Sample records for airglow imaging observations

  1. The advances in airglow study and observation by the ground-based airglow observation network over China

    NASA Astrophysics Data System (ADS)

    Xu, Jiyao; Li, Qinzeng; Yuan, Wei; Liu, Xiao; Liu, Weijun; Sun, Longchang

    2017-04-01

    Ground-based airglow observation networks over China used to study airglow have been established, which contains 15 stations. Some new results were obtained using the networks. For OH airglow observations, firstly, an unusual outbreak of Concentric Gravity Wave (CGW) events were observed by the first no-gap network nearly every night during the first half of August 2013. Combination of the ground imager network with satellites provides multilevel observations of the CGWs from the troposphere to the mesopause region. Secondly, three-year OH airglow images (2012-2014) from Qujing (25.6°N, 103.7°E) were used to study how orographic features of the Tibetan Plateau (TP) affect the geographical distributions of gravity wave (GW) sources. We find the orographic forcings have a significant impact on the gravity wave propagation features. Thirdly, ground-based observations of the OH (9-4, 8-3, 6-2, 5-1, 3-0) band airglow over Xinglong (40°2N, 117°4E) in northern China from 2012 to 2014 are used to calculate rotational temperatures. By comparing the ground-based OH rotational temperature with SABER's observations, five Einstein coefficient datasets are evaluated. We find rotational temperatures determined using any of the available Einstein coefficient datasets have systematic errors. We have obtained a set of optimal Einstein coefficients ratios for rotational temperature derivation using three years data from ground-based OH spectra and SABER temperatures. For the OI 630.0 nm airglow observations, we used three-year (2011-2013) observations of thermospheric winds (at 250 km) by Fabry-Perot interferometers at Xinglong to study the climatology of atmospheric planetary wave-type oscillations (PWTOs) with periods of 4-19 days. We found these PWTOs occur more frequently in the months from May to October. They are consistent with the summertime preference of middle-latitude ionospheric electron density oscillations noted in other studies. By using an all-sky airglow imager

  2. MANGO Imager Network Observations of Geomagnetic Storm Impact on Midlatitude 630 nm Airglow Emissions

    NASA Astrophysics Data System (ADS)

    Kendall, E. A.; Bhatt, A.

    2017-12-01

    The Midlatitude Allsky-imaging Network for GeoSpace Observations (MANGO) is a network of imagers filtered at 630 nm spread across the continental United States. MANGO is used to image large-scale airglow and aurora features and observes the generation, propagation, and dissipation of medium and large-scale wave activity in the subauroral, mid and low-latitude thermosphere. This network consists of seven all-sky imagers providing continuous coverage over the United States and extending south into Mexico. This network sees high levels of medium and large scale wave activity due to both neutral and geomagnetic storm forcing. The geomagnetic storm observations largely fall into two categories: Stable Auroral Red (SAR) arcs and Large-scale traveling ionospheric disturbances (LSTIDs). In addition, less-often observed effects include anomalous airglow brightening, bright swirls, and frozen-in traveling structures. We will present an analysis of multiple events observed over four years of MANGO network operation. We will provide both statistics on the cumulative observations and a case study of the "Memorial Day Storm" on May 27, 2017.

  3. Statistical analysis of gravity waves characteristics observed by airglow imaging at Syowa Station (69S, 39E), Antarctica

    NASA Astrophysics Data System (ADS)

    Matsuda, Takashi S.; Nakamura, Takuji; Shiokawa, Kazuo; Tsutsumi, Masaki; Suzuki, Hidehiko; Ejiri, Mitsumu K.; Taguchi, Makoto

    Atmospheric gravity waves (AGWs), which are generated in the lower atmosphere, transport significant amount of energy and momentum into the mesosphere and lower thermosphere and cause the mean wind accelerations in the mesosphere. This momentum deposit drives the general circulation and affects the temperature structure. Among many parameters to characterize AGWs, horizontal phase velocity is very important to discuss the vertical propagation. Airglow imaging is a useful technique for investigating the horizontal structures of AGWs at around 90 km altitude. Recently, there are many reports about statistical characteristics of AGWs observed by airglow imaging. However, comparison of these results obtained at various locations is difficult because each research group uses its own method for extracting and analyzing AGW events. We have developed a new statistical analysis method for obtaining the power spectrum in the horizontal phase velocity domain from airglow image data, so as to deal with huge amounts of imaging data obtained on different years and at various observation sites, without bias caused by different event extraction criteria for the observer. This method was applied to the data obtained at Syowa Station, Antarctica, in 2011 and compared with a conventional event analysis in which the phase fronts were traced manually in order to estimate horizontal characteristics. This comparison shows that our new method is adequate to deriving the horizontal phase velocity characteristics of AGWs observed by airglow imaging technique. We plan to apply this method to airglow imaging data observed at Syowa Station in 2002 and between 2008 and 2013, and also to the data observed at other stations in Antarctica (e.g. Rothera Station (67S, 68W) and Halley Station (75S, 26W)), in order to investigate the behavior of AGWs propagation direction and source distribution in the MLT region over Antarctica. In this presentation, we will report interim analysis result of the data

  4. Comparison with the horizontal phase velocity distribution of gravity waves observed airglow imaging data of different sampling periods

    NASA Astrophysics Data System (ADS)

    Matsuda, T. S.; Nakamura, T.; Ejiri, M. K.; Tsutsumi, M.; Shiokawa, K.

    2014-12-01

    Atmospheric gravity waves (AGWs), which are generated in the lower atmosphere, transport significant amount of energy and momentum into the mesosphere and lower thermosphere. Among many parameters to characterize AGWs, horizontal phase velocity is very important to discuss the vertical propagation. Airglow imaging is a useful technique for investigating the horizontal structures of AGWs around mesopause. There are many airglow imagers operated all over the world, and a large amount of data which could improve our understanding of AGWs propagation direction and source distribution in the MLT region. We have developed a new statistical analysis method for obtaining the power spectrum in the horizontal phase velocity domain (phase velocity spectrum), from airglow image data, so as to deal with huge amounts of imaging data obtained on different years and at various observation sites, without bias caused by different event extraction criteria for the observer. From a series of images projected onto the geographic coordinates, 3-D Fourier transform is applied and 3-D power spectrum in horizontal wavenumber and frequency domain is obtained. Then, it is converted into phase velocity and frequency domain. Finally, the spectrum is integrated along the frequency for the range of interest and 2-D spectrum in horizontal phase velocity is calculated. This method was applied to the data obtained at Syowa Station (69ºS, 40ºE), Antarctica, in 2011 and compared with a conventional event analysis in which the phase fronts were traced manually in order to estimate horizontal propagation characteristics. This comparison shows that our new method is adequate to deriving the horizontal phase velocity characteristics of AGWs observed by airglow imaging technique. Airglow imaging observation has been operated with various sampling intervals. We also presents how the images with different sample interval should be treated.

  5. Telescopic Imaging of Heater-Induced Airglow at HAARP

    DTIC Science & Technology

    2007-01-01

    03-01-2007 Final1 10-09-2003 - 10-09-2006 4. TITLE AND SUBTITLE Ba. CONTRACT NUMBER Telescopic Imaging of Heater-Induced Airglow at HAARP N00014-03-1... HAARP to optically measure fine structure in the ionosphere and to study airglow sources. In the presence of aurora and a strong blanketing E layer... HAARP was modulated at intervals of several seconds. For several cycles, small bright airglow spots were observed whenever HAARP was on. These spots

  6. Nighttime Medium-Scale Traveling Ionospheric Disturbances From Airglow Imager and Global Navigation Satellite Systems Observations

    NASA Astrophysics Data System (ADS)

    Huang, Fuqing; Lei, Jiuhou; Dou, Xiankang; Luan, Xiaoli; Zhong, Jiahao

    2018-01-01

    In this study, coordinated airglow imager, GPS total electron content (TEC), and Beidou geostationary orbit (GEO) TEC observations for the first time are used to investigate the characteristics of nighttime medium-scale traveling ionospheric disturbances (MSTIDs) over central China. The results indicated that the features of nighttime MSTIDs from three types of observations are generally consistent, whereas the nighttime MSTID features from the Beidou GEO TEC are in better agreement with those from airglow images as compared with the GPS TEC, given that the nighttime MSTID characteristics from GPS TEC are significantly affected by Doppler effect due to satellite movement. It is also found that there are three peaks in the seasonal variations of the occurrence rate of nighttime MSTIDs in 2016. Our study revealed that the Beidou GEO satellites provided fidelity TEC observations to study the ionospheric variability.

  7. Gravity Wave Detection through All-sky Imaging of Airglow

    NASA Astrophysics Data System (ADS)

    Nguyen, T. V.; Martinez, A.; Porat, I.; Hampton, D. L.; Bering, E., III; Wood, L.

    2017-12-01

    Airglow, the faint glow of the atmosphere, is caused by the interaction of air molecules with radiation from the sun. Similarly, the aurora is created by interactions of air molecules with the solar wind. It has been shown that airglow emissions are altered by gravity waves passing through airglow source region (100-110km), making it possible to study gravity waves and their sources through airglow imaging. University of Houston's USIP - Airglow team designed a compact, inexpensive all-sky imager capable of detecting airglow and auroral emissions using a fisheye lens, a simple optical train, a filter wheel with 4 specific filters, and a CMOS camera. This instrument has been used in USIP's scientific campaign in Alaska throughout March 2017. During this period, the imager captured auroral activity in the Fairbanks region. Due to lunar conditions and auroral activity images from the campaign did not yield visible signs of airglow. Currently, the team is trying to detect gravity wave patterns present in the images through numerical analysis. Detected gravity wave patterns will be compared to local weather data, and may be used to make correlations between gravity waves and weather events. Such correlations could provide more data on the relationship between the mesosphere and lower layers of the atmosphere. Practical applications of this research include weather prediction and detection of air turbulence.

  8. WINDII atmospheric wave airglow imaging

    NASA Technical Reports Server (NTRS)

    Armstrong, W. T.; Hoppe, U.-P.; Solheim, B. H.; Shepherd, G. G.

    1996-01-01

    Preliminary WINDII nighttime airglow wave-imaging data in the UARS rolldown attitude has been analyzed with the goal to survey gravity waves near the upper boundary of the middle atmosphere. Wave analysis is performed on O[sub 2](0,0) emissions from a selected 1[sup 0] x 1[sup 0] oblique view of the airglow layer at approximately 95 km altitude, which has no direct earth background and only an atmospheric background which is optically thick for the 0[sub 2](0,0) emission. From a small data set, orbital imaging of atmospheric wave structures is demonstrated, with indication of large variations in wave activity across land and sea. Comparison ground-based imagery is discussed with respect to similarity of wave variations across land/sea boundaries and future orbital mosaic image construction.

  9. First OH Airglow Observation of Mesospheric Gravity Waves Over European Russia Region

    NASA Astrophysics Data System (ADS)

    Li, Qinzeng; Yusupov, Kamil; Akchurin, Adel; Yuan, Wei; Liu, Xiao; Xu, Jiyao

    2018-03-01

    For the first time, we perform a study of mesospheric gravity waves (GWs) for four different seasons of 1 year in the latitudinal band from 45°N to 75°N using an OH all-sky airglow imager over Kazan (55.8°N, 49.2°E), Russia, during the period of August 2015 to July 2016. Our observational study fills a huge airglow imaging observation gap in Europe and Russia region. In total, 125 GW events and 28 ripple events were determined by OH airglow images in 98 clear nights. The observed GWs showed a strong preference of propagation toward northeast in all seasons, which was significantly different from airglow imager observations at other latitudes that the propagation directions were seasonal dependent. The middle atmosphere wind field is used to explain the lack of low phase speed GWs since these GWs were falling into the blocking region due to the filtering effects. Deep tropospheric convections derived from the European Centre for Medium-Range Weather Forecasts reanalysis data are determined near Caucasus Mountains region, which suggests that the convections are the dominant source of the GWs in spring, summer, and autumn seasons. This finding extends our knowledge that convection might also be an important source of GWs in the higher latitudes. In winter the generation mechanism of the GWs are considered to be jet stream systems. In addition, the occurrence frequency of ripple is much lower than other stations. This study provides some constraints on the range of GW parameters in GW parameterization in general circulation models in Europe and Russia region.

  10. New statistical analysis of the horizontal phase velocity distribution of gravity waves observed by airglow imaging

    NASA Astrophysics Data System (ADS)

    Matsuda, Takashi S.; Nakamura, Takuji; Ejiri, Mitsumu K.; Tsutsumi, Masaki; Shiokawa, Kazuo

    2014-08-01

    We have developed a new analysis method for obtaining the power spectrum in the horizontal phase velocity domain from airglow intensity image data to study atmospheric gravity waves. This method can deal with extensive amounts of imaging data obtained on different years and at various observation sites without bias caused by different event extraction criteria for the person processing the data. The new method was applied to sodium airglow data obtained in 2011 at Syowa Station (69°S, 40°E), Antarctica. The results were compared with those obtained from a conventional event analysis in which the phase fronts were traced manually in order to estimate horizontal characteristics, such as wavelengths, phase velocities, and wave periods. The horizontal phase velocity of each wave event in the airglow images corresponded closely to a peak in the spectrum. The statistical results of spectral analysis showed an eastward offset of the horizontal phase velocity distribution. This could be interpreted as the existence of wave sources around the stratospheric eastward jet. Similar zonal anisotropy was also seen in the horizontal phase velocity distribution of the gravity waves by the event analysis. Both methods produce similar statistical results about directionality of atmospheric gravity waves. Galactic contamination of the spectrum was examined by calculating the apparent velocity of the stars and found to be limited for phase speeds lower than 30 m/s. In conclusion, our new method is suitable for deriving the horizontal phase velocity characteristics of atmospheric gravity waves from an extensive amount of imaging data.

  11. Enhanced airglow at Titan

    NASA Astrophysics Data System (ADS)

    Royer, Emilie; Esposito, Larry; Wahlund, Jan-Erik

    2016-06-01

    The Cassini Ultraviolet Imaging Spectrograph (UVIS) instrument made thousand of observations of Titan since its arrival in the Saturnian system in 2004, but only few of them have been analyzed yet. Using the imaging capability of UVIS combined to a big data analytics approach, we have been able to uncover an unexpected pattern in this observations: on several occasions the Titan airglow exhibits an enhanced brightness by approximately a factor of 2, generally combined with a lower altitude of the airglow emission peak. These events typically last from 10 to 30 minutes and are followed and preceded by an airglow of regular and expected level of brightness and altitude. Observations made by the Cassini Plasma Spectrometer (CAPS) instrument onboard Cassini allowed us to correlate the enhanced airglow observed on T-32 with an electron burst. The timing of the burst and the level of energetic electrons (1 keV) observed by CAPS correspond to a brighter and lower than typical airglow displayed on the UVIS data. Furthermore, during T-32 Titan was inside the Saturn's magnetosheath and thus more subject to bombardment by energetic particles. However, our analysis demonstrates that the presence of Titan inside the magnetosheath is not a necessary condition for the production of an enhanced airglow, as we detected other similar events while Titan was within Saturn's magnetosphere. The study presented here aims to a better understanding of the interactions of Titan's upper atmosphere with its direct environment.

  12. Detection of large-scale concentric gravity waves from a Chinese airglow imager network

    NASA Astrophysics Data System (ADS)

    Lai, Chang; Yue, Jia; Xu, Jiyao; Yuan, Wei; Li, Qinzeng; Liu, Xiao

    2018-06-01

    Concentric gravity waves (CGWs) contain a broad spectrum of horizontal wavelengths and periods due to their instantaneous localized sources (e.g., deep convection, volcanic eruptions, or earthquake, etc.). However, it is difficult to observe large-scale gravity waves of >100 km wavelength from the ground for the limited field of view of a single camera and local bad weather. Previously, complete large-scale CGW imagery could only be captured by satellite observations. In the present study, we developed a novel method that uses assembling separate images and applying low-pass filtering to obtain temporal and spatial information about complete large-scale CGWs from a network of all-sky airglow imagers. Coordinated observations from five all-sky airglow imagers in Northern China were assembled and processed to study large-scale CGWs over a wide area (1800 km × 1 400 km), focusing on the same two CGW events as Xu et al. (2015). Our algorithms yielded images of large-scale CGWs by filtering out the small-scale CGWs. The wavelengths, wave speeds, and periods of CGWs were measured from a sequence of consecutive assembled images. Overall, the assembling and low-pass filtering algorithms can expand the airglow imager network to its full capacity regarding the detection of large-scale gravity waves.

  13. Thermospheric Airglow Perturbations in the Upper Atmosphere Caused by Hurricane Harvey

    NASA Astrophysics Data System (ADS)

    Bhatt, A.; Kendall, E. A.

    2017-12-01

    The Midlatitude Allsky imaging Network for Geophysical Observations (MANGO) consists of seven allsky imagers distributed across the United States recording observations of large-scale airglow perturbations. The imagers are filtered at 630 nm, a forbidden oxygen line, in order to record the predominant source of airglow at 250 km altitude. While the ubiquitous airglow layer is challenging to observe when under uniform conditions, waves in the upper atmosphere cause ripples in the airglow layer which can easily be imaged by appropriate instrumentation. MANGO is the first network to record perturbations in the airglow layer on a continent-size scale. Large and Mid-scale Traveling Ionospheric Disturbances (LSTIDs and MSTIDs) are recorded that are caused by auroral forcing, mountain turbulence, and tidal variations. On August 25, airglow perturbations centered on the Hurricane Harvey path were observed by MANGO. These images and connections to other complimentary data sets such as GPS will be presented.

  14. Imaging observations of lower thermospheric O(1S) and O2 airglow emissions from STS 9 - Implications of height variations

    NASA Technical Reports Server (NTRS)

    Swenson, G. R.; Mende, S. B.; Llewellyn, E. J.

