainty in wintertime <span class="hlt">temperature</span> trends over North America due to internal variability is largely overestimated by CESM1, on average by a factor of 32%. Our <span class="hlt">observation</span>-based resampling approach is combined with the forced signal from LENS to produce an '<span class="hlt">Observational</span> Large Ensemble' (OLENS). The members of OLENS indicate a range of spatially coherent fields of <span class="hlt">temperature</span> trends resulting from different sequences of internal variability consistent with <span class="hlt">observations</span>. The smaller trend variability in OLENS suggests that uncertainty in the historical climate change signal in <span class="hlt">observations</span> due to internal variability is less than suggested by LENS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22703420','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22703420"><span>Intrauterine <span class="hlt">temperature</span> during intrapartum amnioinfusion: a prospective <span class="hlt">observational</span> study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tomlinson, T M; Schaecher, C; Sadovsky, Y; Gross, G</p> <p>2012-07-01</p> <p>To determine the influence of routine intrapartum amnioinfusion (AI) on intrauterine <span class="hlt">temperature</span>. Prospective <span class="hlt">observational</span> study. Maternity unit, Barnes Jewish Hospital, St Louis, MO, USA. Forty women with singleton gestations and an indication for intrapartum intrauterine pressure catheter placement. Using a <span class="hlt">temperature</span> probe, we digitally recorded intrauterine <span class="hlt">temperature</span> every 10 minutes during labour. Amnioinfusion was administered according to a standard protocol using saline equilibrated to the ambient <span class="hlt">temperature</span>. Mean intrauterine <span class="hlt">temperature</span> during labour. Participants were monitored for a mean of 280 minutes (range 20-820). A total of 164 intrauterine <span class="hlt">temperature</span> readings in the AI cohort were compared with 797 control measurements. When compared with controls, we <span class="hlt">observed</span> a lower intrauterine <span class="hlt">temperature</span> in the AI cohort (36.4 versus 37.4°C, P<0.01). More measurements in the AI cohort were recorded in the presence of intrapartum fever (40% versus 30%). A subgroup analysis of measurements recorded in afebrile parturients revealed an even greater effect of AI (1.5°C decrease, 37.3 versus 35.8°C, P<0.01). Routine intrapartum AI using saline equilibrated to a mean ambient <span class="hlt">temperature</span> of 25.0°C reduces intrauterine <span class="hlt">temperature</span> and may thereby affect fetal core <span class="hlt">temperature</span>. © 2012 The Authors BJOG An International Journal of Obstetrics and Gynaecology © 2012 RCOG.</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://www.ncbi.nlm.nih.gov/pubmed/17367827','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17367827"><span>Subjective time runs faster under the influence of <span class="hlt">bright</span> rather than dim light conditions during the forenoon.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Morita, Takeshi; Fukui, Tomoe; Morofushi, Masayo; Tokura, Hiromi</p> <p>2007-05-16</p> <p>The study investigated if 6 h morning <span class="hlt">bright</span> light exposure, compared with dim light exposure, could influence time sense (range: 5-15 s). Eight women served as participants. The participant entered a bioclimatic chamber at 10:00 h on the day before the test day, where an ambient <span class="hlt">temperature</span> and relative humidity were controlled at 25 degrees C and 60%RH. She sat quietly in a sofa in 50 lx until 22:00 h, retired at 22:00 h and then slept in total darkness. She rose at 07:00 h the following morning and again sat quietly in a sofa till 13:00 h, either in <span class="hlt">bright</span> (2500 lx) or dim light (50 lx), the order of light intensities between the two occasions being randomized. The time-estimation test was performed from 13:00 to 13:10 h in 200 lx. The participant estimated the time that had elapsed between two buzzers, ranging over 5-15 s, and inputting the estimate into a computer. The test was carried out separately upon each individual. Results showed that the participants estimated higher durations of the given time intervals after previous exposure to 6 h of <span class="hlt">bright</span> rather than dim light. The finding is discussed in terms of different load errors (difference between the actual core <span class="hlt">temperature</span> and its thermoregulatory set-point) following 6-h exposure to <span class="hlt">bright</span> or dim light in the morning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.9489P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.9489P"><span>Metre-size <span class="hlt">bright</span> spots at the surface of comet 67P/Churyumov-Gerasimenko: Interpretation of OSIRIS data using laboratory experiments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>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</p> <p>2015-04-01</p> <p>Since the beginning of Rosetta's orbital <span class="hlt">observations</span>, over a hundred small <span class="hlt">bright</span> spots have been identified in images returned by its OSIRIS NAC camera, in all types of morphological regions on the nucleus. <span class="hlt">Bright</span> 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 <span class="hlt">observed</span> 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 <span class="hlt">observed</span> at the surfaces of boulders of larger dimension. The <span class="hlt">brightness</span> of these spots is up to ten times the average <span class="hlt">brightness</span> 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 <span class="hlt">bright</span> spots. In particular, recent sublimation experiments conducted at the University of Bern reproduce the spectro-photometric variability <span class="hlt">observed</span> 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 <span class="hlt">observations</span> and laboratory results, our favoured interpretation of the <span class="hlt">observed</span> features is that the <span class="hlt">bright</span> 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</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900031734&hterms=Symbiotic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DSymbiotic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900031734&hterms=Symbiotic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DSymbiotic"><span>Multifrequency <span class="hlt">observations</span> of symbiotic stars</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kenyon, Scott J.