    1989-01-01

    The lower thermospheric nightglow in the Southern Hemisphere was observed with the Atmospheric Emissions Photometric Imager during the Spacelab 1 mission in December, 1983. Observations of emission from O(1S) at 2972 and 5577A, O2 at 7620 A, OH near 6300 A, and the combined emission from the three upper states of O2 which lead to the Herzberg I and II and Chamberlain band emissions in B and near UV are discussed. The altitudes of peak emission heights are determined, showing that the peak heights are not constant with latitude. It is found that airglow heights varied with latitude by as much as 8 km. The observed airglow height pattern near the equator is similar to that of Wasser and Donahue (1979).

  15. First spaceborne observation of the entire concentric airglow structure caused by tropospheric disturbance

    NASA Astrophysics Data System (ADS)

    Akiya, Y.; Saito, A.; Sakanoi, T.; Hozumi, Y.; Yamazaki, A.; Otsuka, Y.; Nishioka, M.; Tsugawa, T.

    2014-10-01

    Spaceborne imagers are able to observe the airglow structures with wide field of views regardless of the tropospheric condition that limits the observational time of the ground-based imagers. Concentric wave structures of the O2 airglow in 762 nm wavelength were observed over North America on 1 June 2013 from the International Space Station. This was the first observation in which the entire image of the structure was captured from space, and its spatial scale size was determined to be 1200 km radius without assumptions. The apparent horizontal wavelength was 80 km, and the amplitude in the intensity was approximately 20% of the background intensity. The propagation velocity of the structure was derived as 125 ± 62 m/s and atmospheric gravity waves were estimated to be generated for 3.5 ± 1.7 h. Concentric structures observed in this event were interpreted to be generated by super cells that caused a tornado in its early phase.

  16. Issues in Quantitative Analysis of Ultraviolet Imager (UV) Data: Airglow

    NASA Technical Reports Server (NTRS)

    Germany, G. A.; Richards, P. G.; Spann, J. F.; Brittnacher, M. J.; Parks, G. K.

    1999-01-01

    The GGS Ultraviolet Imager (UVI) has proven to be especially valuable in correlative substorm, auroral morphology, and extended statistical studies of the auroral regions. Such studies are based on knowledge of the location, spatial, and temporal behavior of auroral emissions. More quantitative studies, based on absolute radiometric intensities from UVI images, require a more intimate knowledge of the instrument behavior and data processing requirements and are inherently more difficult than studies based on relative knowledge of the oval location. In this study, UVI airglow observations are analyzed and compared with model predictions to illustrate issues that arise in quantitative analysis of UVI images. These issues include instrument calibration, long term changes in sensitivity, and imager flat field response as well as proper background correction. Airglow emissions are chosen for this study because of their relatively straightforward modeling requirements and because of their implications for thermospheric compositional studies. The analysis issues discussed here, however, are identical to those faced in quantitative auroral studies.

  17. Satellite-based observations of tsunami-induced mesosphere airglow perturbations

    NASA Astrophysics Data System (ADS)

    Yang, Yu-Ming; Verkhoglyadova, Olga; Mlynczak, Martin G.; Mannucci, Anthony J.; Meng, Xing; Langley, Richard B.; Hunt, Linda A.

    2017-01-01

    Tsunami-induced airglow emission perturbations were retrieved by using space-based measurements made by the Sounding of the Atmosphere using Broad-band Emission Radiometry (SABER) instrument on board the Thermosphere-Ionosphere-Mesosphere Energetics Dynamics spacecraft. At and after the time of the Tohoku-Oki earthquake on 11 March 2011, and the Chile earthquake on 16 September 2015, the spacecraft was performing scans over the Pacific Ocean. Significant ( 10% relative to the ambient emission profiles) and coherent nighttime airglow perturbations were observed in the mesosphere following Sounding of the Atmosphere using Broad-band Emission Radiometry limb scans intercepting tsunami-induced atmospheric gravity waves. Simulations of emission variations are consistent with the physical characteristics of the disturbances at the locations of the corresponding SABER scans. Airglow observations and model simulations suggest that atmospheric neutral density and temperature perturbations can lead to the observed amplitude variations and multipeak structures in the emission profiles. This is the first time that airglow emission rate perturbations associated with tsunamis have been detected with space-based measurements.

  18. Enhanced 630nm equatorial airglow emission observed by Limb Viewing Hyper Spectral Imager (LiVHySI) onboard YOUTHSAT-1

    NASA Astrophysics Data System (ADS)

    Bisht, R. S.; Thapa, N.; Babu, P. N.

    2016-04-01

    The Earth's airglow layer, when observed in the limb view mode, appears to be a double layer. LiVHySI onboard YOUTHSAT (inclination 98.730, apogee 817 km, launched by Indian Space Research Organization in April, 2011) is an Earth's limb viewing camera measuring airglow emissions in the spectral window of 550-900 nm. Total altitude coverage is about 500 km with command selectable lowest altitude. During few of the orbits we have observed the double layer structure and obtained absolute spectral intensity and altitude profile for 630 nm airglow emission. Our night time observations of upper atmosphere above dip equator carried out on 3rd May, 2011 show a prominent 630 nm double layer structure. The upper airglow layer consists of the 630 nm atomic oxygen O(1D) emission line and lower layer consists of OH(9-3) meinel band emission at 630 nm. The volume emission rate as a function of altitude is simulated for our observational epoch and the modeled limb intensity distribution is compared with the observations. The observations are in good agreement with the simulated intensity distribution.