</p> <p>1988-01-01</p> <p>The discovery of symbiotic stars is described, and the results of multifrequency <span class="hlt">observations</span> made during the past two decades are presented. <span class="hlt">Observational</span> data identify symbiotic stars as long-period binary systems that can be divided into two basic physical classes: detached symbiotics containing a red giant (or a Mira variable), and semidetached symbiotics containing a lobe-filling red giant and a solar-type main sequence star. Three components are typically <span class="hlt">observed</span>: (1) the cool giant component with an effective <span class="hlt">temperature</span> of 2500-4000 K, which can be divided by the IR spectral classification into normal M giants (S-types) and heavily reddened Mira variables (D-types); (2) the hot companion displaying a <span class="hlt">bright</span> blue continuum at UV wavelengths, which is sometimes also an X-ray source; and (3) a gaseous nebula enveloping the binary.</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://ntrs.nasa.gov/search.jsp?R=20080031657&hterms=Qbo&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DQbo','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080031657&hterms=Qbo&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DQbo"><span><span class="hlt">Temperatures</span> and Composition in the Saturn System from Cassini CIRS</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flasar, F. Michael</p> <p>2008-01-01</p> <p>We summarize recent <span class="hlt">observations</span> by the Composite Infrared Spectrometer of Saturn, its rings, Titan, and the icy satellites. Limb <span class="hlt">observations</span> of Saturn show vertical oscillations of <span class="hlt">temperatures</span> and zonal-wind shears in the equatorial region that may be related to a temporal oscillation similar to the terrestrial QBO and Jupiter's QQO. There is also evidence of subsidence at mid-northern latitudes driven by the equatorial activity. Nadir-viewing <span class="hlt">observations</span> show compact warm spots in the troposphere and stratosphere at both (summer and winter) poles, likely associated with subsidence. <span class="hlt">Observations</span> of Titan have defined better the characteristics of the northern winter polar vortex, with 190 m/s winds surrounding a cold atmosphere at 1 microbar. The very warm polar stratopause at 10 microbar and the enhanced abundances of organic compounds suggest subsidence within the vortex. Analysis of the zonal structure in <span class="hlt">temperature</span> indicates that the stratospheric zonal winds rotate about an axis that is displaced approximately 4.1 deg from the IAU pole. Additional flybys, including a close one in March 2008, continue to characterize the endogenic activity in Enceladus s south polar region. <span class="hlt">Temperature</span> maps of <span class="hlt">bright</span> and dark terrains on Iapetus indicate that its ice is approximately stable to sublimation in the <span class="hlt">bright</span> regions and highly unstable in the dark regions. Thermal mapping of Saturn s rings continues to constrain their composition, and <span class="hlt">observations</span> at different solar phase angles, spacecraft elevations, solar elevations, and local hour angles have elucidated the effects of ring-particle shadowing and vertical motions on the thermal structure, and revealed the presence of small-scale structure associated with self-gravity wakes.</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/1999AAS...195.5312C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999AAS...195.5312C"><span>Millimeter Wavelength <span class="hlt">Observations</span> of Galactic Sources with the Mobile Anisotropy Telescope (MAT)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cruz, K. L.; Caldwell, R.; Devlin, M. J.; Dorwart, W. B.; Herbig, T.; Miller, A. D.; Nolta, M. R.; Page, L. A.; Puchalla, J. L.; Torbet, E.; Tran, H. T.</p> <p>1999-12-01</p> <p>The Mobile Anisotropy Telescope (MAT) has completed two <span class="hlt">observing</span> seasons (1997 and 1998) in Chile from the Cerro Toco site. Although the primary goal of MAT was to measure anisotropy in the Cosmic Microwave Background (CMB) radiation, the chosen <span class="hlt">observation</span> scheme also allowed daily viewing of the Galactic Plane. We present filtered maps at 30, 40 and 144 GHz of a region of the Galactic Plane which contains several millimeter-<span class="hlt">bright</span> regions including the Carinae nebula and IRAS 11097-6102. We report the best fit <span class="hlt">brightness</span> <span class="hlt">temperatures</span> as well as the total flux densities in the MAT beams (0.9, 0.6 and 0.2 degrees FWHM) . The data are calibrated with respect to Jupiter whose flux is known to better than 8% in all frequency bands. This work was funded by the National Science Foundation and the Packard Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140006483','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140006483"><span>Anti­-parallel Filament Flows and <span class="hlt">Bright</span> Dots <span class="hlt">Observed</span> in the EUV with Hi-­C</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Alexander, Caroline E.; Regnier, Stephane; Walsh, Robert; Winebarger, Amy</p> <p>2013-01-01</p> <p>Hi-C obtained the highest spatial and temporal resolution <span class="hlt">observations</span> ever taken in the solar EUV corona. Hi-C reveals dynamics and structure at the limit of its temporal and spatial resolution. Hi-C <span class="hlt">observed</span> various fine-scale features that SDO/AIA could not pick out. For the first time in the corona, Hi-C revealed magnetic braiding and component reconnection consistent with coronal heating. Hi-C shows evidence of reconnection and heating in several different regions and magnetic configurations with plasma being heated to 0.3 - 8 x 10(exp 6) K <span class="hlt">temperatures</span>. Surprisingly, many of the first results highlight plasma at <span class="hlt">temperatures</span> that are not at the peak of the response functions.</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 surface <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 <span class="hlt">temperatures</span>, 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://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/2018AAS...23122008G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23122008G"><span>Physical Conditions in the Solar Corona Derived from the Total Solar Eclipse <span class="hlt">Observations</span> obtained on 2017 August 21 Using a Polarization Camera</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gopalswamy, N.; Yashiro, Seiji; Reginald, Nelson; Thakur, Neeharika; Thompson, Barbara J.; Gong, Qian</p> <p>2018-01-01</p> <p>We present preliminary results obtained by <span class="hlt">observing</span> the solar corona during the 2017 August 21 total solar eclipse using a polarization camera mounted on an eight-inch Schmidt-Cassegrain telescope. The <span class="hlt">observations</span> were made from Madras Oregon during 17:19 to 17:21 UT. Total and polarized <span class="hlt">brightness</span> images were obtained at four wavelengths (385, 398.5, 410, and 423 nm). The polarization camera had a polarization mask mounted on a 2048x2048 pixel CCD with a pixel size of 7.4 microns. The resulting images had a size of 975x975 pixels because four neighboring pixels were summed to yield the polarization and total <span class="hlt">brightness</span> images. The ratio of 410 and 385 nm images is a measure of the coronal <span class="hlt">temperature</span>, while that at 423 and 398.5 nm images is a measure of the coronal flow speed. We compared the <span class="hlt">temperature</span> map from the eclipse <span class="hlt">observations</span> with that obtained from the Solar Dynamics Observatory’s Atmospheric Imaging Assembly images at six EUV wavelengths, yielding consistent <span class="hlt">temperature</span> information of the corona.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19760038549&hterms=right+Bless+you&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dright%2BBless%2Byou','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760038549&hterms=right+Bless+you&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dright%2BBless%2Byou"><span>Empirical effective <span class="hlt">temperatures</span> and bolometric corrections for early-type stars</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Code, A. D.; Bless, R. C.; Davis, J.; Brown, R. H.</p> <p>1976-01-01</p> <p>An empirical effective <span class="hlt">temperature</span> for a star can be found by measuring its apparent angular diameter and absolute flux distribution. The angular diameters of 32 <span class="hlt">bright</span> stars in the spectral range O5f to F8 have recently been measured with the stellar interferometer at Narrabri Observatory, and their absolute flux distributions have been found by combining <span class="hlt">observations</span> of ultraviolet flux from the Orbiting Astronomical Observatory (OAO-2) with ground-based photometry. In this paper, these data have been combined to derive empirical effective <span class="hlt">temperatures</span> and bolometric corrections for these 32 stars.</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> </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/2060/19920016088','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19920016088"><span>An atlas of upper tropospheric radiances <span class="hlt">observed</span> in the 6 to 7-micrometer water vapor band using TOVS data from the NOAA weather satellites during 1979-1991</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chesters, Dennis; Sharma, OM</p> <p>1992-01-01</p> <p>This document is a pictorial atlas of the Earth's radiance emitted in the 6 to 7 micro-m water vapor band. At these wavelengths, the infrared <span class="hlt">brightness</span> <span class="hlt">temperature</span> corresponds to the layer-average <span class="hlt">temperature</span> of the top few millimeters of water vapor in the atmosphere. At low altitudes, <span class="hlt">bright</span> regions are dry slots in the upper troposphere. The satellite <span class="hlt">observations</span> were obtained from NOAA's cloud and angle corrected measurements made by a series of polar orbiting TOVS (TIROS Operational Vertical Sounder) instruments flown from 1979 to 1991. TOVS 6.7 micro-m and 7.2 micro-m channels were converted to a single <span class="hlt">brightness</span> <span class="hlt">temperature</span> that simulates a high altitude channel near '6.5' micro-m. For climatological studies, the daily '6.5' micro-m overpass data were gridded to a cartesian projection with 5 by 5 degree horizontal resolution between 40 degrees N and 40 degrees S latitude. This atlas presents greyscale images of the '6.5' micro-m <span class="hlt">brightness</span> fields for every day in every month for 13 years. The mean <span class="hlt">brightness</span> for each of the 12 months for 13 years is presented to display interannual variability, and the annual cycle of 12 monthly means is summarized on a single page. Statistical summaries are presented from other investigations in progress.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/14646259','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14646259"><span>Effect of <span class="hlt">bright</span> light on EEG activities and subjective sleepiness to mental task during nocturnal sleep deprivation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yokoi, Mari; Aoki, Ken; Shimomura, Yoshihiro; Iwanaga, Koichi; Katsuura, Tetsuo; Shiomura, Yoshihiro</p> <p>2003-11-01</p> <p>The purpose of this study was to investigate the effect of the exposure to <span class="hlt">bright</span> light on EEG activity and subjective sleepiness at rest and at the mental task during nocturnal sleep deprivation. Eight male subjects lay awake in semi-supine in a reclining seat from 21:00 to 04:30 under the <span class="hlt">bright</span> (BL; >2500 lux) or the dim (DL; <150 lux) light conditions. During the sleep deprivation, the mental task (Stroop color-word conflict test: CWT) was performed each 15 min in one hour. EEG, subjective sleepiness, rectal and mean skin <span class="hlt">temperatures</span> and urinary melatonin concentrations were measured. The subjective sleepiness increased with time of sleep deprivation during both rest and CWT under the DL condition. The exposure to <span class="hlt">bright</span> light delayed for 2 hours the increase in subjective sleepiness at rest and suppressed