  19. 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 "bright nights." Currently, scientists study <span class="hlt">airglow</span>, an emission of light from the high atmosphere, 100 km above us. Satellite <span class="hlt">observations</span> of a green <span class="hlt">airglow</span> 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 <span class="hlt">airglow</span> brightening, and this article demonstrates that this event produces a bright 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('http://adsabs.harvard.edu/abs/2003EAEJA....13380K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003EAEJA....13380K"><span>OH line selection for nadir <span class="hlt">airglow</span> gravity wave <span class="hlt">imaging</span> in the auroral zone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumer, J. B.; Hecht, J.; Geballe, T. R.; Mergenthaler, J. L.; Rinaldi, M.; Claflin, E. S.; Swenson, G. R.</p> <p>2003-04-01</p> <p>For satellite borne nadir OH <span class="hlt">airglow</span> wave <span class="hlt">imaging</span> in the auroral zone the <span class="hlt">observed</span> lines must be strong enough to give good signal to noise, coincident with strong atmospheric absorption lines to suppress structure in the <span class="hlt">image</span> due to reflection of <span class="hlt">airglow</span> and moonlight from tops of clouds and from high altitude terrain, and in a spectral region coincident with relatively weak aurora that its contribution to the <span class="hlt">observed</span> structure can be corrected by data obtained in a guard band containing relatively strong auroral emission, and relatively weak, or no <span class="hlt">airglow</span>. OH <span class="hlt">airglow</span> spectra <span class="hlt">observed</span> from high altitude, in our case Mauna Kea by the UKIRT CGS4 grating instrument, (see website http://www.jach.hawaii.edu/JACpublic/UKIRT/instruments/cgs4/maunakea/ohlines.html) provide an opportunity to identify lines that ARE NOT <span class="hlt">observed</span> at that high altitude. These are most absorbed in the earths atmosphere. These occur in the regions near 1400 and 1900 nm of strong water vapor absorption. Our preliminary determination is that the 7-5 p1(2) line at 1899.01 nm and the p1(3) at 1911.41 nm are the best candidates. These are missing in the <span class="hlt">observed</span> spectra, and this is confirmed by running FASCODE transmission calculations from top of Mauna Kea to space at .01 cm-1 resolution. Similar calculations for conditions at which the high resolution Kitt peak atlas data were taken confirmed the calculations. OH line positions and relative strengths within the band were derived from the HITRAN data base, and transmitted lines in the 7-5 band were used to determine the strength of these lines. Each are the order 10 kR, and are about four to six times brighter than atmospheric absorbed candidate lines in the 1400 nm region. Also, the aurora in the 1900nm region is considerably weaker than in the 1400nm region. In fact the region 1351 to 1358 contains relatively strong aurora, and practically no <span class="hlt">airglow</span>, and is candidate for an instrumental auroral guard band. The nadir <span class="hlt">imaging</span> instrument which</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li class="active"><span>1</span></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_1 --> <div id="page_2" 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_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><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="21"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003090&hterms=plasma&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dplasma','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003090&hterms=plasma&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dplasma"><span>Hemispheric Asymmetry in Transition from Equatorial Plasma Bubble to Blob as Deduced from 630.0 nm <span class="hlt">Airglow</span> <span class="hlt">Observations</span> at Low Latitudes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Park, Jaeheung; Martinis, Carlos R.; Luehr, Hermann; Pfaff, Robert F.; Kwak, Young-Sil</p> <p>2016-01-01</p> <p>Transitions from depletions to enhancements of 630.0 nm nighttime <span class="hlt">airglow</span> have been <span class="hlt">observed</span> at Arecibo. Numerical simulations by Krall et al. (2009) predicted that they should occur only in one hemisphere, which has not yet been confirmed <span class="hlt">observationally</span>. In this study we investigate the hemispheric conjugacy of the depletion-to-enhancement transition using multiple instruments. We focus on one event <span class="hlt">observed</span> in the American longitude sector on 22 December 2014: 630.0 nm <span class="hlt">airglow</span> depletions evolved into enhancements in the Northern Hemisphere while the evolution did not occur in the conjugate location in the Southern Hemisphere. Concurrent plasma density measured by low Earth orbit (LEO) satellites and 777.4 nm <span class="hlt">airglow</span> <span class="hlt">images</span> support that the depletions and enhancements of 630.0 nm night time <span class="hlt">airglow</span> reflect plasma density decreases and increases (blobs), respectively. Characteristics of the <span class="hlt">airglow</span> depletions, in the context of the LEO satellite data, further suggest that the plasma density depletion deduced from the <span class="hlt">airglow</span> data represents equatorial plasma bubbles (EPBs) rather than medium-scale traveling ionospheric disturbances from midlatitudes. Hence, the event in this study can be interpreted as EPB-to-blob transition.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRD..121..650C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRD..121..650C"><span>Intermittency of gravity wave momentum flux in the mesopause region <span class="hlt">observed</span> with an all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Bing; Liu, Alan Z.</p> <p>2016-01-01</p> <p>The intermittency of gravity wave momentum flux (MF) near the OH <span class="hlt">airglow</span> layer (˜87 km) in the mesopause region is investigated for the first time using <span class="hlt">observation</span> of all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span> over Maui, Hawaii (20.7°N, 156.3°W), and Cerro Pachón, Chile (30.3°S, 70.7°W). At both sites, the probability density function (pdf) of gravity wave MF shows two distinct distributions depending on the magnitude of the MF. For MF smaller (larger) than ˜16 m2 s-2 (0.091 mPa), the pdf follows a lognormal (power law) distribution. The intermittency represented by the Bernoulli proxy and the percentile ratio shows that gravity waves have higher intermittency at Maui than at Cerro Pachón, suggesting more intermittent background variation above Maui. It is found that most of the MF is contributed by waves that occur very infrequently. But waves that individually contribute little MF are also important because of their higher occurrence frequencies. The peak contribution is from waves with MF around ˜2.2 m2 s-2 at Cerro Pachón and ˜5.5 m2 s-2 at Maui. Seasonal variations of the pdf and intermittency imply that the background atmosphere has larger influence on the <span class="hlt">observed</span> intermittency in the mesopause region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016cosp...41E1065K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E1065K"><span>Mesopause region wind, temperature and <span class="hlt">airglow</span> irradiance above Eureka, Nunavut</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kristoffersen, Samuel; Ward, William E.; Vail, Christopher; Shepherd, Marianna</p> <p>2016-07-01</p> <p>The PEARL All Sky <span class="hlt">Imager</span> (PASI, <span class="hlt">airglow</span> <span class="hlt">images</span>), the Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span> (SATI, <span class="hlt">airglow</span> irradiance and temperature) and the E-Region Wind Interferometer II (ERWIN2, wind, <span class="hlt">airglow</span> irradiance and temperature) are co-located at the Polar Environment Atmospheric Research Laboratory (PEARL)in Eureka, Nunavut (80 N, 86 W). These instruments view the wind, temperature and <span class="hlt">airglow</span> irradiance of hydroxyl (all three) O2 (ERWIN2 and SATI), sodium (PASI), and oxygen green line (PASI and ERWIN2). The viewing locations and specific emissions of the various instruments differ. Nevertheless, the co-location of these instruments provides an excellent opportunity for case studies of specific events and for intercomparison between the different techniques. In this paper we discuss the approach we are using to combine <span class="hlt">observations</span> from the different instruments. Case studies show that at times the various instruments are in good agreement but at other times they differ. Of particular interest are situations where gravity wave signatures are evident for an extended period of time and one such situation is presented. The discussion includes consideration of the filtering effect of viewing through <span class="hlt">airglow</span> layers and the extent to which wind, <span class="hlt">airglow</span> and temperature variations can be associated with the same gravity wave.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EP%26S...68..155H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EP%26S...68..155H"><span>Calibration of <span class="hlt">imaging</span> parameters for space-borne <span class="hlt">airglow</span> photography using city light positions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hozumi, Yuta; Saito, Akinori; Ejiri, Mitsumu K.</p> <p>2016-09-01</p> <p>A new method for calibrating <span class="hlt">imaging</span> parameters of photographs taken from the International Space Station (ISS) is presented in this report. <span class="hlt">Airglow</span> in the mesosphere and the F-region ionosphere was captured on the limb of the Earth with a digital single-lens reflex camera from the ISS by astronauts. To utilize the photographs as scientific data, <span class="hlt">imaging</span> parameters, such as the angle of view, exact position, and orientation of the camera, should be determined because they are not measured at the time of <span class="hlt">imaging</span>. A new calibration method using city light positions shown in the photographs was developed to determine these <span class="hlt">imaging</span> parameters with high accuracy suitable for <span class="hlt">airglow</span> study. Applying the pinhole camera model, the apparent city light positions on the photograph are matched with the actual city light locations on Earth, which are derived from the global nighttime stable light map data obtained by the Defense Meteorological Satellite Program satellite. The correct <span class="hlt">imaging</span> parameters are determined in an iterative process by matching the apparent positions on the <span class="hlt">image</span> with the actual city light locations. We applied this calibration method to photographs taken on August 26, 2014, and confirmed that the result is correct. The precision of the calibration was evaluated by comparing the results from six different photographs with the same <span class="hlt">imaging</span> parameters. The precisions in determining the camera position and orientation are estimated to be ±2.2 km and ±0.08°, respectively. The 0.08° difference in the orientation yields a 2.9-km difference at a tangential point of 90 km in altitude. The <span class="hlt">airglow</span> structures in the photographs were mapped to geographical points using the calibrated <span class="hlt">imaging</span> parameters and compared with a simultaneous <span class="hlt">observation</span> by the Visible and near-Infrared Spectral <span class="hlt">Imager</span> of the Ionosphere, Mesosphere, Upper Atmosphere, and Plasmasphere mapping mission installed on the ISS. The comparison shows good agreements and supports the validity</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17794901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17794901"><span>Night <span class="hlt">Airglow</span> <span class="hlt">Observations</span> from Orbiting Spacecraft Compared with Measurements from Rockets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Koomen, M J; Gulledge, I S; Packer, D M; Tousey, R</p> <p>1963-06-07</p> <p>A luminous band around the night-time horizon, <span class="hlt">observed</span> from orbiting capsules by J. H. Glenn and M. S. Carpenter, and identified as the horizon enhancement of the night <span class="hlt">airglow</span>, is detected regularly in rocket-borne studies of night <span class="hlt">airglow</span>. Values of luminance and dip angle of this band derived from Carpenter's <span class="hlt">observations</span> agree remarkably well with values obtained from rocket data. The rocket results, however, do not support Carpenter's <span class="hlt">observation</span> that the emission which he saw was largely the atomic oxygen line at 5577 A, but assign the principal luminosity to the green continuum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930005120','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930005120"><span>Near-infrared oxygen <span class="hlt">airglow</span> from the Venus nightside</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Crisp, D.; Meadows, V. S.; Allen, D. A.; Bezard, B.; Debergh, C.; Maillard, J.-P.</p> <p>1992-01-01</p> <p>Groundbased <span class="hlt">imaging</span> and spectroscopic <span class="hlt">observations</span> of Venus reveal intense near-infrared oxygen <span class="hlt">airglow</span> emission from the upper atmosphere and provide new constraints on the oxygen photochemistry and dynamics near the mesopause (approximately 100 km). Atomic oxygen is produced by the Photolysis of CO2 on the dayside of Venus. These atoms are transported by the general circulation, and eventually recombine to form molecular oxygen. Because this recombination reaction is exothermic, many of these molecules are created in an excited state known as O2(delta-1). The <span class="hlt">airglow</span> is produced as these molecules emit a photon and return to their ground state. New <span class="hlt">imaging</span> and spectroscopic <span class="hlt">observations</span> acquired during the summer and fall of 1991 show unexpected spatial and temporal variations in the O2(delta-1) <span class="hlt">airglow</span>. The implications of these <span class="hlt">observations</span> for the composition and general circulation of the upper venusian atmosphere are not yet understood but they provide important new constraints on comprehensive dynamical and chemical models of the upper mesosphere and lower thermosphere of Venus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA33A2422S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA33A2422S"><span>Performance evaluation of low-cost <span class="hlt">airglow</span> cameras for mesospheric gravity wave measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suzuki, S.; Shiokawa, K.</p> <p>2016-12-01</p> <p>Atmospheric gravity waves significantly contribute to the wind/thermal balances in the mesosphere and lower thermosphere (MLT) through their vertical transport of horizontal momentum. It has been reported that the gravity wave momentum flux preferentially associated with the scale of the waves; the momentum fluxes of the waves with a horizontal scale of 10-100 km are particularly significant. <span class="hlt">Airglow</span> <span class="hlt">imaging</span> is a useful technique to <span class="hlt">observe</span> two-dimensional structure of small-scale (<100 km) gravity waves in the MLT region and has been used to investigate global behaviour of the waves. Recent studies with simultaneous/multiple <span class="hlt">airglow</span> cameras have derived spatial extent of the MLT waves. Such network <span class="hlt">imaging</span> <span class="hlt">observations</span> are advantageous to ever better understanding of coupling between the lower and upper atmosphere via gravity waves. In this study, we newly developed low-cost <span class="hlt">airglow</span> cameras to enlarge the <span class="hlt">airglow</span> <span class="hlt">imaging</span> network. Each of the cameras has a fish-eye lens with a 185-deg field-of-view and equipped with a CCD video camera (WATEC WAT-910HX) ; the camera is small (W35.5 x H36.0 x D63.5 mm) and inexpensive, much more than the <span class="hlt">airglow</span> camera used for the existing ground-based network (Optical Mesosphere Thermosphere <span class="hlt">Imagers</span> (OMTI) operated by Solar-Terrestrial Environmental Laboratory, Nagoya University), and has a CCD sensor with 768 x 494 pixels that is highly sensitive enough to detect the mesospheric OH <span class="hlt">airglow</span> emission perturbations. In this presentation, we will report some results of performance evaluation of this camera made at Shigaraki (35-deg N, 136-deg E), Japan, where is one of the OMTI station. By summing 15-<span class="hlt">images</span> (i.e., 1-min composition of the <span class="hlt">images</span>) we recognised clear gravity wave patterns in the <span class="hlt">images</span> with comparable quality to the OMTI's <span class="hlt">image</span>. Outreach and educational activities based on this research will be also reported.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA13B..07G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA13B..07G"><span>NIRAC: Near Infrared <span class="hlt">Airglow</span> Camera for the International Space Station</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gelinas, L. J.; Rudy, R. J.; Hecht, J. H.</p> <p>2017-12-01</p> <p>NIRAC is a space based infrared <span class="hlt">airglow</span> <span class="hlt">imager</span> that will be deployed to the International Space Station in late 2018, under the auspices of the Space Test Program. NIRAC will survey OH <span class="hlt">airglow</span> emissions in the 1.6 micron wavelength regime, exploring the spatial and temporal variability of emission intensities at latitudes from 51° south to 51° north. Atmospheric perturbations in the 80-100 km altitude range, including those produced by atmospheric gravity waves (AGWs), are <span class="hlt">observable</span> in the OH <span class="hlt">airglow</span>. The objective of the NIRAC experiment is to make near global measurement of the OH <span class="hlt">airglow</span> and <span class="hlt">airglow</span> perturbations. These emissions also provide a bright source of illumination at night, allowing for nighttime detection of clouds and surface characteristics. The instrument, developed by the Aerospace Space Science Applications Laboratory, employs a space-compatible FPGA for camera control and data collection and a novel, custom optical system to eliminate <span class="hlt">image</span> smear due to orbital motion. NIRAC utilizes a high-performance, large format infrared focal plane array, transitioning technology used in the existing Aerospace Corporation ground-based <span class="hlt">airglow</span> <span class="hlt">imager</span> to a space based platform. The high-sensitivity, four megapixel <span class="hlt">imager</span> has a native spatial resolution of 100 meters at ISS altitudes. The 23° x 23° FOV sweeps out a 150 km swath of the OH <span class="hlt">airglow</span> layer as viewed from the ISS, and is sensitive to OH intensity perturbations down to 0.1%. The detector has a 1.7 micron cutoff that precludes the need for cold optics and reduces cooling requirements (to 180 K). Detector cooling is provided by a compact, lightweight cryocooler capable of reaching 120K, providing a great deal of margin.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JASTP.154...33C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JASTP.154...33C"><span>Case study of convective instability <span class="hlt">observed</span> in <span class="hlt">airglow</span> <span class="hlt">images</span> over the Northeast of Brazil</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Carvalho, A. J. A.; Paulino, I.; Medeiros, A. F.; Lima, L. M.; Buriti, R. A.; Paulino, A. R.; Wrasse, C. M.; Takahashi, H.</p> <p>2017-02-01</p> <p>An intense activity of ripples during the nighttime was <span class="hlt">observed</span> in <span class="hlt">airglow</span> <span class="hlt">images</span> over São João do Cariri (36.5° W, 7.4° S) on 10 October 2004 which lasted for two hours. Those ripples appeared simultaneously with the crossing of a mesospheric front and medium scale gravity waves. The ripples occurred ahead of the mesospheric front and their phase front were almost parallel to the phase of the mesospheric front and were almost perpendicular to the phase front of the gravity wave. Using wind measurements from a meteor radar located at São João do Cariri and simultaneous vertical temperature profiles from the TIMED/SABER satellite, on the night of the events and within the <span class="hlt">imager</span> field of view, the atmospheric background environment in the mesosphere and lower thermosphere (MLT) was investigated in order to understand the instability process that caused the appearance of the ripples. Dynamic and convective instabilities have been pointed out as responsible for creation of ripples in the MLT. The <span class="hlt">observed</span> ripples were advected by the neutral wind, they occurred into a region with negative lapse rate of the potential temperature and the Richardson number was negative as well. According to these characteristics, the ripple structures could be generated in the MLT region due to the predominance of convective instability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA31A2387B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA31A2387B"><span>Mid-latitude response to geomagnetic storms <span class="hlt">observed</span> in 630nm <span class="hlt">airglow</span> over continental 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>Bhatt, A.; Kendall, E. A.</p> <p>2016-12-01</p> <p>We present analysis of mid-latitude response <span class="hlt">observed</span> to geomagnetic storms using the MANGO network consisting of all-sky cameras <span class="hlt">imaging</span> 630nm emission over the continental United States. The response largely falls in two categories: Stable Auroral Red (SAR) arc and Large-scale traveling ionospheric disturbances (LSTIDs). However, outside of these phenomena, less often <span class="hlt">observed</span> response include anomalous <span class="hlt">airglow</span> brightening, bright swirls, and frozen in traveling structures. We will present an analysis of various events <span class="hlt">observed</span> over 3 years of MANGO network operation, which started with two <span class="hlt">imagers</span> in the western US with addition of new <span class="hlt">imagers</span> in the last year. We will also present unusual north and northeastward propagating waves often <span class="hlt">observed</span> in conjunction with diffuse aurora. Wherever possible, we will compare with <span class="hlt">observations</span> from Boston University <span class="hlt">imagers</span> located in Massachusetts and Texas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.4628Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.4628Y"><span>Peak height of OH <span class="hlt">airglow</span> derived from simultaneous <span class="hlt">observations</span> a Fabry-Perot interferometer and a meteor radar</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yu, Tao; Zuo, Xiaomin; Xia, Chunliang; Li, Mingyuan; Huang, Cong; Mao, Tian; Zhang, Xiaoxin; Zhao, Biqiang; Liu, Libo</p> <p>2017-04-01</p> <p>A new method for estimating daily averaged peak height of the OH <span class="hlt">airglow</span> layer from a ground-based meteor radar (MR) and a Fabry-Perot interferometer (FPI) is presented. The first results are derived from 4 year simultaneous measurements of winds by a MR and a FPI at two adjacent stations over center China and are compared with <span class="hlt">observations</span> from the Thermosphere Ionosphere Mesosphere Energetics and Dynamics/Sounding of the Atmosphere using Broadband Emission Radiometry (SABER) instrument. The OH <span class="hlt">airglow</span> peak heights, which are derived by using correlation analysis between winds of the FPI and MR, are found to generally peak at an altitude of 87 km and frequently varied between 80 km and 90 km day to day. In comparison with SABER OH 1.6 μm <span class="hlt">observations</span>, reasonable similarity of <span class="hlt">airglow</span> peak heights is found, and rapid day-to-day variations are also pronounced. Lomb-Scargle analysis is used to determine cycles of temporal variations of <span class="hlt">airglow</span> peak heights, and there are obvious periodic variations both in our <span class="hlt">airglow</span> peak heights and in the satellite <span class="hlt">observations</span>. In addition to the annual, semiannual, monthly, and three monthly variations, the shorter time variations, e.g., day-to-day and several days' variations, are also conspicuous. The day-to-day variations of <span class="hlt">airglow</span> height obviously could reduce <span class="hlt">observation</span> accuracy and lead to some deviations in FPI measurements. These FPI wind deviations arising from <span class="hlt">airglow</span> height variations are also estimated to be about 3-5 m/s from 2011 to 2015, with strong positive correlation with <span class="hlt">airglow</span> peak height variation. More attention should be paid to the wind deviations associated with <span class="hlt">airglow</span> height variation when using and interpreting winds measured by FPI.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSA21B2025K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSA21B2025K"><span>HF-induced <span class="hlt">airglow</span> structure as a proxy for ionospheric irregularity detection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kendall, E. A.</p> <p>2013-12-01</p> <p>The High Frequency Active Auroral Research Program (HAARP) heating facility allows scientists to test current theories of plasma physics to gain a better understanding of the underlying mechanisms at work in the lower ionosphere. One powerful technique for diagnosing radio frequency interactions in the ionosphere is to use ground-based optical instrumentation. High-frequency (HF), heater-induced artificial <span class="hlt">airglow</span> <span class="hlt">observations</span> can be used to diagnose electron energies and distributions in the heated region, illuminate natural and/or artificially induced ionospheric irregularities, determine ExB plasma drifts, and measure quenching rates by neutral species. Artificial <span class="hlt">airglow</span> is caused by HF-accelerated electrons colliding with various atmospheric constituents, which in turn emit a photon. The most common emissions are 630.0 nm O(1D), 557.7 nm O(1S), and 427.8 nm N2+(1NG). Because more photons will be emitted in regions of higher electron energization, it may be possible to use <span class="hlt">airglow</span> <span class="hlt">imaging</span> to map artificial field-aligned irregularities at a particular altitude range in the ionosphere. Since fairly wide field-of-view <span class="hlt">imagers</span> are typically deployed in <span class="hlt">airglow</span> campaigns, it is not well-known what meter-scale features exist in the artificial <span class="hlt">airglow</span> emissions. Rocket data show that heater-induced electron density variations, or irregularities, consist of bundles of ~10-m-wide magnetic field-aligned filaments with a mean depletion depth of 6% [Kelley et al., 1995]. These bundles themselves constitute small-scale structures with widths of 1.5 to 6 km. Telescopic <span class="hlt">imaging</span> provides high resolution spatial coverage of ionospheric irregularities and goes hand in hand with other <span class="hlt">observing</span> techniques such as GPS scintillation, radar, and ionosonde. Since <span class="hlt">airglow</span> <span class="hlt">observations</span> can presumably <span class="hlt">image</span> ionospheric irregularities (electron density variations), they can be used to determine the spatial scale variation, the fill factor, and the lifetime characteristics of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AnGeo..34..293P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AnGeo..34..293P"><span>Periodic waves in the lower thermosphere <span class="hlt">observed</span> by OI630 nm <span class="hlt">airglow</span> <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paulino, I.; Medeiros, A. F.; Vadas, S. L.; Wrasse, C. M.; Takahashi, H.; Buriti, R. A.; Leite, D.; Filgueira, S.; Bageston, J. V.; Sobral, J. H. A.; Gobbi, D.</p> <p>2016-02-01</p> <p>Periodic wave structures in the thermosphere have been <span class="hlt">observed</span> at São João do Cariri (geographic coordinates: 36.5° W, 7.4° S; geomagnetic coordinates based on IGRF model to 2015: 35.8° E, 0.48° N) from September 2000 to November 2010 using OI630.0 nm <span class="hlt">airglow</span> <span class="hlt">images</span>. During this period, which corresponds to almost one solar cycle, characteristics of 98 waves were studied. Similarities between the characteristics of these events and <span class="hlt">observations</span> at other places around the world were noted, primarily the spectral parameters. The <span class="hlt">observed</span> periods were mostly found between 10 and 35 min; horizontal wavelengths ranged from 100 to 200 km, and phase speed from 30 to 180 m s-1. These parameters indicated that some of the waves, presented here, are slightly faster than those <span class="hlt">observed</span> previously at low and middle latitudes (Indonesia, Carib and Japan), indicating that the characteristics of these waves may change at different places. Most of <span class="hlt">observed</span> waves have appeared during magnetically quiet nights, and the occurrence of those waves followed the solar activity. Another important characteristic is the quasi-monochromatic periodicity that distinguish them from the single-front medium-scale traveling ionospheric disturbances (MSTIDs) that have been <span class="hlt">observed</span> previously over the Brazilian region. Moreover, most of the <span class="hlt">observed</span> waves did not present a phase front parallel to the northeast-southwest direction, which is predicted by the Perkins instability process. It strongly suggests that most of these waves must have had different generation mechanisms from the Perkins instability, which have been pointed out as being a very important mechanism for the generation of MSTIDs in the lower thermosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1912782S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1912782S"><span>High resolution <span class="hlt">observations</span> of small-scale gravity waves and turbulence features in the OH <span class="hlt">airglow</span> layer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sedlak, René; Hannawald, Patrick; Schmidt, Carsten; Wüst, Sabine; Bittner, Michael</p> <p>2017-04-01</p> <p>A new version of the Fast <span class="hlt">Airglow</span> <span class="hlt">Imager</span> (FAIM) for the detection of atmospheric waves in the OH <span class="hlt">airglow</span> layer has been set up at the German Remote Sensing Data Centre (DFD) of the German Aerospace Centre (DLR) at Oberpfaffenhofen (48.09 ° N, 11.28 ° E), Germany. The spatial resolution of the instrument is 17 m/pixel in zenith direction with a field of view (FOV) of 11.1 km x 9.0 km at the OH layer height of ca. 87 km. Since November 2015, the system has been in operation in two different setups (zenith angles 46 ° and 0 °) with a temporal resolution of 2.5 to 2.8 s. In a first case study we present <span class="hlt">observations</span> of two small wave-like features that might be attributed to gravity wave instabilities. In order to spectrally analyse harmonic structures even on small spatial scales down to 550 m horizontal wavelength, we made use of the Maximum Entropy Method (MEM) since this method exhibits an excellent wavelength resolution. MEM further allows analysing relatively short data series, which considerably helps to reduce problems such as stationarity of the underlying data series from a statistical point of view. We present an <span class="hlt">observation</span> of the subsequent decay of well-organized wave fronts into eddies, which we tentatively interpret in terms of an indication for the onset of turbulence. Another remarkable event which demonstrates the technical capabilities of the instrument was <span class="hlt">observed</span> during the night of 4th to 5th April 2016. It reveals the disintegration of a rather homogenous brightness variation into several filaments moving in different directions and with different speeds. It resembles the formation of a vortex with a horizontal axis of rotation likely related to a vertical wind shear. This case shows a notable similarity to what is expected from theoretical modelling of Kelvin-Helmholtz instabilities (KHIs). The comparatively high spatial resolution of the presented new version of the FAIM <span class="hlt">airglow</span> <span class="hlt">imager</span> provides new insights into the structure of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AMT....10.3093F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AMT....10.3093F"><span>Optimizing hydroxyl <span class="hlt">airglow</span> retrievals from long-slit astronomical spectroscopic <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>Franzen, Christoph; Hibbins, Robert Edward; Espy, Patrick Joseph; Djupvik, Anlaug Amanda</p> <p>2017-08-01</p> <p>Astronomical spectroscopic <span class="hlt">observations</span> from ground-based telescopes contain background emission lines from the terrestrial atmosphere's <span class="hlt">airglow</span>. In the near infrared, this background is composed mainly of emission from Meinel bands of hydroxyl (OH), which is produced in highly excited vibrational states by reduction of ozone near 90 km. This emission contains a wealth of information on the chemical and dynamical state of the Earth's atmosphere. However, <span class="hlt">observation</span> strategies and data reduction processes are usually optimized to minimize the influence of these features on the astronomical spectrum. Here we discuss a measurement technique to optimize the extraction of the OH <span class="hlt">airglow</span> signal itself from routine J-, H-, and K-band long-slit astronomical spectroscopic <span class="hlt">observations</span>. As an example, we use data recorded from a point-source <span class="hlt">observation</span> by the Nordic Optical Telescope's intermediate-resolution spectrograph, which has a spatial resolution of approximately 100 m at the <span class="hlt">airglow</span> layer. Emission spectra from the OH vibrational manifold from v' = 9 down to v' = 3, with signal-to-noise ratios up to 280, have been extracted from 10.8 s integrations. Rotational temperatures representative of the background atmospheric temperature near 90 km, the mesosphere and lower thermosphere region, can be fitted to the OH rotational lines with an accuracy of around 0.7 K. Using this measurement and analysis technique, we derive a rotational temperature distribution with v' that agrees with atmospheric model conditions and the preponderance of previous work. We discuss the derived rotational temperatures from the different vibrational bands and highlight the potential for both the archived and future <span class="hlt">observations</span>, which are at unprecedented spatial and temporal resolutions, to contribute toward the resolution of long-standing problems in atmospheric physics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMNH51A1922G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMNH51A1922G"><span>Methods for analyzing optical <span class="hlt">observations</span> of tsunami-induced signatures in <span class="hlt">airglow</span> emissions from ground-based and space-based platforms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grawe, M.; Makela, J. J.</p> <p>2016-12-01</p> <p><span class="hlt">Airglow</span> <span class="hlt">imaging</span> of the 630.0-nm redline emission has emerged as a useful tool for studying the properties of tsunami-ionospheric coupling in recent years, offering spatially continuous coverage of the sky with a single instrument. Past studies have shown that <span class="hlt">airglow</span> signatures induced by tsunamis are inherently anisotropic due to the <span class="hlt">observation</span> geometry and effects from the geomagnetic field. Here, we present details behind the techniques used to determine the parameters of the signature (orientation, wavelength, etc) with potential extensions to real or quasi-real time and a tool for interpreting the location and strength of the signatures in the field of view. We demonstrate application of the techniques to ground-based optical measurements of several tsunami-induced signatures taking place over the past five years from an <span class="hlt">imaging</span> system in Hawaii. Additionally, these methods are extended for use on space-based <span class="hlt">observation</span> platforms, offering advantages over ground-based installations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA33A2575F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA33A2575F"><span>Small-Scale Dynamical Structures Using OH <span class="hlt">Airglow</span> From Astronomical <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>Franzen, C.; Espy, P. J.; Hibbins, R. E.; Djupvik, A. A.</p> <p>2017-12-01</p> <p>Remote sensing of perturbations in the hydroxyl (OH) Meinel <span class="hlt">airglow</span> has often been used to <span class="hlt">observe</span> gravity, tidal and planetary waves travelling through the 80-90 km region. While large scale (>1 km) gravity waves and the winds caused by their breaking are widely documented, information on the highest frequency waves and instabilities occurring during the breaking process is often limited by the temporal and spatial resolution of the available <span class="hlt">observations</span>. In an effort to better quantify the full range of wave scales present near the mesopause, we present a series of <span class="hlt">observations</span> of the OH Meinel (9,7) transition that were executed with the Nordic Optical Telescope on La Palma (18°W, 29°N). These measurements have a 24 s repetition rate and horizontal spatial resolutions at 87 km as small as 10 cm, allowing us to quantify the transition in the mesospheric wave domains as the gravity waves break. Temporal scales from hours to minutes, as well as sub-100 m coherent structures in the OH <span class="hlt">airglow</span> have been <span class="hlt">observed</span> and will be presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.1363T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.1363T"><span>Plans of lightning and <span class="hlt">airglow</span> measurements with LAC/Akatsuki</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takahashi, Yukihiro; Hoshino, Naoya; Sato, Mitsuteru; Yair, Yoav; Galand, Marina; Fukuhara, Tetsuya</p> <p></p> <p>Though there are extensive researches on the existence of lightning discharge in Venus over few decades, this issue is still under controversial. Recently it is reported that the magnetometer on board Venus Express detected whistler mode waves whose source could be lightning discharge occurring well below the spacecraft. However, it is too early to determine the origin of these waves. On the other hand, night <span class="hlt">airglow</span> is expected to provide essential information on the atmospheric circulation in the upper atmosphere of Venus. But the number of consecutive <span class="hlt">images</span> of <span class="hlt">airglow</span> obtained by spacecraft is limited and even the variations of most enhanced location is still unknown. In order to identify the discharge phenomena in the atmosphere of Venus separating from noises and to know the daily variation of <span class="hlt">airglow</span> distribution in night-side disk, we plan to <span class="hlt">observe</span> the lightning and <span class="hlt">airglow</span> optical emissions with high-speed and high-sensitivity optical detector with narrow-band filters on board Akatsuki. We are ready to launch the flight model of lightning and <span class="hlt">airglow</span> detector, LAC (Lightning and <span class="hlt">Airglow</span> Camera). Main difference from other previous equipments which have provided evidences of lightning existence in Venus is the high-speed sampling rate at 32 us interval for each pixel, enabling us to distinguish the optical lightning flash from other pulsing noises. In this presentation the <span class="hlt">observation</span> strategies, including ground-based support with optical telescopes, are shown and discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850067068&hterms=ionospheric+modification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dionospheric%2Bmodification','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850067068&hterms=ionospheric+modification&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dionospheric%2Bmodification"><span>Simulations and <span class="hlt">observations</span> of plasma depletion, ion composition, and <span class="hlt">airglow</span> emissions in two auroral ionospheric depletion experiments</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yau, A. W.; Whalen, B. A.; Harris, F. R.; Gattinger, R. L.; Pongratz, M. B.</p> <p>1985-01-01</p> <p><span class="hlt">Observations</span> of plasma depletion, ion composition modification, and <span class="hlt">airglow</span> emissions in the Waterhole experiments are presented. The detailed ion chemistry and <span class="hlt">airglow</span> emission processes related to the ionospheric hole formation in the experiment are examined, and <span class="hlt">observations</span> are compared with computer simulation results. The latter indicate that the overall depletion rates in different parts of the depletion region are governed by different parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AdSpR..54..554B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AdSpR..54..554B"><span>Limb Viewing Hyper Spectral <span class="hlt">Imager</span> (LiVHySI) for <span class="hlt">airglow</span> measurements onboard YOUTHSAT-1</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bisht, R. S.; Hait, A. K.; Babu, P. N.; Sarkar, S. S.; Benerji, A.; Biswas, A.; Saji, A. K.; Samudraiah, D. R. M.; Kirankumar, A. S.; Pant, T. K.; Parimalarangan, T.</p> <p>2014-08-01</p> <p>The Limb Viewing Hyper Spectral <span class="hlt">Imager</span> (LiVHySI) is one of the Indian payloads onboard YOUTHSAT (inclination 98.73°, apogee 817 km) launched in April, 2011. The Hyper-spectral <span class="hlt">imager</span> has been operated in Earth’s limb viewing mode to measure <span class="hlt">airglow</span> emissions in the spectral range 550-900 nm, from terrestrial upper atmosphere (i.e. 80 km altitude and above) with a line-of-sight range of about 3200 km. The altitude coverage is about 500 km with command selectable lowest altitude. This <span class="hlt">imaging</span> spectrometer employs a Linearly Variable Filter (LVF) to generate the spectrum and an Active Pixel Sensor (APS) area array of 256 × 512 pixels, placed in close proximity of the LVF as detector. The spectral sampling is done at 1.06 nm interval. The optics used is an eight element f/2 telecentric lens system with 80 mm effective focal length. The detector is aligned with respect to the LVF such that its 512 pixel dimension covers the spectral range. The radiometric sensitivity of the <span class="hlt">imager</span> is about 20 Rayleigh at noise floor through the signal integration for 10 s at wavelength 630 nm. The <span class="hlt">imager</span> is being operated during the eclipsed portion of satellite orbits. The integration in the time/spatial domain could be chosen depending upon the season, solar and geomagnetic activity and/or specific target area. This paper primarily aims at describing LiVHySI, its in-orbit operations, quality, potential of the data and its first <span class="hlt">observations</span>. The <span class="hlt">images</span> reveal the thermospheric <span class="hlt">airglow</span> at 630 nm to be the most prominent. These first LiVHySI <span class="hlt">observations</span> carried out on the night of 21st April, 2011 are presented here, while the variability exhibited by the thermospheric nightglow at O(1D) 630 nm has been described in detail.</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_1");'>1</a></li> <li class="active"><span>2</span></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_2 --> <div id="page_3" 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_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</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="41"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..253G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..253G"><span>Mesospheric front <span class="hlt">observations</span> by the OH <span class="hlt">airglow</span> <span class="hlt">imager</span> carried out at Ferraz Station on King George Island, Antarctic Peninsula, in 2011</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Giongo, Gabriel Augusto; Valentin Bageston, José; Prado Batista, Paulo; Wrasse, Cristiano Max; Dornelles Bittencourt, Gabriela; Paulino, Igo; Paes Leme, Neusa Maria; Fritts, David C.; Janches, Diego; Hocking, Wayne; Schuch, Nelson Jorge</p> <p>2018-02-01</p> <p>The main goals of this work are to characterize and investigate the potential wave sources of four mesospheric fronts identified in the hydroxyl near-infrared (OH-NIR) <span class="hlt">airglow</span> <span class="hlt">images</span>, obtained with an all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span> installed at Comandante Ferraz Antarctic Station (EACF, as per its Portuguese acronym) located on King George Island in the Antarctic Peninsula. We identified and analyzed four mesospheric fronts in 2011 over King George Island. In addition, we investigate the atmospheric background environment between 80 and 100 km altitude and discuss the ducts and propagation conditions for these waves. For that, we used wind data obtained from a meteor radar operated at EACF and temperature data obtained from the TIMED/SABER satellite. The vertical wavenumber squared, m2, was calculated for each of the four waves. Even though no clearly defined duct (indicated by positive values of m2 sandwiched between layers above and below with m2 < 0) was found in any of the events, favorable propagation conditions for horizontal propagation of the fronts were found in three cases. In the fourth case, the wave front did not find any duct support and it appeared to dissipate near the zenith, transferring energy and momentum to the medium and, consequently, accelerating the wind in the wave propagation direction (near to south) above the OH peak (88-92 km). The likely wave sources for these four cases were investigated by using meteorological satellite <span class="hlt">images</span> and in two cases we could find that strong instabilities were potential sources, i.e., a cyclonic activity and a large convective cloud cell. In the other two cases it was not possible to associate troposphere sources as potential candidates for the generation of such wave fronts <span class="hlt">observed</span> in the mesosphere and secondary wave sources were attributed to these cases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSA42A..04H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSA42A..04H"><span>Optical <span class="hlt">imaging</span> of <span class="hlt">airglow</span> structure in equatorial plasma bubbles at radio scintillation scales</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Holmes, J. M.; Pedersen, T.; Parris, R. T.; Stephens, B.; Caton, R. G.; Dao, E. V.; Kratochvil, S.; Morton, Y.; Xu, D.; Jiao, Y.; Taylor, S.; Carrano, C. S.