the increase in that during CWT. The <span class="hlt">bright</span> light exposure also delayed the increase in the theta and alpha wave activities in EEG at rest. In contrast, the effect of the <span class="hlt">bright</span> light exposure on the theta and alpha wave activities disappeared by CWT. Additionally, under the BL condition, the entire theta activity during CWT throughout nocturnal sleep deprivation increased significantly from that in a rest condition. Our results suggest that the exposure to <span class="hlt">bright</span> light throughout nocturnal sleep deprivation influences the subjective sleepiness during the mental task and the EEG activity, as well as the subjective sleepiness at rest. However, the effect of the <span class="hlt">bright</span> light exposure on the EEG activity at the mental task diminishes throughout nocturnal sleep deprivation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=316933','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=316933"><span>Applications of Land Surface <span class="hlt">Temperature</span> from Microwave <span class="hlt">Observations</span></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>Land surface <span class="hlt">temperature</span> (LST) is a key input for physically-based retrieval algorithms of hydrological states and fluxes. Yet, it remains a poorly constrained parameter for global scale studies. The main two <span class="hlt">observational</span> methods to remotely measure T are based on thermal infrared (TIR) <span class="hlt">observation</span>...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100033466','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100033466"><span>Improving a Spectral Bin Microphysical Scheme Using TRMM 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>Li, Xiaowen; Tao, Wei-Kuo; Matsui, Toshihisa; Liu, Chuntao; Masunaga, Hirohiko</p> <p>2010-01-01</p> <p>Comparisons between cloud model simulations and <span class="hlt">observations</span> are crucial in validating model performance and improving physical processes represented in the mod Tel.hese modeled physical processes are idealized representations and almost always have large rooms for improvements. In this study, we use data from two different sensors onboard TRMM (Tropical Rainfall Measurement Mission) satellite to improve the microphysical scheme in the Goddard Cumulus Ensemble (GCE) model. TRMM <span class="hlt">observed</span> mature-stage squall lines during late spring, early summer in central US over a 9-year period are compiled and compared with a case simulation by GCE model. A unique aspect of the GCE model is that it has a state-of-the-art spectral bin microphysical scheme, which uses 33 different bins to represent particle size distribution of each of the seven hydrometeor species. A forward radiative transfer model calculates TRMM Precipitation Radar (PR) reflectivity and TRMM Microwave Imager (TMI) 85 GHz <span class="hlt">brightness</span> <span class="hlt">temperatures</span> from simulated particle size distributions. Comparisons between model outputs and <span class="hlt">observations</span> reveal that the model overestimates sizes of snow/aggregates in the stratiform region of the squall line. After adjusting <span class="hlt">temperature</span>-dependent collection coefficients among ice-phase particles, PR comparisons become good while TMI comparisons worsen. Further investigations show that the partitioning between graupel (a high-density form of aggregate), and snow (a low-density form of aggregate) needs to be adjusted in order to have good comparisons in both PR reflectivity and TMI <span class="hlt">brightness</span> <span class="hlt">temperature</span>. This study shows that long-term satellite <span class="hlt">observations</span>, especially those with multiple sensors, can be very useful in constraining model microphysics. It is also the first study in validating and improving a sophisticated spectral bin microphysical scheme according to long-term satellite <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22364543-resolving-bright-hcn-emission-toward-seyfert-nucleus-m51-shock-enhancement-radio-jets-weak-masing-infrared-pumping','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22364543-resolving-bright-hcn-emission-toward-seyfert-nucleus-m51-shock-enhancement-radio-jets-weak-masing-infrared-pumping"><span>RESOLVING THE <span class="hlt">BRIGHT</span> HCN(1–0) EMISSION TOWARD THE SEYFERT 2 NUCLEUS OF M51: SHOCK ENHANCEMENT BY RADIO JETS AND WEAK MASING BY INFRARED PUMPING?</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>Matsushita, Satoki; Trung, Dinh-V-; Boone, Frédéric</p> <p>2015-01-20</p> <p>We present high angular resolution <span class="hlt">observations</span> of the HCN(1-0) emission (at ∼1'' or ∼34 pc), together with CO J = 1-0, 2-1, and 3-2 <span class="hlt">observations</span>, toward the Seyfert 2 nucleus of M51 (NGC 5194). The overall HCN(1-0) distribution and kinematics are very similar to that of the CO lines, which have been indicated as the jet-entrained molecular gas in our past <span class="hlt">observations</span>. In addition, high HCN(1-0)/CO(1-0) <span class="hlt">brightness</span> <span class="hlt">temperature</span> ratio of about unity is <span class="hlt">observed</span> along the jets, similar to that <span class="hlt">observed</span> at the shocked molecular gas in our Galaxy. These results strongly indicate that both diffuse and dense gases are entrained bymore » the jets and outflowing from the active galactic nucleus. The channel map of HCN(1-0) at the systemic velocity shows a strong emission right at the nucleus, where no obvious emission has been detected in the CO lines. The HCN(1-0)/CO(1-0) <span class="hlt">brightness</span> <span class="hlt">temperature</span> ratio at this region reaches >2, a value that cannot be explained considering standard physical/chemical conditions. Based on our calculations, we suggest infrared pumping and possibly weak HCN masing, but still requiring an enhanced HCN abundance for the cause of this high ratio. This suggests the presence of a compact dense obscuring molecular gas in front of the nucleus of M51, which remains unresolved at our ∼1'' (∼34 pc) resolution, and consistent with the Seyfert 2 classification picture.