</p> <p>2015-12-01</p> <p>Imagery of optical emissions from F-region plasma is one of the few means available todetermine plasma density structure in two dimensions. However, the smallest spatial scalesobservable with this technique are typically limited not by magnification of the lens or resolutionof the detector but rather by the optical throughput of the system, which drives the integrationtime, which in turn causes smearing of the features that are typically moving at speeds of 100m/s or more. In this paper we present high spatio-temporal imagery of equatorial plasma bubbles(EPBs) from an <span class="hlt">imaging</span> system called the Large Aperture Ionospheric Structure <span class="hlt">Imager</span>(LAISI), which was specifically designed to capture short-integration, high-resolution <span class="hlt">images</span> ofF-region recombination <span class="hlt">airglow</span> at λ557.7 nm. The <span class="hlt">imager</span> features 8-inch diameter entranceoptics comprised of a unique F/0.87 lens, combined with a monolithic 8-inch diameterinterference filter and a 2x2-inch CCD detector. The LAISI field of view is approximately 30degrees. Filtered all-sky <span class="hlt">images</span> at common <span class="hlt">airglow</span> wavelengths are combined with magneticfield-aligned LAISI <span class="hlt">images</span>, GNSS scintillation, and VHF scintillation data obtained atAscension Island (7.98S, 14.41W geographic). A custom-built, multi-constellation GNSS datacollection system was employed that sampled GPS L1, L2C, L5, GLONASS L1 and L2, BeidouB1, and Galileo E1 and E5a signals. Sophisticated processing software was able to maintainlock of all signals during strong scintillation, providing unprecedented spatial <span class="hlt">observability</span> ofL band scintillation. The smallest-resolvable scale sizes above the noise floor in the EPBs, as viewed byLAISI, are illustrated for integration times of 1, 5 and 10 seconds, with concurrentzonal irregularity drift speeds from both spaced-receiver VHF measurements and single-stationGNSS measurements of S4 and σφ. These <span class="hlt">observable</span> optical scale sizes are placed in thecontext of those that give rise to radio scintillation in VHF and L band signals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA42A..07R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA42A..07R"><span>Nighttime medium scale traveling ionospheric disturbances in southern hemisphere using FORMOSAT-2/ISUAL 630.0 nm <span class="hlt">airglow</span> <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rajesh, P. K.; Lin, C. C. H.; Liu, T. J. Y.; Chen, A. B. C.; Hsu, R. R.; Chen, C. H.; Huba, J. D.</p> <p>2017-12-01</p> <p>In this work characteristics of nighttime medium-scale travelling ionospheric disturbances (MSTID) are investigated using 630.0 nm limb <span class="hlt">images</span> by <span class="hlt">Imager</span> of Sprites and Upper Atmospheric Lightnings (ISUAL), onboard FORMOSAT-2 satellite. The limb integrated measurements, when projected to a horizontal plane, reveal bands of intensity perturbation with distinct southwest to northeast orientation in the southern hemisphere. <span class="hlt">Airglow</span> simulations are carried out by artificially introducing MSTID fluctuations in model electron density to confirm if such azimuthally oriented features could be identified in the ISUAL viewing geometry. Further statistical analysis shows more MSTID occurrence in solstices with peak in June-July months. The wavelengths of the <span class="hlt">observed</span> perturbations were in the range 150-300 km. The wave fronts were oriented about 30°-50° from the east-west plane, indicating that coupled Perkins and Es-layer instability might be important in the MSTID generation. The results demonstrate that space based <span class="hlt">airglow</span> <span class="hlt">imaging</span> is an effective method for global investigation of MSTID events that are appropriately aligned with the viewing geometry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..12212430O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..12212430O"><span>First Study on the Occurrence Frequency of Equatorial Plasma Bubbles over West Africa Using an All-Sky <span class="hlt">Airglow</span> <span class="hlt">Imager</span> and GNSS Receivers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Okoh, Daniel; Rabiu, Babatunde; Shiokawa, Kazuo; Otsuka, Yuichi; Segun, Bolaji; Falayi, Elijah; Onwuneme, Sylvester; Kaka, Rafiat</p> <p>2017-12-01</p> <p>This is the first paper that reports the occurrence frequency of equatorial plasma bubbles and their dependences of local time, season, and geomagnetic activity based on <span class="hlt">airglow</span> <span class="hlt">imaging</span> <span class="hlt">observations</span> at West Africa. The all-sky <span class="hlt">imager</span>, situated in Abuja (Geographic: 8.99°N, 7.38°E; Geomagnetic: 1.60°S), has a 180° fisheye view covering almost the entire airspace of Nigeria. Plasma bubbles are <span class="hlt">observed</span> for 70 nights of the 147 clear-sky nights from 9 June 2015 to 31 January 2017. Differences between nighttime and daytime ROTIs were also computed as a proxy of plasma bubbles using Global Navigation Satellite Systems (GNSS) receivers within the coverage of the all-sky <span class="hlt">imager</span>. Most plasma bubble occurrences are found during equinoxes and least occurrences during solstices. The occurrence rate of plasma bubbles was highest around local midnight and lower for hours farther away. Most of the postmidnight plasma bubbles were <span class="hlt">observed</span> around the months of December to March, a period that coincides with the harmattan period in Nigeria. The on/off status of plasma bubble in <span class="hlt">airglow</span> and GNSS <span class="hlt">observations</span> were in agreement for 67.2% of the total 768 h, while we suggest several reasons responsible for the remaining 32.8% when the <span class="hlt">airglow</span> and GNSS bubble status are inconsistent. A majority of the plasma bubbles were <span class="hlt">observed</span> under relatively quiet geomagnetic conditions (Dst ≥ -40 and Kp ≤ 3), but there was no significant pattern <span class="hlt">observed</span> in the occurrence rate of plasma bubbles as a function of geomagnetic activity. We suggest that geomagnetic activities could have either suppressed or promoted the occurrence of plasma bubbles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010JGRA..115.9315A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JGRA..115.9315A"><span>Midnight latitude-altitude distribution of 630 nm <span class="hlt">airglow</span> in the Asian sector measured with FORMOSAT-2/ISUAL</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Adachi, Toru; Yamaoka, Masashi; Yamamoto, Mamoru; Otsuka, Yuichi; Liu, Huixin; Hsiao, Chun-Chieh; Chen, Alfred B.; Hsu, Rue-Ron</p> <p>2010-09-01</p> <p>The <span class="hlt">Imager</span> for Sprites and Upper Atmospheric Lightning (ISUAL) payload on board the FORMOSAT-2 satellite carried out the first limb <span class="hlt">imaging</span> <span class="hlt">observation</span> of 630 nm <span class="hlt">airglow</span> for the purpose of studying physical processes in the F region ionosphere. For a total of 14 nights in 2006-2008, ISUAL scanned the midnight latitude-altitude distribution of 630 nm <span class="hlt">airglow</span> in the Asian sector. On two nights of relatively active conditions (ΣKp = 26, 30+) we found several bright <span class="hlt">airglow</span> regions, which were highly variable each night in terms of luminosity and location. In relatively quiet conditions (ΣKp = 4-20) near May/June we found two bright regions which were stably located in the midlatitude region of 40°S-10°S (50°S-20°S magnetic latitude (MLAT)) and in the equatorial region of 0°-10°N (10°S-0° MLAT). On one of the quiet nights, FORMOSAT-3/COSMIC and CHAMP simultaneously measured the plasma density in the same region where ISUAL <span class="hlt">observed</span> <span class="hlt">airglow</span>. The plasma density data generally show good agreement, suggesting that plasma enhancements were the primary source of these two bright <span class="hlt">airglow</span> regions. From detailed comparison with past studies we explain that the <span class="hlt">airglow</span> in the equatorial region was due to the midnight brightness wave produced in association with the midnight temperature maximum, while that in the midlatitude region was due to the typical plasma distribution usually formed in the midnight sector. The fact that the equatorial <span class="hlt">airglow</span> was much brighter than the midlatitude <span class="hlt">airglow</span> and was <span class="hlt">observed</span> on most nights during the campaign period strongly suggests the importance of further studies on the MTM/MBW phenomenology, which is not well reproduced in the current general circulation model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004cosp...35..231A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004cosp...35..231A"><span>Zonal drift velocities of the ionospheric plasma bubbles over brazilian region using oi630nm <span class="hlt">airglow</span> digital <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arruda, D. C. S.; Sobral, J. H. A.; Abdu, M. A.; Castilho, V. M.; Takahashi, H.</p> <p></p> <p>The zonal drift velocities of the ionospheric plasma bubbles over the Brazilian region are analyzed in this study that is based on OI630nm <span class="hlt">airglow</span> digital <span class="hlt">images</span>. These digital <span class="hlt">images</span> were obtained by an all-sky <span class="hlt">imager</span> system between October 1998 and August 2000, at Cachoeira Paulista (22.5°S, 45°W), a low latitude region. In this period, 138 nights of OI 630 nm <span class="hlt">airglow</span> experiments were carried out of which 30 nights detected the ionospheric plasma bubbles. These 30 nights correspond to magnetically quiet days (ΣK_P<24+) and were grouped according approximately to their season. KEY WORDS: <span class="hlt">Imager</span> System, Ionospheric Plasma Bubbles, Zonal drift velocities, OI630nm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA13B..03S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA13B..03S"><span>Optical <span class="hlt">Imaging</span> <span class="hlt">Observation</span> of the Geospace from the International Space Station by ISS-IMAP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saito, A.; Sakanoi, T.; Yoshikawa, I.; Yamazaki, A.; Hozumi, Y.; Perwitasari, S.; Otsuka, Y.; Yamamoto, M.</p> <p>2017-12-01</p> <p>Optical <span class="hlt">imaging</span> <span class="hlt">observation</span> of the mesosphere, thermosphere, ionosphere, and plasmasphere was carried out from the International Space Station (ISS) with ISS-IMAP (Ionosphere, Mesosphere, upper Atmosphere, and Plasmasphere mapping) mission instruments. ISS-IMAP instruments was installed on the Exposed Facility of Japanese Experiment Module of the ISS in August, 2012, and removed in August, 2015. They are two <span class="hlt">imagers</span>, Visible-light and Infrared Spectrum <span class="hlt">Imager</span> (VISI) and Extreme UltraViolet <span class="hlt">Imager</span> (EUVI). VISI made <span class="hlt">imaging</span> <span class="hlt">observations</span> of the <span class="hlt">airglow</span> and aurora in the nadir direction. It had two slits perpendicular to the trajectory of ISS, and the movement of ISS made the two-dimensional <span class="hlt">observation</span> whose field-of-view width is 600km at 100km altitude. It covered the wave length range from 500nm to 900nm. The <span class="hlt">airglow</span> of 730nm (OH, Alt. 85km), 762nm (O2, Alt. 95km), and 630nm (O, Alt. 250km) were mainly <span class="hlt">observed</span> besides the other <span class="hlt">airglow</span>, such as 589nm (Na) and 557 (O). EUVI made <span class="hlt">imaging</span> <span class="hlt">observation</span> of the resonant scattering from ions. It had two telescopes, and <span class="hlt">observed</span> the resonant scattering of He+ in 30.4nm, and O+ in 83.4nm in the limb direction. VISI captured the <span class="hlt">airglow</span> structures whose wavelength from 80km to 500km. The concentric wave structures were frequently <span class="hlt">observed</span> in the mesosphere and lower thermosphere region. They are strong evidence of the vertical coupling between the lower atmosphere and the upper atmosphere by vertical propagation of the atmospheric gravity waves. The other <span class="hlt">airglow</span> structures, such as mesospheric bores, were also detected by ISS-IMAP/VISI. The meso-scale structures in the ionosphere, such as plasma bubbles, and traveling ionospheric disturbances were also <span class="hlt">observed</span>. EUVI revealed the longitudinal structures of He+ in the top side of the ionosphere. It was attributed to the neutral wind in the thermosphere. In the presentation, the outline and results of the ISS-IMAP's VISI and EUVI <span class="hlt">observations</span> will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMNH21C1828O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMNH21C1828O"><span>Internal Gravity Wave Induced by the Queen Charlotte Event (27 October 2012, Mw 7.8): <span class="hlt">Airglow</span> <span class="hlt">Observation</span> and Modeling.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Occhipinti, G.; Bablet, A.; Makela, J. J.</p> <p>2015-12-01</p> <p>The detection of the tsunami related internal gravity waves (IGWtsuna) by <span class="hlt">airglow</span> camera has been recently validated by <span class="hlt">observation</span> (Makela et al., 2011) and modeling (Occhipinti et al., 2011) in the case of the Tohoku event (11 March 2011, Mw 9.0). The <span class="hlt">airglow</span> is measuring the photon emission at 630 nm, indirectly linked to the plasma density of O2+ (Link & Cogger, 1988) and it is commonly used to detect transient event in the ionosphere (Kelley et al., 2002, Makela et al., 2009, Miller et al., 2009). The modeling of the IGWtsuna clearly reproduced the pattern of the <span class="hlt">airglow</span> measurement <span class="hlt">observed</span> over Hawaii and the comparison between the <span class="hlt">observation</span> and the modeling allows to recognize the wave form and allow to explain the IGWtsuna arriving before the tsunami wavefront at the sea level (Occhipinti et al., 2011). Approaching the Hawaiian archipelagos the tsunami propagation is slowed down (reduction of the sea depth), instead, the IGWtsuna, propagating in the atmosphere/ionosphere, conserves its speed. In this work, we present the modeling of the new <span class="hlt">airglow</span> <span class="hlt">observation</span> following the Queen Charlotte event (27 October 2012, Mw 7.8) that has been recently detected, proving that the technique can be generalized for smaller events. Additionally, the effect of the wind on the IGWtsuna, already evocated in the past, is included in the modeling to better reproduce the <span class="hlt">airglow</span> <span class="hlt">observations</span>. All ref. here @ www.ipgp.fr/~ninto</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AMT.....9.5955S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AMT.....9.5955S"><span>High-resolution <span class="hlt">observations</span> of small-scale gravity waves and turbulence features in the OH <span class="hlt">airglow</span> layer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sedlak, René; Hannawald, Patrick; Schmidt, Carsten; Wüst, Sabine; Bittner, Michael</p> <p>2016-12-01</p> <p>A new version of the Fast <span class="hlt">Airglow</span> <span class="hlt">Imager</span> (FAIM) for the detection of atmospheric waves in the OH <span class="hlt">airglow</span> layer has been set up at the German Remote Sensing Data Center (DFD) of the German Aerospace Center (DLR) at Oberpfaffenhofen (48.09° N, 11.28° E), Germany. The spatial resolution of the instrument is 17 m pixel-1 in zenith direction with a field of view (FOV) of 11.1 km × 9.0 km at the OH layer height of ca. 87 km. Since November 2015, the system has been in operation in two different setups (zenith angles 46 and 0°) with a temporal resolution of 2.5 to 2.8 s. In a first case study we present <span class="hlt">observations</span> of two small wave-like features that might be attributed to gravity wave instabilities. In order to spectrally analyse harmonic structures even on small spatial scales down to 550 m horizontal wavelength, we made use of the maximum entropy method (MEM) since this method exhibits an excellent wavelength resolution. MEM further allows analysing relatively short data series, which considerably helps to reduce problems such as stationarity of the underlying data series from a statistical point of view. We present an <span class="hlt">observation</span> of the subsequent decay of well-organized wave fronts into eddies, which we tentatively interpret in terms of an indication for the onset of turbulence. Another remarkable event which demonstrates the technical capabilities of the instrument was <span class="hlt">observed</span> during the night of 4-5 April 2016. It reveals the disintegration of a rather homogenous brightness variation into several filaments moving in different directions and with different speeds. It resembles the formation of a vortex with a horizontal axis of rotation likely related to a vertical wind shear. This case shows a notable similarity to what is expected from theoretical modelling of Kelvin-Helmholtz instabilities (KHIs). The comparatively high spatial resolution of the presented new version of the FAIM provides new insights into the structure of atmospheric wave instability and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910058277&hterms=optics+interference&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doptics%2Binterference','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910058277&hterms=optics+interference&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doptics%2Binterference"><span>Thin film interference optics for <span class="hlt">imaging</span> the O II 834-A <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Seely, John F.; Hunter, William R.</p> <p>1991-01-01</p> <p>Normal incidence thin film interference mirrors and filters have been designed to <span class="hlt">image</span> the O II 834-A <span class="hlt">airglow</span>. It is shown that MgF2 is a useful spacer material for this wavelength region. The mirrors consist of thin layers of MgF2 in combination with other materials that are chosen to reflect efficiently in a narrow band centered at 834 A. Peak reflectance of 60 percent can be obtained with a passband 200 A wide. Al/MgF2/Si and Al/MgF2/SiC interference coatings have been designed to reflect 834 A and to absorb the intense H I 1216 A <span class="hlt">airglow</span>. An In/MgF2/In interference filter is designed to transmit 834 A and attenuate 1216 A radiation. Interference photocathode coatings for rejecting 1216 A radiation are also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSA23B..06H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSA23B..06H"><span>Heater-induced ionization inferred from spectrometric <span class="hlt">airglow</span> measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hysell, D. L.; Miceli, R. J.; Varney, R. H.; Schlatter, N.; Huba, J. D.</p> <p>2013-12-01</p> <p>Spectrographic <span class="hlt">airglow</span> measurements were made during an ionospheric modification experiment at HAARP on March 12, 2013. Artificial <span class="hlt">airglow</span> enhancements at 427.8, 557.7, 630.0, 777.4, and 844.6 nm were <span class="hlt">observed</span>. On the basis of these emissions and using a methodology based on the method of Backus and Gilbert [1968, 1970], we estimate the suprathermal electron population and the subsequent equilibrium electron density profile, including contributions from electron impact ionization. We find that the <span class="hlt">airglow</span> is consistent with significant induced ionization in view of the spatial intermittency of the <span class="hlt">airglow</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.1411S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.1411S"><span>The Martian <span class="hlt">airglow</span>: <span class="hlt">observations</span> by Mars Express and kinetic modelling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Simon, Cyril; Leblanc, François; Gronoff, Guillaume; Witasse, Olivier; Lilensten, Jean; Barthelemy, Mathieu; Bertaux, Jean-Loup</p> <p></p> <p>The photoemissions on Mars are the result of physical chemistry reactions in the upper atmo-sphere that depend on the planet's plasma environment. They arise on the dayside from UV photo-excitation (Barth et al., 1971) and on the nightside from chemical reactions and electron precipitation above regions of strong crustal magnetism (Bertaux et al., 2005). The physics of <span class="hlt">airglow</span> generation at Mars is discussed both in terms of <span class="hlt">observations</span> (satellites) and models (especially transport codes). A review of <span class="hlt">observations</span> made by SPICAM, the UV spectrometer onboard Mars Express, is first presented. The Cameron bands of CO(a - X), the CO+ (A - X) 2 doublet at 289.0 nm and the trans-auroral line of OI (297.2 nm) are mainly seen on the dayside. On the nightside both Cameron emissions and NO(C - X and A - X) emissions are present. In a second step, an updated <span class="hlt">airglow</span> model has been developed and compared to the latest SPICAM data. Several interesting implications are highlighted regarding neutral atmosphere variations for the dayglow (Simon et al., 2009) and electron precipitation mechanisms at the origin of the auroral intensities measured by SPICAM in conjunction with the particle detector ASPERA and the radar MARSIS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19780004642','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19780004642"><span>A rocket-borne <span class="hlt">airglow</span> photometer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Paarmann, L. D.; Smith, L. G.</p> <p>1977-01-01</p> <p>The design of a rocket-borne photometer to measure the <span class="hlt">airglow</span> emission of ionized molecular nitrogen in the 391.4 nm band is presented. This <span class="hlt">airglow</span> is a well known and often <span class="hlt">observed</span> phenomenon of auroras, where the principal source of ionization is energetic electrons. It is believed that at some midlatitude locations energetic electrons are also a source of nighttime ionization in the E region of the ionosphere. If this is so, then significant levels of 391.4 nm <span class="hlt">airglow</span> should be present. The intensity of this <span class="hlt">airglow</span> will be measured in a rocket payload which also contains instrumentation to measured in a rocket payload which also contains instrumentation to measure energetic electron differential flux and the ambient electron density. An intercomparison of the 3 experiments in a nightime launch will allow a test of the importance of energetic electrons as a nighttime source of ionization in the upper E region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JGRA..119.2038H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JGRA..119.2038H"><span>Heater-induced ionization inferred from spectrometric <span class="hlt">airglow</span> measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hysell, D. L.; Miceli, R. J.; Kendall, E. A.; Schlatter, N. M.; Varney, R. H.; Watkins, B. J.; Pedersen, T. R.; Bernhardt, P. A.; Huba, J. D.</p> <p>2014-03-01</p> <p>Spectrographic <span class="hlt">airglow</span> measurements were made during an ionospheric modification experiment at High Frequency Active Auroral Research Program on 12 March 2013. Artificial <span class="hlt">airglow</span> enhancements at 427.8, 557.7, 630.0, 777.4, and 844.6 nm were <span class="hlt">observed</span>. On the basis of these emissions and using a methodology based on the method of Backus and Gilbert (1968, 1970), we estimate the suprathermal electron population and the subsequent equilibrium electron density profile, including contributions from electron impact ionization. We find that the <span class="hlt">airglow</span> is consistent with heater-induced ionization in view of the spatial intermittency of the <span class="hlt">airglow</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA12A..03S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA12A..03S"><span>Recent <span class="hlt">observations</span> of traveling ionospheric disturbances and plasma bubbles using Optical Mesosphere Thermosphere <span class="hlt">Imagers</span> in Asian and African sectors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shiokawa, K.; Otsuka, Y.; Tsuchiya, S.; Moral, A. C.; Okoh, D.</p> <p>2017-12-01</p> <p>We review recent <span class="hlt">observational</span> results of medium-scale traveling ionospheric disturbances (MSTIDs) and equatorial plasma bubbles obtained by using <span class="hlt">airglow</span> <span class="hlt">imagers</span> and Fabry-Perot interferometers of the Optical Mesosphere Thermosphere <span class="hlt">Imagers</span> (OMTIs) at Asian and African sectors. The OMTIs contains 20 <span class="hlt">airglow</span> <span class="hlt">imagers</span> and 5 Fabry-Perot interferometers (FPIs) at Canada, USA (Alaska), Russia, Finland, Norway, Iceland, Japan, Thailand, Indonesia, Australia, and Nigeria (http://stdb2.isee.nagoya-u.ac.jp/omti/). The 3-dimentional Fast Fourier Transformation of <span class="hlt">airglow</span> <span class="hlt">images</span> makes it possible to analyze 16-year <span class="hlt">airglow</span> <span class="hlt">images</span> obtained at Shigaraki (34.8N) and Rikubetsu (43.5N), Japan, to obtain phase velocity spectra of gravity waves and MSTIDs. The MSTIDs spectra show clear southwestward preference of propagation and minor northeastward propagation over Japan. We also found clear negative correlation between MSTID power and solar F10.7 flux, indicating that MSTIDs becomes more active during solar quiet time. This fact suggest the control of ionospheric Perkins and E-F coupling instabilities by solar activities. Three TIDs in <span class="hlt">airglow</span> <span class="hlt">images</span> over Indonesia, including midnight brightness waves (MBWs), were compared with CHAMP-satellite overpass to investigate neutral density variations in the thermosphere associated with the TIDs. We found clear correspondence in variations between the <span class="hlt">airglow</span> intensities and neutral densities, suggesting that the <span class="hlt">observed</span> TIDs over the equatorial region is caused by gravity waves. We also compare average thermospheric temperatures measured by the four FPIs for 3-4 years with the MSIS90E and GAIA models. The comparison shows that GAIA generally shows better fitting than the MSIS90E, but at the equatorial stations, GAIA tends to fail to reproduce the FPI temperature, probably due to ambiguity of location of the midnight temperature maximum. We also made statistics of plasma bubble occurrence using <span class="hlt">airglow</span> <span class="hlt">imager</span> and GNSS receiver at Abuja (9</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.1919F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.1919F"><span>Very high resolution <span class="hlt">observations</span> of waves in the OH <span class="hlt">airglow</span> 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>Franzen, Christoph; Espy, Patrick J.; Hibbins, Robert E.; Djupvik, Amanda A.</p> <p>2017-04-01</p> <p>Vibrationally excited hydroxyl (OH) is produced in the mesosphere by the reaction of atomic hydrogen and ozone. This excited OH radiates a strong, near-infrared <span class="hlt">airglow</span> emission in a thin ( 8 km thick) layer near 87 km. In the past, remote sensing of perturbations in the OH Meinel <span class="hlt">airglow</span> has often been used to <span class="hlt">observe</span> gravity, tidal and planetary waves travelling through this region. However, information on the highest frequency gravity waves is often limited by the temporal and spatial resolution of the available <span class="hlt">observations</span>. In an effort to expand the wave scales present near the mesopause, we present a series of <span class="hlt">observations</span> of the OH Meinel (9,7) transition that were executed with the Nordic Optical Telescope on La Palma (18°W, 29°N). These measurements are taken with a 10 s integration time (24 s repetition rate), and the spatial resolution at 87 km is as small as 10 m, allowing us to quantify the transition between the gravity and acoustic wave domains in the mesosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9910E..1BR','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9910E..1BR"><span>Measurements of <span class="hlt">airglow</span> on Maunakea at Gemini Observatory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Roth, Katherine C.; Smith, Adam; Stephens, Andrew; Smirnova, Olesja</p> <p>2016-07-01</p> <p>Gemini Observatory on Maunakea has been collecting optical and infrared science data for almost 15 years. We have begun a program to analyze <span class="hlt">imaging</span> data from two of the original facility instruments, GMOS and NIRI, in order to measure sky brightness levels in multiple infrared and optical broad-band filters. The present work includes data from mid-2016 back through late-2008. We present measured background levels as a function of several operational quantities (e.g. moon phase, hours from twilight, season). We find that <span class="hlt">airglow</span> is a significant contributor to background levels in several filters. Gemini is primarily a queue scheduled telescope, with <span class="hlt">observations</span> being optimally executed in order to provide the most efficient use of telescope time. We find that while most parameters are well-understood, the atmospheric <span class="hlt">airglow</span> remains challenging to predict. This makes it difficult to schedule <span class="hlt">observations</span> which require dark skies in these filters, and we suggest improvements to ensure data quality.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JASTP.164..116L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JASTP.164..116L"><span>Semidiurnal tidal activity of the middle atmosphere at mid-latitudes derived from O2 atmospheric and OH(6-2) <span class="hlt">airglow</span> SATI <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>López-González, M. J.; Rodríguez, E.; García-Comas, M.; López-Puertas, M.; Olivares, I.; Ruiz-Bueno, J. A.; Shepherd, M. G.; Shepherd, G. G.; Sargoytchev, S.</p> <p>2017-11-01</p> <p>In this paper, we investigate the tidal activity in the mesosphere and lower thermosphere region at 370N using OH Meinel and O2 atmospheric <span class="hlt">airglow</span> <span class="hlt">observations</span> from 1998 to 2015. The <span class="hlt">observations</span> were taken with a Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span> (SATI) installed at Sierra Nevada Observatory (SNO) (37.060N, 3.380W) at 2900 m height. From these <span class="hlt">observations</span> a seasonal dependence of the amplitudes of the semidiurnal tide is inferred. The maximum tidal amplitude occurs in winter and the minimum in summer. The vertically averaged rotational temperatures and vertically integrated volume emission rate (rotational temperatures and intensities here in after), from the O2 atmospheric band measurements and the rotational temperature derived from OH Meinel band measurements reach the maximum amplitude about 1-4 h after midnight during almost all the year except in August-September where the maximum is found 2-4 h earlier. The amplitude of the tide in the OH intensity reaches the minimum near midnight in midwinter, then it is progressively delayed until 4:00 LT in August-September, and from there on it moves again forward towards midnight. The mean Krassovsky numbers for OH and O2 emissions are 5.9 ±1.8 and 5.6 ±1.0, respectively, with negative Krassovsky phases for almost all the year, indicating an upward energy transport. The mean vertical wavelengths for the vertical tidal propagation derived from OH and O2 emissions are 35 ±20 km and 33 ±18 km, respectively. The vertical wavelengths together with the phase shift in the temperature derived from both <span class="hlt">airglow</span> emissions indicate that these <span class="hlt">airglow</span> emission layers are separated by 7 ±3 km, on average.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.3748L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.3748L"><span>Characteristics of ripple structures revealed in OH <span class="hlt">airglow</span> <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Jing; Li, Tao; Dou, Xiankang; Fang, Xin; Cao, Bing; She, Chiao-Yao; Nakamura, Takuji; Manson, Alan; Meek, Chris; Thorsen, Denise</p> <p>2017-03-01</p> <p>Small-scale ripple structures <span class="hlt">observed</span> in OH <span class="hlt">airglow</span> <span class="hlt">images</span> are most likely induced by either dynamic instability due to large wind shear or convective instability due to superadiabatic lapse rate. Using the data set taken in the mesopause region with an OH all-sky <span class="hlt">imager</span> at Yucca Ridge Field Station, Colorado (40.7°N, 104.9°W), from September 2003 to December 2005, we study the characteristics and seasonal variations of ripple structures. By analyzing the simultaneous background wind and temperature <span class="hlt">observed</span> by the nearby sodium temperature/wind lidar at Fort Collins, Colorado (40.6°N, 105°W), and a nearby medium-frequency radar at Platteville, Colorado (40.2°N, 105.8°W), we are able to statistically study the possible relation between ripples and the background atmosphere conditions. Characteristics and seasonal variations of ripples are presented in detail in this study. The occurrence frequency of ripples exhibits clear seasonal variability, with peak in autumn. The occurrence of ripples shows a local time dependence, which is most likely associated with the solar tides. The lifetime and spatial scale of these ripples are typically 5-20 min and 5-10 km, respectively, and most of the ripples move preferentially either southward or northward. However, more than half of the <span class="hlt">observed</span> ripples do not advect with background flow; they have higher Richardson numbers than those ripples that advect with background flow. It is possible that they are not instability features but wave structures that are hard to be distinguished from the real instability features.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880042140&hterms=Abreu&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAbreu%252C%2Bc.','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880042140&hterms=Abreu&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAbreu%252C%2Bc."><span>The Visible <span class="hlt">Airglow</span> Experiment - A review</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hays, Paul B.; Abreu, Vincent J.; Solomon, Stanley C.; Yee, Jeng-Hwa</p> <p>1988-01-01</p> <p>Contributions of the Visible <span class="hlt">Airglow</span> Experiment (VAE) to the understanding of various <span class="hlt">airglow</span> and auroral processes are reviewed. The impact of instrumental design and operation on the <span class="hlt">observations</span> is discussed, and emphasis is placed on the relationship between <span class="hlt">observations</span> and inversion for optical measurements of light emissions from a diffuse medium. VAE data are used to explain the physical mechanisms responsible for the production and destruction of excited species including O(+)(1P), O(1D), O(1S), N(2D), and Mg(+)2P1/2.</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_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" 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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</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="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850029402&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850029402&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight"><span>Rocket <span class="hlt">observations</span> of the ultraviolet <span class="hlt">airglow</span> during morning twilight</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cebula, R. P.; Feldman, P. D.</p> <p>1984-01-01</p> <p>Rocket-borne (Astrobee) UV measurements were made of the terrestrial <span class="hlt">airglow</span> at morning twilight from 82 and 90 deg zenith angles at altitudes of 90 and 246 km in September 1979. Data were acquired on the NO gamma and delta bands, the 2470 A O II, 1356 A and the 1304 A O I lines, the Lyman-Berge-Hopfield N2 and the Herzberg 02 lines. The zodiacal contribution was substracted to obtain pure <span class="hlt">airglow</span> data. Spectral analyses supported a larger nighttime decrease of N(4S) than for NO, the latter being in diffusive equilibrium above 190 km altitude. The NO gamma band was directly related to the thermospheric N(4S) contribution, the latter having a density of 2-8 million/cu cm at 200 km. Finally, self-consistent photoionization and photoelectron impact ionization models were derived for the atomic and ionic oxygen emissions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720021779','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720021779"><span>The equatorial <span class="hlt">airglow</span> and the ionospheric geomagnetic anomaly</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chandra, S.; Reed, E. I.; Troy, B. E., Jr.; Blamont, J. E.</p> <p>1972-01-01</p> <p>OGO D <span class="hlt">observations</span> of OI (6300A) emissions reveal a global pattern in the equatorial <span class="hlt">airglow</span> undetected from the ground-based <span class="hlt">observations</span>. The post sunset emission rate of OI is generally asymmetrical with respect to the geomagnetic equator and shows no apparent correlation with the ultraviolet <span class="hlt">airglow</span> (OI 1304 and 1356A) and F region electron density measured simultaneously from the same spacecraft. Both the ultraviolet <span class="hlt">airglow</span> and the ion density measured in the altitude region of 450 km follow similar latitudinal variations and exhibit properties of the equatorial ionospheric anomaly. The asymmetry in OI emission can be attributed to the asymmetry in the height of the F 2 maximum inferred from the height of the maximum emission. From correlative studies of the <span class="hlt">airglow</span> and the ionospheric measurements, the mechanisms for the ultraviolet and the 6300A emission are discussed in terms of the processes involving radiative and dissociative recombinations. A relationship between molecular oxygen density and the integrated OI emission rate is derived and the feasibility of using this relationship for estimating O2 density is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930018294','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930018294"><span>An assessment of twilight <span class="hlt">airglow</span> inversion procedures using atmosphere explorer <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>Mcdade, I. C.; Sharp, W. E.</p> <p>1993-01-01</p> <p>The aim of this research project was to test and truth some recently developed methods for recovering thermospheric oxygen atom densities and thermospheric temperatures from ground-based <span class="hlt">observations</span> of the 7320 A O(+)((sup 2)D - (sup 2)P) twilight air glow emission. The research plan was to use twilight <span class="hlt">observations</span> made by the Visible <span class="hlt">Airglow</span> Experiment (VAE) on the Atmosphere Explorer 'E' satellite as proxy ground based twilight <span class="hlt">observations</span>. These <span class="hlt">observations</span> were to be processed using the twilight inversion procedures, and the recovered oxygen atom densities and thermospheric temperatures were then to be examined to see how they compared with the densities and temperatures that were measured by the Open Source Mass Spectrometer and the Neutral Atmosphere Temperature Experiment on the satellite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..705N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..705N"><span>Determination of gravity wave parameters in the <span class="hlt">airglow</span> combining photometer and <span class="hlt">imager</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nyassor, Prosper K.; Arlen Buriti, Ricardo; Paulino, Igo; Medeiros, Amauri F.; Takahashi, Hisao; Wrasse, Cristiano M.; Gobbi, Delano</p> <p>2018-05-01</p> <p>Mesospheric <span class="hlt">airglow</span> measurements of two or three layers were used to characterize both vertical and horizontal parameters of gravity waves. The data set was acquired coincidentally from a multi-channel filter (Multi-3) photometer and an all-sky <span class="hlt">imager</span> located at São João do Cariri (7.4° S, 36.5° W) in the equatorial region from 2001 to 2007. Using a least-square fitting and wavelet analysis technique, the phase and amplitude of each <span class="hlt">observed</span> wave were determined, as well as the amplitude growth. Using the dispersion relation of gravity waves, the vertical and horizontal wavelengths were estimated and compared to the horizontal wavelength obtained from the keogram analysis of the <span class="hlt">images</span> <span class="hlt">observed</span> by an all-sky <span class="hlt">imager</span>. The results show that both horizontal and vertical wavelengths, obtained from the dispersion relation and keogram analysis, agree very well for the waves <span class="hlt">observed</span> on the nights of 14 October and 18 December 2006. The determined parameters showed that the <span class="hlt">observed</span> wave on the night of 18 December 2006 had a period of ˜ 43.8 ± 2.19 min, with the horizontal wavelength of 235.66 ± 11.78 km having a downward phase propagation, whereas that of 14 October 2006 propagated with a period of ˜ 36.00 ± 1.80 min with a horizontal wavelength of ˜ 195 ± 9.80 km, and with an upward phase propagation. The <span class="hlt">observation</span> of a wave taken by a photometer and an all-sky <span class="hlt">imager</span> allowed us to conclude that the same wave could be <span class="hlt">observed</span> by both instruments, permitting the investigation of the two-dimensional wave parameter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMSA21B..07W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMSA21B..07W"><span>An Intense Traveling <span class="hlt">Airglow</span> Front in the Upper Mesosphere-Lower Thermosphere with Characteristic of a Turbulent Bore <span class="hlt">Observed</span> over Alice Springs, Australia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walterscheid, R. L.; Hecht, J. H.; Hickey, M. P.; Gelinas, L. J.; Vincent, R. A.; Reid, I. M.; Woithe, J.</p> <p>2010-12-01</p> <p>The Aerospace Corporation’s Nightglow <span class="hlt">Imager</span> <span class="hlt">observed</span> a large step-function change in <span class="hlt">airglow</span> in the form of a traveling front in the OH and O2 <span class="hlt">airglow</span> emissions over Alice Springs Australia on February 2, 2003. The front exhibited a stepwise increase of nearly a factor two in the OH brightness and a stepwise decrease in the O2 brightness. The change in brightness in each layer was associated with a strong leading disturbance followed by a train of weak barely visible waves. The OH <span class="hlt">airglow</span> brightness behind the front was the brightness night for 02 at Alice Springs that we have measured in seven years of <span class="hlt">observations</span>. The OH brightness was among the five brightest. The event was associated with a strong phase-locked two-day wave (TDW).We have analyzed the stability conditions for the upper mesosphere and lower thermosphere and found that the <span class="hlt">airglow</span> layers were found in a region of strong ducting. The thermal structure was obtained from combining data from the SABER instrument on the TIMED satellite and the NRLMSISE-00 model. The wind profile was obtained by combining the HWM07 model and MF radar winds from Buckland Park Australia. We found that the TDW-disturbed profile was significantly more effective in supporting a high degree of ducting than a profile based only on HWM07 winds. Dramatic wall events have been interpreted as manifestations of undular bores (e.g., Smith et al. [2003]). Undular bores are nonlinear high Froude number events that must generate an ever increasing train of waves to carry the excess energy away from the bore front. Only a very weak wave train behind the initial disturbance was seen for the Alice Springs event. The form of the amplitude ordering was not typical of a nonlinear wave train. Therefore a bore interpretation requires another means of energy dissipation, namely turbulent dissipation. We suggest that a reasonable interpretation of the <span class="hlt">observed</span> event is a turbulent bore. We are unaware of any previous event having</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910050764&hterms=1085&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D%2526%25231085','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910050764&hterms=1085&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3D%2526%25231085"><span>Nitrogen <span class="hlt">airglow</span> sources - Comparison of Triton, Titan, and earth</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Strobel, Darrell F.; Meier, R. R.; Summers, Michael E.; Strickland, Douglas J.