« less</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> <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('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 surface <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 surface <span class="hlt">brightness</span> depressions was displaced as the radio lobes expanded subsonically, leaving cavities in the hot atmosphere. The gas <span class="hlt">temperature</span> 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://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/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('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('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://www.osti.gov/servlets/purl/1336923','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1336923"><span><span class="hlt">Bright</span> and durable field emission source derived from refractory taylor cones</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>Hirsch, Gregory</p> <p></p> <p>A method of producing field emitters having improved <span class="hlt">brightness</span> and durability relying on the creation of a liquid Taylor cone from electrically conductive materials having high melting points. The method calls for melting the end of a wire substrate with a focused laser beam, while imposing a high positive potential on the material. The resulting molten Taylor cone is subsequently rapidly quenched by cessation of the laser power. Rapid quenching is facilitated in large part by radiative cooling, resulting in structures having characteristics closely matching that of the original liquid Taylor cone. Frozen Taylor cones thus obtained yield desirable tipmore » end forms for field emission sources in electron beam applications. Regeneration of the frozen Taylor cones in-situ is readily accomplished by repeating the initial formation procedures. The high <span class="hlt">temperature</span> liquid Taylor cones can also be employed as <span class="hlt">bright</span> ion sources with chemical elements previously considered impractical to implement.« less</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/2060/19760012432','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19760012432"><span>Use of Skylab EREP data in a sea-surface <span class="hlt">temperature</span> experiment. [Monroe Reservoir and Key West, Fla.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Anding, D. C. (Principal Investigator); Walker, J. P.</p> <p>1975-01-01</p> <p>The author has identified the following significant results. A sea surface <span class="hlt">temperature</span> experiment was studied, demonstrating the feasibility of a procedure for the remote measurement of sea surface <span class="hlt">temperature</span> which inherently corrects for the effect of the intervening atmosphere without recourse to climatological data. The procedure was applied to Skylab EREP S191 spectrometer data, and it is demonstrated that atmospheric effects on the <span class="hlt">observed</span> <span class="hlt">brightness</span> <span class="hlt">temperature</span> can be reduced to less than 1.0 K.</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> <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/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> </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/2060/19840016444','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('https://ntrs.nasa.gov/search.jsp?R=20050041715&hterms=Lamb&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DLamb','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050041715&hterms=Lamb&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DLamb"><span>Chandra <span class="hlt">Observations</span> of Extended X-Ray Emission in ARP 220</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>McDowell, J. C.; Clements, D. L.; Lamb, S. A.; Shaked, S.; Hearn, N. C.; Colina, L.; Mundell, C.; Borne, K.; Baker, A. C.; Arribas, S.</p> <p>2003-01-01</p> <p>We resolve the extended X-ray emission from the prototypical ultraluminous infrared galaxy Arp 220. Extended, faint, edge-brightened, soft X-ray lobes outside the optical galaxy are <span class="hlt">observed</span> to a distance of 1CL 15 kpc on each side of the nuclear region. <span class="hlt">Bright</span> plumes inside the optical isophotes coincide with the optical line emission and extend 1 1 kpc from end to end across the nucleus. The data for the plumes cannot be fitted by a single-<span class="hlt">temperature</span> plasma and display a range of <span class="hlt">temperatures</span> from 0.2 to 1 keV. The plumes emerge from <span class="hlt">bright</span>, diffuse circumnuclear emission in the inner 3 kpc centered on the Ha peak, which is displaced from the radio nuclei. There is a close morphological correspondence between the Ha and soft X-ray emission on all spatial scales. We interpret the plumes as a starburst-driven superwind and discuss two interpretations of the emission from the lobes in the context of simulations of the merger dynamics of Arp 220.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19730057181&hterms=induction+melting&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dinduction%2Bmelting','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19730057181&hterms=induction+melting&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dinduction%2Bmelting"><span>The emissivities of liquid metals at their fusion <span class="hlt">temperatures</span>.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bonnell, D. W.; Treverton, J. A.; Valerga, A. J.; Margrave , J. L.</p> <p>1972-01-01</p> <p>The emissivities for several transition metals and various other metals and compounds in the liquid state at their fusion <span class="hlt">temperatures</span> have been determined in this laboratory. The technique used involves electromagnetic levitation-induction heating of the materials in an inert atmosphere. The <span class="hlt">brightness</span> <span class="hlt">temperature</span> of the liquid phase of the material is measured as the material is heated through fusion. Given a reliable value of the fusion <span class="hlt">temperature</span>, which is available for most pure substances, one may readily calculate an emissivity for the liquid phase at the fusion <span class="hlt">temperature</span>. Even in cases where melting points are poorly known, the <span class="hlt">brightness</span> <span class="hlt">temperatures</span> are unique parameters, independent of the <span class="hlt">temperature</span> scale and measured for a chemically defined system at a fixed point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19860015159','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19860015159"><span>High-spectral resolution solar 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>Hurford, G. J.