</p> <p>1991-01-01</p> <p>The individual contributions of direct solar excitation, photoelectron excitation, and magnetospheric electron excitation of Triton and Titan <span class="hlt">airglow</span> <span class="hlt">observed</span> by the Voyager Ultraviolet Spectrometer (UVS) are quantified. The principal spectral features of Triton's <span class="hlt">airglow</span> are shown to be consistent with precipitation of magnetospheric electrons with power dissipation about 500 million W. Solar excitation rates of the dominant N2 and N(+) emission features are factors of 2-7 weaker than magnetospheric electron excitation. On Titan, the calculated disk center and bright limb N(+) 1085 A intensities due to solar excitation agree with <span class="hlt">observed</span> values, while the 970 A feature is mostly N21 c5 band emission. The calculated LBH intensity by photoelectrons suggests that magnetospheric electrons play a minor role in Titan's UV <span class="hlt">airglow</span>. On earth, solar/photoelectron excitation explains the <span class="hlt">observed</span> N(+) 1085 A and LBH intensites and accounts for only 40 percent of the N(+) 916 A intensity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010DPS....42.3625S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010DPS....42.3625S"><span>The Production of Titan's Ultraviolet Nitrogen <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stevens, Michael H.; Gustin, J.; Ajello, J. M.; Evans, J. S.; Meier, R. R.; Stewart, A. I. F.; Esposito, L. W.; McClintock, W. E.; Stephan, A. W.</p> <p>2010-10-01</p> <p>The Cassini Ultraviolet <span class="hlt">Imaging</span> Spectrograph (UVIS) <span class="hlt">observed</span> Titan's dayside limb on 22 June, 2009, obtaining high quality extreme ultraviolet (EUV) and far ultraviolet (FUV) spectra from a distance of only 60,000 km (23 Titan radii). The <span class="hlt">observations</span> reveal the same EUV and FUV emissions arising from photoelectron excitation and photofragmentation of molecular nitrogen (N2) on Earth but with the altitude of peak emission much higher on Titan near 1000 km altitude. In the EUV, emission bands from the photoelectron excited N2 Carroll-Yoshino c4'-X system and N I and N II multiplets arising from photofragmentation of N2 dominate, with no detectable c4'(0,0) emission near 958 Å, contrary to many interpretations of the lower resolution Voyager 1 Ultraviolet Spectrometer data. The FUV is dominated by emission bands from the N2 Lyman-Birge-Hopfield a-X system and additional N I multiplets. We also identify several N2 Vegard-Kaplan A-X bands between 1500-1900 Å, two of which are located near 1561 and 1657 Å where C I multiplets were previously identified from a separate UVIS disk <span class="hlt">observation</span>. We compare these limb emissions to predictions from a terrestrial <span class="hlt">airglow</span> model adapted to Titan that uses a solar spectrum appropriate for these June, 2009 <span class="hlt">observations</span>. Volume production rates and limb radiances are calculated, including extinction by methane and allowance for multiple scattering within the readily excited c4'(0,v') system, and compared to UVIS <span class="hlt">observations</span>. We find that for these <span class="hlt">airglow</span> data only emissions arising from processes involving N2 are present.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010JGRA..11510326H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JGRA..11510326H"><span>Further investigations of lightning-induced transient emissions in the OH <span class="hlt">airglow</span> layer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Tai-Yin; Kuo, C. L.; Chiang, C. Y.; Chen, A. B.; Su, H. T.; Hsu, R. R.</p> <p>2010-10-01</p> <p>A previous study of lightning-induced transient emissions in and below the OH <span class="hlt">airglow</span> layer using <span class="hlt">observations</span> by the <span class="hlt">Imager</span> of Sprites and Upper Atmospheric Lightning (ISUAL) CCD camera onboard the FORMOSAT-II satellite showed that intensity enhancements occurred more frequently in the OH <span class="hlt">airglow</span> layer. Here we show the results of new <span class="hlt">observations</span> made in December 2009 and January 2010 using a narrowband 630 nm filter and spectrophotometer and present further analysis. We estimated the N21P intensity enhancements to be ˜65% and 53% of the total intensity enhancements for the two events we analyzed using ISUAL and the spectrophotometer data in conjunction with a model for emissions of light and VLF perturbations from electromagnetic pulse sources (elves). Our analysis indicates that there is still somewhat considerable intensity enhancement (˜1.25 kR) unaccounted for after the N21P contribution has been removed. Our study suggests that there might be OH emissions in elves and that OH species might also be involved in the lightning-induced process and contribute to the intensity enhancements that we <span class="hlt">observed</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhDT........82A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhDT........82A"><span><span class="hlt">Airglow</span> studies using <span class="hlt">observations</span> made with the GLO instrument on the Space Shuttle</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alfaro Suzan, Ana Luisa</p> <p>2009-12-01</p> <p>Our understanding of Earth's upper atmosphere has advanced tremendously over the last few decades due to our enhanced capacity for making remote <span class="hlt">observations</span> from space. Space based <span class="hlt">observations</span> of Earth's daytime and nighttime <span class="hlt">airglow</span> emissions are very good examples of such enhancements to our knowledge. The terrestrial nighttime <span class="hlt">airglow</span>, or nightglow, is barely discernible to the naked eye as viewed from Earth's surface. However, it is clearly visible from space - as most astronauts have been amazed to report. The nightglow consists of emissions of ultraviolet, visible and near-infrared radiation from electronically excited oxygen molecules and atoms and vibrationally excited OH molecules. It mostly emanates from a 10 km thick layer located about 100 km above Earth's surface. Various photochemical models have been proposed to explain the production of the emitting species. In this study some unique <span class="hlt">observations</span> of Earth's nightglow made with the GLO instrument on NASA's Space Shuttle, are analyzed to assess the proposed excitation models. Previous analyses of these <span class="hlt">observations</span> by Broadfoot and Gardner (2001), performed using a 1-D inversion technique, have indicated significant spatial structures and have raised serious questions about the proposed nightglow excitation models. However, the <span class="hlt">observation</span> of such strong spatial structures calls into serious question the appropriateness of the adopted 1-D inversion technique and, therefore, the validity of the conclusions. In this study a more rigorous 2-D tomographic inversion technique is developed and applied to the available GLO data to determine if some of the apparent discrepancies can be explained by the limitations of the previously applied 1-D inversion approach. The results of this study still reveal some potentially serious inadequacies in the proposed photochemical models. However, alternative explanations for the discrepancies between the GLO <span class="hlt">observations</span> and the model expectations are suggested. These</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18360466','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18360466"><span>Three-channel <span class="hlt">imaging</span> fabry-perot interferometer for measurement of mid-latitude <span class="hlt">airglow</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shiokawa, K; Kadota, T; Ejiri, M K; Otsuka, Y; Katoh, Y; Satoh, M; Ogawa, T</p> <p>2001-08-20</p> <p>We have developed a three-channel <span class="hlt">imaging</span> Fabry-Perot interferometer with which to measure atmospheric wind and temperature in the mesosphere and thermosphere through nocturnal <span class="hlt">airglow</span> emissions. The interferometer measures two-dimensional wind and temperature for wavelengths of 630.0 nm (OI, altitude, 200-300 km), 557.7 nm (OI, 96 km), and 839.9 nm (OH, 86 km) simultaneously with a time resolution of 20 min, using three cooled CCD detectors with liquid-N(2) Dewars. Because we found that the CCD sensor moves as a result of changes in the level of liquid N(2) in the Dewars, the cooling system has been replaced by thermoelectric coolers. The fringe drift that is due to changes in temperature of the etalon is monitored with a frequency-stabilized He-Ne laser. We also describe a data-reduction scheme for calculating wind and temperature from the <span class="hlt">observed</span> fringes. The system is fully automated and has been in operation since June 1999 at the Shigaraki Observatory (34.8N, 136.1E), Shiga, Japan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001596.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-GSFC_20171208_Archive_e001596.html"><span>Waves in <span class="hlt">Airglow</span></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>In April 2012, waves in Earth’s “airglow” spread across the nighttime skies of northern Texas like ripples in a pond. In this case, the waves were provoked by a massive thunderstorm. <span class="hlt">Airglow</span> is a layer of nighttime light emissions caused by chemical reactions high in Earth’s atmosphere. A variety of reactions involving oxygen, sodium, ozone and nitrogen result in the production of a very faint amount of light. In fact, it’s approximately one billion times fainter than sunlight (~10-11 to 10-9 W·cm-2· sr-1). This chemiluminescence is similar to the chemical reactions that light up a glow stick or glow-in-the-dark silly putty. The “day-night band,” of the Visible Infrared <span class="hlt">Imaging</span> Radiometer Suite (VIIRS) on the Suomi NPP satellite captured these glowing ripples in the night sky on April 15, 2012 (top <span class="hlt">image</span>). The day-night band detects lights over a range of wavelengths from green to near-infrared and uses highly sensitive electronics to <span class="hlt">observe</span> low light signals. (The absolute minimum signals detectable are at the levels of nightglow emission.) The lower <span class="hlt">image</span> shows the thunderstorm as <span class="hlt">observed</span> by a thermal infrared band on VIIRS. This thermal band, which is sensitive only to heat emissions (cold clouds appear white), is not sensitive to the subtle visible-light wave structures seen by the day-night band. Technically speaking, <span class="hlt">airglow</span> occurs at all times. During the day it is called “dayglow,” at twilight “twilightglow,” and at night “nightglow.” There are slightly different processes taking place in each case, but in the <span class="hlt">image</span> above the source of light is nightglow. The strongest nightglow emissions are mostly constrained to a relatively thin layer of atmosphere between 85 and 95 kilometers (53 and 60 miles) above the Earth’s surface. Little emission occurs below this layer since there’s a higher concentration of molecules, allowing for dissipation of chemical energy via collisions rather than light production. Likewise, little</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016STP.....2c.106M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016STP.....2c.106M"><span>Night <span class="hlt">airglow</span> in RGB mode</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mikhalev, Aleksandr; Podlesny, Stepan; Stoeva, Penka</p> <p>2016-09-01</p> <p>To study dynamics of the upper atmosphere, we consider results of the night sky photometry, using a color CCD camera and taking into account the night <span class="hlt">airglow</span> and features of its spectral composition. We use night <span class="hlt">airglow</span> <span class="hlt">observations</span> for 2010-2015, which have been obtained at the ISTP SB RAS Geophysical Observatory (52° N, 103° E) by the camera with KODAK KAI-11002 CCD sensor. We estimate the average brightness of the night sky in R, G, B channels of the color camera for eastern Siberia with typical values ranging from ~0.008 to 0.01 erg*cm-2*s-1. Besides, we determine seasonal variations in the night sky luminosities in R, G, B channels of the color camera. In these channels, luminosities decrease in spring, increase in autumn, and have a pronounced summer maximum, which can be explained by scattered light and is associated with the location of the Geophysical Observatory. We consider geophysical phenomena with their optical effects in R, G, B channels of the color camera. For some geophysical phenomena (geomagnetic storms, sudden stratospheric warmings), we demonstrate the possibility of the quantitative relationship between enhanced signals in R and G channels and increases in intensities of discrete 557.7 and 630 nm emissions, which are predominant in the <span class="hlt">airglow</span> spectrum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.8770T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.8770T"><span>Sixteen year variation of horizontal phase velocity and propagation direction of mesospheric and thermospheric waves in <span class="hlt">airglow</span> <span class="hlt">images</span> at Shigaraki, Japan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeo, D.; Shiokawa, K.; Fujinami, H.; Otsuka, Y.; Matsuda, T. S.; Ejiri, M. K.; Nakamura, T.; Yamamoto, M.</p> <p>2017-08-01</p> <p>We analyzed the horizontal phase velocity of gravity waves and medium-scale traveling ionospheric disturbances (MSTIDs) by using the three-dimensional fast Fourier transform method developed by Matsuda et al. (2014) for 557.7 nm (altitude: 90-100 km) and 630.0 nm (altitude: 200-300 km) <span class="hlt">airglow</span> <span class="hlt">images</span> obtained at Shigaraki MU Observatory (34.8°N, 136.1°E, dip angle: 49°) over ˜16 years from 16 March 1999 to 20 February 2015. The analysis of 557.7 nm <span class="hlt">airglow</span> <span class="hlt">images</span> shows clear seasonal variation of the propagation direction of gravity waves in the mesopause region. In spring, summer, fall, and winter, the peak directions are northeastward, northeastward, northwestward, and southwestward, respectively. The difference in east-west propagation direction between summer and winter is probably caused by the wind filtering effect due to the zonal mesospheric jet. Comparison with tropospheric reanalysis data shows that the difference in north-south propagation direction between summer and winter is caused by differences in the latitudinal location of wave sources due to convective activity in the troposphere relative to Shigaraki. The analysis of 630.0 nm <span class="hlt">airglow</span> <span class="hlt">images</span> shows that the propagation direction of MSTIDs is mainly southwestward with a minor northeastward component throughout the 16 years. A clear negative correlation is seen between the yearly power spectral density of MSTIDs and F10.7 solar flux. This negative correlation with solar activity may be explained by the linear growth rate of the Perkins instability and secondary wave generation of gravity waves in the thermosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSM41A2219S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSM41A2219S"><span>Effect of severe geomagnetic disturbances on the atomic oxygen <span class="hlt">airglow</span> emissions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sunil Krishna, M.; Bag, T.</p> <p>2013-12-01</p> <p>The atomic oxygen greenline (557.7nm) and redline emission (630.0 nm) are the most readily <span class="hlt">observed</span> and prominent lines in the nightglow. These emissions can be used as precursors for a variety of physical and chemical processes that occur in the upper mesosphere and lower thermosphere. There are a multitude of effects of space weather on the Earth's atmosphere. The decay of ring current is a very important parameter which can induce variation in the densities of few important species in the atmosphere which are of <span class="hlt">airglow</span> interest. The connection of variation of <span class="hlt">airglow</span> emissions with the extreme space weather conditions is not very well established. In the present study, severe geomagnetic storms and their effect on the <span class="hlt">airglow</span> emissions such as 557.7 nm and 630.0 nm emissions is studied. This study is primarily based on photochemical models with the necessary input obtained from a combination of experimental <span class="hlt">observations</span> and empirical models. We have tried to understand the effect of severe space weather conditions on few very important <span class="hlt">airglow</span> emissions in terms of volume emission rates, change in the peak emission height. Based on the variation an attempt has been made to understand the cause of the variation and further to link the variations in the ring current to the <span class="hlt">airglow</span> chemistry. The study presents the results of calculations performed for the most severe geomagnetic storms occurred over the recent past because of variety of causes on Sun.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.7834S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.7834S"><span>Global modeling of thermospheric <span class="hlt">airglow</span> in the far ultraviolet</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Solomon, Stanley C.</p> <p>2017-07-01</p> <p>The Global <span class="hlt">Airglow</span> (GLOW) model has been updated and extended to calculate thermospheric emissions in the far ultraviolet, including sources from daytime photoelectron-driven processes, nighttime recombination radiation, and auroral excitation. It can be run using inputs from empirical models of the neutral atmosphere and ionosphere or from numerical general circulation models of the coupled ionosphere-thermosphere system. It uses a solar flux module, photoelectron generation routine, and the Nagy-Banks two-stream electron transport algorithm to simultaneously handle energetic electron distributions from photon and auroral electron sources. It contains an ion-neutral chemistry module that calculates excited and ionized species densities and the resulting <span class="hlt">airglow</span> volume emission rates. This paper describes the inputs, algorithms, and code structure of the model and demonstrates example outputs for daytime and auroral cases. Simulations of far ultraviolet emissions by the atomic oxygen doublet at 135.6 nm and the molecular nitrogen Lyman-Birge-Hopfield bands, as viewed from geostationary orbit, are shown, and model calculations are compared to limb-scan <span class="hlt">observations</span> by the Global Ultraviolet <span class="hlt">Imager</span> on the TIMED satellite. The GLOW model code is provided to the community through an open-source academic research license.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994AdSpR..14..177K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994AdSpR..14..177K"><span>Moon based global field <span class="hlt">airglow</span>: For Artemis or any common Lunar Lander</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kozlowski, R. W. H.; Sprague, A. L.; Sandel, B. R.; Hunten, D. M.; Broadfoot, A. L.</p> <p>1994-06-01</p> <p>An inexpensive, small mass, <span class="hlt">airglow</span> experiment consisting of a suite of <span class="hlt">airglow</span> detectors is planned for one or more lunar landers. Solid state detectors measuring light through narrow band filters or concave gratings can integrate emissions from lunar atmospheric constituents and store the information for relay to earth when convenient. The proposed instrument is a simplified version of the Shuttle-borne Arizona <span class="hlt">Imager</span>-Spectrograph. These zenith and near horizon viewing detectors may allow us to monitor fluctuations in atomic species of oxygen, calcium, sodium, potassium, argon, and neon and OH, if present. This choice of <span class="hlt">observations</span> would monitor outgassing from the interior (Ar), meteoritic dust flux (Na, K) solar wind sputtering (O, Ca), and outgassing from the surface (implanted Ne, Na, K). A global network could be inexpensively deployed aboard landers carrying a variety of other selenographic instrumentation. Powered by solar cells such a field network will return data applicable to a wide variety of interplanetary medium and solar-lunar interaction problems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JASTP..93...70P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JASTP..93...70P"><span>Statistical analysis of infrasound signatures in <span class="hlt">airglow</span> <span class="hlt">observations</span>: Indications for acoustic resonance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pilger, Christoph; Schmidt, Carsten; Bittner, Michael</p> <p>2013-02-01</p> <p>The detection of infrasonic signals in temperature time series of the mesopause altitude region (at about 80-100 km) is performed at the German Remote Sensing Data Center of the German Aerospace Center (DLR-DFD) using GRIPS instrumentation (GRound-based Infrared P-branch Spectrometers). Mesopause temperature values with a temporal resolution of up to 10 s are derived from the <span class="hlt">observation</span> of nocturnal <span class="hlt">airglow</span> emissions and permit the identification of signals within the long-period infrasound range.Spectral intensities of wave signatures with periods between 2.5 and 10 min are estimated applying the wavelet analysis technique to one minute mean temperature values. Selected events as well as the statistical distribution of 40 months of <span class="hlt">observation</span> are presented and discussed with respect to resonant modes of the atmosphere. The mechanism of acoustic resonance generated by strong infrasonic sources is a potential explanation of distinct features with periods between 3 and 5 min <span class="hlt">observed</span> in the dataset.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA33A2573T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA33A2573T"><span>Statistical analysis of 16-year phase velocity distribution of mesospheric and ionospheric waves in <span class="hlt">airglow</span> <span class="hlt">images</span>: Comparison between Rikubetsu and Shigaraki, Japan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsuchiya, S.; Shiokawa, K.; Fujinami, H.; Otsuka, Y.; Nakamura, T.; Yamamoto, M.