</p> <p>1986-01-01</p> <p>The application of high-spectral resolution microwave <span class="hlt">observations</span> to the study of solar activity is discussed with particular emphasis on the frequency dependence of microwave emission from solar active regions. A shell model of gyroresonance emission from active regions is described which suggest that high-spectral resolution, spatially-resolved <span class="hlt">observations</span> can provide quantitative information about the magnetic field distribution at the base of the corona. Corresponding <span class="hlt">observations</span> of a single sunspot with the Owens Valley frequency-agile interferometer at 56 frequencies between 1.2 and 14 Ghs are presented. The overall form of the <span class="hlt">observed</span> size and <span class="hlt">brightness</span> <span class="hlt">temperature</span> spectra was consistent with expectations based on the shell model, although there were differences of potential physical significance. The merits and weaknesses of microwave spectroscopy as a technique for measuring magnetic fields in the solar corona are briefly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.101i1910S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.101i1910S"><span><span class="hlt">Temperature</span> dependent recombination dynamics in InP/ZnS colloidal nanocrystals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shirazi, R.; Kopylov, O.; Kovacs, A.; Kardynał, B. E.</p> <p>2012-08-01</p> <p>In this letter, we investigate exciton recombination in InP/ZnS core-shell colloidal nanocrystals over a wide <span class="hlt">temperature</span> range. Over the entire range between room <span class="hlt">temperature</span> and liquid helium <span class="hlt">temperature</span>, multi-exponential exciton decay curves are <span class="hlt">observed</span> and well explained by the presence of <span class="hlt">bright</span> and dark exciton states, as well as defect states. Two different types of defect are present: one located at the core-shell interface and the other on the surface of the nanocrystal. Based on the <span class="hlt">temperature</span> dependent contributions of all four states to the total photoluminescence signal, we estimate that the four states are distributed within a 20 meV energy band in nanocrystals that emit at 1.82 eV.</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/2014ApPhL.105a4104D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ApPhL.105a4104D"><span><span class="hlt">Observations</span> of a mode transition in a hydrogen hollow cathode discharge using phase resolved optical emission spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dixon, Sam; Charles, Christine; Dedrick, James; Gans, Timo; O'Connell, Deborah; Boswell, Rod</p> <p>2014-07-01</p> <p>Two distinct operational modes are <span class="hlt">observed</span> in a radio frequency (rf) low pressure hydrogen hollow cathode discharge. The mode transition is characterised by a change in total light emission and differing expansion structures. An intensified CCD camera is used to make phase resolved images of Balmer α emission from the discharge. The low emission mode is consistent with a typical γ discharge, and appears to be driven by secondary electrons ejected from the cathode surface. The <span class="hlt">bright</span> mode displays characteristics common to an inductive discharge, including increased optical emission, power factor, and <span class="hlt">temperature</span> of the H2 gas. The <span class="hlt">bright</span> mode precipitates the formation of a stationary shock in the expansion, <span class="hlt">observed</span> as a dark region adjacent to the source-chamber interface.</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.osti.gov/biblio/22660906-quiescent-prominences-era-alma-simulated-observations-using-whole-prominence-fine-structure-model','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22660906-quiescent-prominences-era-alma-simulated-observations-using-whole-prominence-fine-structure-model"><span>QUIESCENT PROMINENCES IN THE ERA OF ALMA: SIMULATED <span class="hlt">OBSERVATIONS</span> USING THE 3D WHOLE-PROMINENCE FINE STRUCTURE MODEL</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>Gunár, Stanislav; Heinzel, Petr; Mackay, Duncan H.</p> <p>2016-12-20</p> <p>We use the detailed 3D whole-prominence fine structure model to produce the first simulated high-resolution ALMA <span class="hlt">observations</span> of a modeled quiescent solar prominence. The maps of synthetic <span class="hlt">brightness</span> <span class="hlt">temperature</span> and optical thickness shown in the present paper are produced using a visualization method for synthesis of the submillimeter/millimeter radio continua. We have obtained the simulated <span class="hlt">observations</span> of both the prominence at the limb and the filament on the disk at wavelengths covering a broad range that encompasses the full potential of ALMA. We demonstrate here extent to which the small-scale and large-scale prominence and filament structures will be visible inmore » the ALMA <span class="hlt">observations</span> spanning both the optically thin and thick regimes. We analyze the relationship between the <span class="hlt">brightness</span> and kinetic <span class="hlt">temperature</span> of the prominence plasma. We also illustrate the opportunities ALMA will provide for studying the thermal structure of the prominence plasma from the cores of the cool prominence fine structure to the prominence–corona transition region. In addition, we show that detailed 3D modeling of entire prominences with their numerous fine structures will be important for the correct interpretation of future ALMA <span class="hlt">observations</span> of prominences.« less</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('https://ntrs.nasa.gov/search.jsp?R=19760036568&hterms=hack&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dhack','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760036568&hterms=hack&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dhack"><span>High <span class="hlt">temperature</span> plasma in beta Lyrae, <span class="hlt">observed</span> from Copernicus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kondo, Y.; Hack, M.; Hutchings, J. B.; Mccluskey, G. E., Jr.; Plavec, M.; Polidan, R. S.</p> <p>1975-01-01</p> <p>High-resolution UV spectrophotometry of the complex close binary system beta Lyrae was performed with a telescope spectrometer on board Copernicus. <span class="hlt">Observations</span> were made at phases 0.0, 0.25, 0.5, and 0.75 with resolutions of 0.2 A (far-UV) and 0.4 A (mid-UV). The far-UV spectrum is completely dominated by emission lines indicating the existence of a high-<span class="hlt">temperature</span> plasma in this binary. The spectrum of this object is unlike that of any other object <span class="hlt">observed</span> from Copernicus. It is believed that this high-<span class="hlt">temperature</span> plasma results from dynamic mass transfer taking place in the binary. The current results are compared with OAO-2 <span class="hlt">observations</span> and other <span class="hlt">observational</span> results. The possibility that the secondary component is a collapsed object is also discussed; the Copernicus <span class="hlt">observations</span> are consistent with the hypothesis that the spectroscopically invisible secondary component is a black hole.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080039561','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080039561"><span>A Catalog of Galaxy Clusters <span class="hlt">Observed</span> by XMM-Newton</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Snowden, S. L.; Mushotzky, R. M.; Kuntz, K. D.; Davis, David S.</p> <p>2007-01-01</p> <p>Images and the radial profiles of the <span class="hlt">temperature</span>, abundance, and <span class="hlt">brightness</span> for 70 clusters of galaxies <span class="hlt">observed</span> by XMM-Newton are presented along with a detailed discussion of the data reduction and analysis methods, including background modeling, which were used in the processing. Proper consideration of the various background components is vital to extend the reliable determination of cluster parameters to the largest possible cluster radii. The various components of the background including the quiescent particle background, cosmic diffuse emission, soft proton contamination, and solar wind charge exchange emission are discussed along with suggested means of their identification, filtering, and/or their modeling and subtraction. Every component is spectrally variable, sometimes significantly so, and all components except the cosmic background are temporally variable as well. The distributions of the events over the FOV vary between the components, and some distributions vary with energy. The scientific results from <span class="hlt">observations</span> of low surface <span class="hlt">brightness</span> objects and the diffuse background itself can be strongly affected by these background components and therefore great care should be taken in their consideration.</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/2014AGUFMSA24A..01C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSA24A..01C"><span>Plasma-Neutral Coupling on the Dark and <span class="hlt">Bright</span> Sides of Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chu, X.; Yu, Z.; Fong, W.; Chen, C.; Zhao, J.; Huang, W.; Roberts, B. R.; Fuller-Rowell, T. J.; Richmond, A. D.; Gerrard, A. J.; Weatherwax, A. T.; Gardner, C. S.</p> <p>2014-12-01</p> <p>The polar mesosphere and thermosphere provide a unique natural laboratory for studying the complex physical, chemical, neutral dynamical and electrodynamics processes in the Earth's atmosphere and space environment. McMurdo (geographic 77.83S, geomagnetic 80S) is located by the poleward edge of the aurora oval; so energetic particles may penetrate into the lower thermosphere and mesosphere along nearly vertical geomagnetic field lines. Lidar <span class="hlt">observations</span> at McMurdo from December 2010 to 2014 have discovered several neutral atmosphere phenomena closely related to ionosphereic parameters and geomagnetic activity. For example, the diurnal tidal amplitude of <span class="hlt">temperatures</span> not only increases super-exponentially from 100 to 110 km but also its growth rate becomes larger at larger Kp index. The lidar discovery of neutral iron (Fe) layers with gravity wave signatures in the thermosphere enabled the direct measurements of neutral <span class="hlt">temperatures</span> from 30 to 170 km, revealing the neutral-ion coupling and aurora-enhanced Joule heating. A lidar 'marathon' of 174-hour continuous <span class="hlt">observations</span> showed dramatic changes of composition (Fe atoms and ice particles) densities (over 40 times) in the mesopause region and their correlations to solar events. In this paper we will study the plasma-neutral coupling on the dark side of Antarctica via <span class="hlt">observation</span> analysis and numerical modeling of the thermospheric Fe layers in the 100-200 km. A newly developed thermospheric Fe/Fe+ model is used to quantify how Fe+ ions are transported from their main deposition region to the E-F region and then neutralized to form Fe layers under dark polar conditions. We will also study the plasma-neutral coupling on the <span class="hlt">bright</span> side of Antarctica via analyzing Fe events in summer. Complementary <span class="hlt">observations</span> will be combined to show how the extreme changes of Fe layers are related to aurora particle precipitation and visible/sub-visible ice particles. These <span class="hlt">observations</span> and studies will open new areas of</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 surface and atmosphere since the 1950's. The emission is generally understood to come from a combination of emission and absorption in the subsurface, surface, 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> <span class="hlt">temperatures</span> 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" surface (e.g., similar to the Moon or Earth) [1-3].No simple surface 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> <span class="hlt">temperature</span> 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/1987ApJ...317..992J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987ApJ...317..992J"><span>Surface <span class="hlt">brightness</span> profiles of 10 comets</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jewitt, D. C.; Meech, K. J.</p> <p>1987-06-01</p> <p>CCD photometric <span class="hlt">observations</span> 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 <span class="hlt">brightness</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017A%26A...604A.111B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017A%26A...604A.111B"><span>Probing the innermost regions of AGN jets and their magnetic fields with RadioAstron. II. <span class="hlt">Observations</span> of 3C 273 at minimum activity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bruni, G.; Gómez, J. L.; Casadio, C.; Lobanov, A.; Kovalev, Y. Y.; Sokolovsky, K. V.; Lisakov, M. M.; Bach, U.; Marscher, A.; Jorstad, S.; Anderson, J. M.; Krichbaum, T. P.; Savolainen, T.; Vega-García, L.; Fuentes, A.; Zensus, J. A.; Alberdi, A.; Lee, S.-S.; Lu, R.-S.; Pérez-Torres, M.; Ros, E.</p> <p>2017-08-01</p> <p>Context. RadioAstron is a 10 m orbiting radio telescope mounted on the Spektr-R satellite, launched in 2011, performing Space Very Long Baseline Interferometry (SVLBI) <span class="hlt">observations</span> supported by a global ground array of radio telescopes. With an apogee of 350 000 km, it is offering for the first time the possibility to perform μas-resolution imaging in the cm-band. Aims: The RadioAstron active galactic nuclei (AGN) polarization Key Science Project (KSP) aims at exploiting the unprecedented angular resolution provided by RadioAstron to study jet launching/collimation and magnetic-field configuration in AGN jets. The targets of our KSP are some of the most powerful blazars in the sky. Methods: We present <span class="hlt">observations</span> at 22 GHz of 3C 273, performed in 2014, designed to reach a maximum baseline of approximately nine Earth diameters. Reaching an angular resolution of 0.3 mas, we study a particularly low-activity state of the source, and estimate the nuclear region <span class="hlt">brightness</span> <span class="hlt">temperature</span>, comparing with the extreme one detected one year before during the RadioAstron early science period. We also make use of the VLBA-BU-BLAZAR survey data, at 43 GHz, to study the kinematics of the jet in a 1.5-yr time window. Results: We find that the nuclear <span class="hlt">brightness</span> <span class="hlt">temperature</span> is two orders of magnitude lower than the exceptionally high value detected in 2013 with RadioAstron at the same frequency (1.4 × 1013 K, source-frame), and even one order of magnitude lower than the equipartition value. The kinematics analysis at 43 GHz shows that a new component was ejected 2 months after the 2013 epoch, visible also in our 22 GHz map presented here. Consequently this was located upstream of the core during the <span class="hlt">brightness</span> <span class="hlt">temperature</span> peak. Fermi-LAT <span class="hlt">observations</span> for the period 2010-2014 do not show any γ-ray flare in conjunction with the passage of the new component by the core at 43 GHz. Conclusions: These <span class="hlt">observations</span> confirm that the previously detected extreme <span class="hlt">brightness</span> <span class="hlt">temperature</span> in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMGC51D1016S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMGC51D1016S"><span>Quantifying <span class="hlt">Observed</span> <span class="hlt">Temperature</span> Extremes in the Southeastern United States</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sura, P.; Stefanova, L. B.; Griffin, M.; Worsnop, R.</p> <p>2011-12-01</p> <p>There is broad consensus that the most hazardous effects of climate change are related to a potential increase (in frequency and/or intensity) of extreme weather and climate events. In particular, the statistics of regional daily <span class="hlt">temperature</span> extremes are of practical interest for the agricultural community and energy suppliers. This is notably true for the Southeastern United States where winter hard freezes are a relatively rare and potentially catastrophic event. Here we use a long record of quality-controlled <span class="hlt">observations</span> collected from 272 National Weather Service (NWS) Cooperative <span class="hlt">Observing</span> Network (COOP) stations throughout Florida, Georgia, Alabama, and South and North Carolina to provide a detailed climatology of <span class="hlt">temperature</span> extremes in the Southeastern United States. We employ two complementary approaches. First, we analyze the effect of El Nino-Southern Oscillation (ENSO) and the Arctic Oscillation (AO) on the non-Gaussian (i.e. higher order) statistics of wintertime daily minimum and maximum <span class="hlt">temperatures</span>. We find a significant and spatially varying impact of ENSO and AO on the non-Gaussian statistics of daily maximum and minimum <span class="hlt">temperatures</span> throughout the domain. Second, the extremes of the <span class="hlt">temperature</span> distributions are studied by calculating the 1st and 99th percentiles, and then analyzing the number of days with record low/high <span class="hlt">temperatures</span> per season. This analysis of daily <span class="hlt">temperature</span> extremes reveals oscillating, multi-decadal patterns with spatially varying centers of action.</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.ncbi.nlm.nih.gov/pubmed/24522258','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24522258"><span>High <span class="hlt">brightness</span> phosphorescent organic light emitting diodes on transparent and flexible cellulose films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Purandare, Sumit; Gomez, Eliot F; Steckl, Andrew J</p> <p>2014-03-07</p> <p>Organic light-emitting diodes (OLED) were fabricated on flexible and transparent reconstituted cellulose obtained from wood pulp. Cellulose is naturally available, abundant, and biodegradable and offers a unique substrate alternative for the fabrication of flexible OLEDs. Transparent cellulose material was formed by dissolution of cellulose in an organic solvent (dimethyl acetamide) at elevated <span class="hlt">temperature</span> (165 °C) in the presence of a salt (LiCl). The optical transmission of 40-μm thick transparent cellulose sheet averaged 85% over the visible spectrum. High <span class="hlt">brightness</span> and high efficiency thin film OLEDs were fabricated on transparent cellulose films using phosphorescent Ir(ppy)3 as the emitter material. The OLEDs achieved current and luminous emission efficiencies as high as 47 cd A(-1) and 20 lm W(-1), respectively, and a maximum <span class="hlt">brightness</span> of 10,000 cd m(-2).</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>