</p> <p>2017-12-01</p> <p>A new spectral analysis technique has been developed to obtain power spectra in the horizontal phase velocity by using the 3-D Fast Fourier Transform [Matsuda et al., JGR, 2014]. Takeo et al. (JGR, 2017) studied spectral parameters of atmospheric gravity waves (AGWs) in the mesopause region and medium-scale traveling ionospheric disturbances (MSTIDs) in the thermosphere over 16 years by using <span class="hlt">airglow</span> <span class="hlt">images</span> at wavelengths of 557.7 nm (emission altitudes: 90-100 km) and 630.0 nm (200-300 km) obtained at Shigaraki (34.8N, 136.1E), Japan. In this study, we have applied the same spectral analysis technique to the 557.7 nm and 630.0-nm <span class="hlt">airglow</span> <span class="hlt">images</span> obtained at Rikubetsu (43.5N, 143.8E), Japan, for 16 years from 1999 to 2014. We compared spectral features of AGWs and MSTIDs over 16 years <span class="hlt">observed</span> at Shigaraki and Rikubetsu, which are separated by 1,174 km. The propagation direction of mesospheric AGWs seen in 557.7-nm <span class="hlt">airglow</span> <span class="hlt">images</span> is northeastward in summer and southwestward in winter at both Shigaraki and Rikubetsu, probably due to wind filtering of these waves by the mesospheric jet. In winter, the propagation direction of AGWs gradually shifted from southwestward to northwestward as time progresses from evening to morning at both stations. We suggest that this local-time shift of propagation direction can also be explained by the wind filtering effect. The propagation direction of AGWs changed from southwestward to northeastward at Rikubetsu on the day of the reversal of eastward zonal wind at 60N and 10 hPa (about 35 km in altitude) by the stratospheric sudden warming (SSW), while such a SSW-associated change was not identified at Shigaraki, indicating that the effect of SSW wind reversal reached only to the Rikubetsu latitudes. For MSTIDs, there is a negative correlation between yearly variation of powers spectral density and F10.7 flux and propagation direction is southwestward in all season at both Shigaraki and Rikubetsu. This negative correlation can be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRA..123.1593X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRA..123.1593X"><span>Longitudinal Thin Structure of Equatorial Plasma Depletions Coincidently <span class="hlt">Observed</span> by Swarm Constellation and all-Sky <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xiong, Chao; Xu, Jiyao; Wu, Kun; Yuan, Wei</p> <p>2018-02-01</p> <p>The lower pair satellites of Swarm mission, flying side-by-side and separated by 1.4° in longitude (about 150 km), usually <span class="hlt">observed</span> equatorial plasma depletions (EPDs) showing quite different structures, and sometime even only one satellite <span class="hlt">observed</span> EPD. In this study, we provided 6-h continuous <span class="hlt">observations</span> of EPDs on the night of 23-24 September 2014, from an all-sky <span class="hlt">imager</span> located at Fuke (geographic:19.5°N,109.1°E), south of China. From the <span class="hlt">airglow</span> <span class="hlt">images</span> the EPDs were found with longitudinal extensions of about 50 km and all tilted from northwest to southeast direction. We further checked the in situ electron density simultaneously measured by the Swarm lower pair satellites and found the differences of Swarm in situ electron densities explained well by the longitudinally thin structure of EPDs <span class="hlt">observed</span> from the all-sky <span class="hlt">imager</span>. During later periods the bifurcation and merging were <span class="hlt">observed</span> by the <span class="hlt">airglow</span> <span class="hlt">images</span>, and it was the first time to report both processes in the evolution of one EPD. The bifurcation was first <span class="hlt">observed</span> at the higher-latitude part, and then <span class="hlt">observed</span> at lower latitudes of EPD. The subbranches generated through bifurcation showed even thinner longitudinal extension of about 20-30 km, and later the subbranches started to merge with each other, forming a really complicated mesh of depleted regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017DPS....4920902P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017DPS....4920902P"><span>Monitoring Saturn's Upper Atmosphere Density Variations Using Helium 584 <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parkinson, Chris</p> <p>2017-10-01</p> <p>The study of He 584 Å brightnesses is interesting as the EUV (Extreme UltraViolet) planetary <span class="hlt">airglow</span> have the potential to yield useful information about mixing and other important parameters in its thermosphere. Resonance scattering of sunlight by He atoms is the principal source of the planetary emission of He 585 Å. The principal parameter involved in determining the He 584 Å albedo are the He volume mixing ratio, f_He, well below the homopause. Our main science objective is to estimate the helium mixing ratio in the lower atmosphere. Specifically, He emissions come from above the homopause where optical depth trau=1 in H2 and therefore the interpretation depends mainly on two parameters: He mixing ratio of the lower atmosphere and K_z. The occultations of Koskinen et al (2015) give K_z with an accuracy that has never been possible before and the combination of occultations and <span class="hlt">airglow</span> therefore provide estimates of the mixing ratio in the lower atmosphere. We make these estimates at several locations that can be reasonably studied with both occultations and <span class="hlt">airglow</span> and then average the results. Our results lead to a greatly improved estimate of the mixing ratio of He in the upper atmosphere and below. The second objective is to constrain the dynamics in the atmosphere by using the estimate of the He mixing ratio from the main objective. Once we have an estimate of the He mixing ratio in the lower atmosphere that agrees with both occultations and <span class="hlt">airglow</span>, helium becomes an effective tracer species as any variations in the Cassini UVIS helium data are direct indicator of changes in K_z i.e., dynamics. Our third objective is to connect this work to our Cassini UVIS data He 584 Å <span class="hlt">airglow</span> analyses as they both cover the time span of the <span class="hlt">observations</span> and allow us to monitor changes in the <span class="hlt">airglow</span> <span class="hlt">observations</span> that may correlate with changes in the state of the atmosphere as revealed by the occultations Saturn's upper thermosphere. This work helps to determine the</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_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" 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_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</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="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006ihy..workE.135S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006ihy..workE.135S"><span>OI 630.0 nm Night <span class="hlt">Airglow</span> <span class="hlt">Observations</span> during the Geomagnetic Storm on November 20, 2003 at Kolhapur (P43)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharma, A. K.; et al.</p> <p>2006-11-01</p> <p>sharma_ashokkumar@yahoo.com The ground based photometric <span class="hlt">observations</span> of OI 630 nm emission line have been carried out from Kolhapur station (Geog. Lat.16.8˚N, Geo. Long 74.2˚E), India during the period of the largest geomagnetic storm of the solar cycle 23 which occurred on 20 November 2003, with minimum Dst index 472 nT occurring around mid-night hours. We <span class="hlt">observed</span> that on 19 November 2003 which was geomagnetically quiet day, the <span class="hlt">airglow</span> activity of OI 630 nm emission was subdued and it was decreasing monotonically. However, on the night of November 20, 2003 the enhancement is <span class="hlt">observed</span> during geomagnetic storm due to the increased electron density at the altitude of the F region which is related to the downward transport of electron from the plasmasphere to the F-region. <span class="hlt">Airglow</span> intensity at OI 630.0 nm showed increase around midnight on November 21, 2003 but comparatively on a smaller scale. On this night the DST index was about 100 nT. This implies that the effect of the geomagnetic storm persisted on that night also. These <span class="hlt">observations</span> have been explained by the penetration magnetospheric electric field to the low latitude region and the subsequent modulation of meridional wind during the magnetic disturbance at night.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.8154K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.8154K"><span>16 years of <span class="hlt">airglow</span> measurement with astronomical facilities</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kausch, Wolfgang; Noll, Stefan; Kimeswenger, Stefan; Unterguggenberger, Stefanie; Jones, Amy; Proxauf, Bastian</p> <p>2017-04-01</p> <p><span class="hlt">Observations</span> taken with ground-based astronomical telescopes are affected by various <span class="hlt">airglow</span> emission processes in the Earth's upper atmosphere. This chemiluminescent emission can be used to investigate the physical state of the meso- and the thermosphere. By applying a modified approach of techniques originally developed to characterise and remove these features from the astronomical spectra, which are not primarily taken for <span class="hlt">airglow</span> studies, these spectra are suitable for <span class="hlt">airglow</span> research. For our studies, we currently use data from two <span class="hlt">observing</span> sites on both hemispheres for our studies: The European Southern Observatory operates four 8m telescopes at the Very Large Telescope (VLT) in the Chilean Atacama desert (24.6°S, 70.4°W). The 2.5m Sloan Digital Sky Survey telescope (SDSS) located in New Mexico/USA (32.8°N, 105.8°W) provides <span class="hlt">observations</span> from the northern hemisphere. Each of these telescopes is equipped with several astronomical instruments. Among them are several spectrographs operating in the optical and near-IR regime with medium to high spectral resolution. Currently, we work on data from the following three spectrographs (1) UVES@VLT (Ultraviolet and Visual Echelle Spectrograph): This instrument provides spectra in the wavelength regime from 0.3 to 1.1μm in small spectral ranges. Its high resolving power (up to R˜110 000) allows a detailed study of oxygen (OI@557nm, OI@630nm), sodium (NaD@589nm), nitrogen (NI@520nm), and many OH bands. UVES has been in operation since 1999 providing the longest time series. (2) X-Shooter@VLT: This spectrograph is unique as it provides the whole wavelength range from 0.3 to 2.5μm at once with medium resolving power (R˜3 300 to 18 000, depending on the setup). This enables us to study the dependency of optical and near-IR <span class="hlt">airglow</span> processes simultaneously, e.g. the OH bands. In addition, weak <span class="hlt">airglow</span> continuum emission, e.g. arising from FeO and NiO can be studied. In operation since 2009, the data cover half a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSA23B..07F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSA23B..07F"><span>HF-enhanced 4278-Å <span class="hlt">airglow</span>: evidence of accelerated ionosphere electrons?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fallen, C. T.; Watkins, B. J.</p> <p>2013-12-01</p> <p>We report calculations from a one-dimensional physics-based self-consistent ionosphere model (SCIM) demonstrating that HF-heating of F-region electrons can produce 4278-Å <span class="hlt">airglow</span> enhancements comparable in magnitude to those reported during ionosphere HF modification experiments at the High-frequency Active Auroral Research Program (HAARP) observatory in Alaska. These artificial 'blue-line' emissions, also <span class="hlt">observed</span> at the EISCAT ionosphere heating facility in Norway, have been attributed to arise solely from additional production of N2+ ions through impact ionization of N2 molecules by HF-accelerated electrons. Each N2+ ion produced by impact ionization or photoionization has a probability of being created in the N2+(1N) excited state, resulting in a blue-line emission from the allowed transition to its ground state. The ionization potential of N2 exceeds 18 eV, so enhanced impact ionization of N2 implies that significant electron acceleration processes occur in the HF-modified ionosphere. Further, because of the fast N2+ emission time, measurements of 4278-Å intensity during ionosphere HF modification experiments at HAARP have also been used to estimate artificial ionization rates. To the best of our knowledge, all <span class="hlt">observations</span> of HF-enhanced blue-line emissions have been made during twilight conditions when resonant scattering of sunlight by N2+ ions is a significant source of 4278-Å <span class="hlt">airglow</span>. Our model calculations show that F-region electron heating by powerful O-mode HF waves transmitted from HAARP is sufficient to increase N2+ ion densities above the shadow height through temperature-enhanced ambipolar diffusion and temperature-suppressed ion recombination. Resonant scattering from the modified sunlit region can cause a 10-20 R increase in 4278-Å <span class="hlt">airglow</span> intensity, comparable in magnitude to artificial emissions measured during ionosphere HF-modification experiments. This thermally-induced artificial 4278-Å aurora occurs independently of any artificial</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720018659','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720018659"><span>An atlas of low latitude 6300A (01) night <span class="hlt">airglow</span> from OGO-4 <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>Reed, E. I.; Fowler, W. B.; Blamont, J. E.</p> <p>1972-01-01</p> <p>The atomic oxygen emission line at 6300 A, measured in the nadir direction by a photometer on the polar orbiting satellite OGO-4, was plotted between 40 deg N and 40 deg S latitude on a series of maps for the moon-free periods between 30 August 1967 and 10 January 1968 The longitudinal and local time variations which occur during the northern fall-winter season are indicated. The northern tropical arc is more widespread while the southern arc is not present at all longitudes. The conditions under which the <span class="hlt">observations</span> were made are described, and four <span class="hlt">airglow</span> maps were selected to show the local time variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA33A2424C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA33A2424C"><span>Relationship between Ripples and Gravity Waves <span class="hlt">Observed</span> in OH <span class="hlt">Airglow</span> over the Andes Lidar Observatory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, B.; Gelinas, L. J.; Liu, A. Z.; Hecht, J. H.</p> <p>2016-12-01</p> <p>Instabilities generated by large amplitude gravity waves are ubiquitous in the mesopause region, and contribute to the strong forcing on the background atmosphere. Gravity waves and ripples generated by instability are commonly detected by high resolution <span class="hlt">airglow</span> <span class="hlt">imagers</span> that measure the hydroxyl emissions near the mesopause ( 87 km). Recently, a method based on 2D wavelet is developed by Gelinas et al. to characterize the statistics of ripple parameters from the Aerospace Infrared Camera at Andes Lidar Observatory located at Cerro Pachón, Chile (70.74°W, 30.25°S). In the meantime, data from a collocated all-sky <span class="hlt">imager</span> is used to derive gravity wave parameters and their statistics. In this study, the relationship between the ripples and gravity waves that appeared at the same time and location are investigated in terms of their orientations, magnitudes and scales, to examine the statistical properties of the gravity wave induced instabilities and the ripples they generate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Icar..307..207N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Icar..307..207N"><span>Extreme ultraviolet spectra of Venusian <span class="hlt">airglow</span> <span class="hlt">observed</span> by EXCEED</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nara, Yusuke; Yoshikawa, Ichiro; Yoshioka, Kazuo; Murakami, Go; Kimura, Tomoki; Yamazaki, Atsushi; Tsuchiya, Fuminori; Kuwabara, Masaki; Iwagami, Naomoto</p> <p>2018-06-01</p> <p>Extreme ultraviolet (EUV) spectra of Venus in the wavelength range 520 - 1480 Å with 3 - 4 Å resolutions were obtained in March 2014 by an EUV <span class="hlt">imaging</span> spectrometer EXCEED (Extreme Ultraviolet Spectroscope for Exospheric Dynamics) on the HISAKI spacecraft. Due to its high sensitivity and long exposure time, many new emission lines and bands were identified. Already known emissions such as the O II 834 Å, O I 989 Å, H ILy - β 1026 Å, and the C I 1277 Å lines (Broadfoot et al., 1974; Bertaux et al., 1980; Feldman et al., 2000) are also detected in the EXCEED spectrum. In addition, N2 band systems such as the Lyman-Birge-Hopfield (a 1Πg - X 1Σg+) (2, 0), (2, 1), (3, 1), (3, 2) and (5, 3) bands, the Birge-Hopfield (b1Πu - X 1 Σg+) (1, 3) band, and the Carroll-Yoshino (c 4‧ 1 Σu+ - X 1Σg+) (0, 0) and (0, 1) bands together are identified for the first time in the Venusian <span class="hlt">airglow</span>. We also identified the CO Hopfield-Birge (B 1Σ+ - X 1Σ+) (1, 0) band in addition to the already known (0, 0) band, and the CO Hopfield-Birge (C 1Σ+ - X 1Σ+) (0, 1), (0, 2) bands in addition to the already known (0, 0) band (Feldman et al., 2000; Gérard et al., 2011).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930026421&hterms=bear&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbear','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930026421&hterms=bear&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbear"><span>Berkeley extreme-ultraviolet <span class="hlt">airglow</span> rocket spectrometer - BEARS</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cotton, D. M.; Chakrabarti, S.</p> <p>1992-01-01</p> <p>The Berkeley EUV <span class="hlt">airglow</span> rocket spectrometer (BEARS) instrument is described. The instrument was designed in particular to measure the dominant lines of atomic oxygen in the FUV and EUV dayglow at 1356, 1304, 1027, and 989 A, which is the ultimate source of <span class="hlt">airglow</span> emissions. The optical and mechanical design of the instrument, the detector, electronics, calibration, flight operations, and results are examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMSA11A..05A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMSA11A..05A"><span>Cassini UVIS <span class="hlt">Observations</span> of Titan Ultraviolet <span class="hlt">Airglow</span> Spectra with Laboratory Modeling from Electron- and Proton-Excited N2 Emission Studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ajello, J. M.; West, R. A.; Malone, C. P.; Gustin, J.; Esposito, L. W.; McClintock, W. E.; Holsclaw, G. M.; Stevens, M. H.</p> <p>2011-12-01</p> <p>Joseph M. Ajello, Robert A. West, Rao S. Mangina Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109 Charles P. Malone Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109 & Department of Physics, California State University, Fullerton, CA 92834 Michael H. Stevens Space Science Division, Naval Research Laboratory, Washington, DC 20375 Jacques Gustin Laboratoire de Physique Atmosphérique et Planétaire, Université de Liège, Liège, Belgium A. Ian F. Stewart, Larry W. Esposito, William E. McClintock, Gregory M. Holsclaw Laboratory for Atmospheric and Space Physics, University of Colorado, Boulder, CO 80303 E. Todd Bradley Department of Physics, University of Central Florida, Orlando, FL 32816 The Cassini Ultraviolet <span class="hlt">Imaging</span> Spectrograph (UVIS) <span class="hlt">observed</span> photon emissions of Titan's day and night limb-<span class="hlt">airglow</span> and disk-<span class="hlt">airglow</span> on multiple occasions, including three eclipse <span class="hlt">observations</span> from 2009 through 2010. The 77 <span class="hlt">airglow</span> <span class="hlt">observations</span> analyzed in this paper show EUV (600-1150 Å) and FUV (1150-1900 Å) atomic multiplet lines and band emissions (lifetimes less than ~100 μs), including the Lyman-Birge-Hopfield (LBH) band system, arising from photoelectron induced fluorescence and solar photo-fragmentation of molecular nitrogen (N2). The altitude of peak UV emission on the limb of Titan during daylight occurred inside the thermosphere/ionosphere (near 1000 km altitude). However, at night on the limb, the same emission features, but much weaker in intensity, arise in the lower atmosphere below 1000 km (lower thermosphere, mesosphere, haze layer) extending downwards to near the surface at ~300 km, possibly resulting from proton- and/or heavier ion-induced emissions as well as secondary-electron-induced emissions. The eclipse <span class="hlt">observations</span> are unique. UV emissions were <span class="hlt">observed</span> during only one of the three eclipse events, and no Vegard-Kaplan (VK) or LBH emissions were seen. Through regression analysis using</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA24A..07G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA24A..07G"><span>Statistical comparisons of gravity wave features derived from OH <span class="hlt">airglow</span> and SABER data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gelinas, L. J.; Hecht, J. H.; Walterscheid, R. L.</p> <p>2017-12-01</p> <p>The Aerospace Corporation's near-IR camera (ANI), deployed at Andes Lidar Observatory (ALO), Cerro Pachon Chile (30S,70W) since 2010, <span class="hlt">images</span> the bright OH Meinel (4,2) <span class="hlt">airglow</span> band. The <span class="hlt">imager</span> provides detailed <span class="hlt">observations</span> of gravity waves and instability dynamics, as described by Hecht et al. (2014). The camera employs a wide-angle lens that views a 73 by 73 degree region of the sky, approximately 120 km x 120 km at 85 km altitude. <span class="hlt">Image</span> cadence of 30s allows for detailed spectral analysis of the horizontal components of wave features, including the evolution and decay of instability features. The SABER instrument on NASA's TIMED spacecraft provides remote soundings of kinetic temperature profiles from the lower stratosphere to the lower thermosphere. Horizontal and vertical filtering techniques allow SABER temperatures to be analyzed for gravity wave variances [Walterscheid and Christensen, 2016]. Here we compare the statistical characteristics of horizontal wave spectra, derived from <span class="hlt">airglow</span> imagery, with vertical wave variances derived from SABER temperature profiles. The analysis is performed for a period of strong mountain wave activity over the Andes spanning the period between June and September 2012. Hecht, J. H., et al. (2014), The life cycle of instability features measured from the Andes Lidar Observatory over Cerro Pachon on March 24, 2012, J. Geophys. Res. Atmos., 119, 8872-8898, doi:10.1002/2014JD021726. Walterscheid, R. L., and A. B. Christensen (2016), Low-latitude gravity wave variances in the mesosphere and lower thermosphere derived from SABER temperature <span class="hlt">observation</span> and compared with model simulation of waves generated by deep tropical convection, J. Geophys. Res. Atmos., 121, 11,900-11,912, doi:10.1002/2016JD024843.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080013140&hterms=saber&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsaber','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080013140&hterms=saber&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsaber"><span>Equatorial Enhancement of the Nighttime OH Mesospheric Infrared <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Baker, D. J.; Mlynczak, M. G.; Russell, J. M.</p> <p>2007-01-01</p> <p>Global measurements of the hydroxyl mesospheric <span class="hlt">airglow</span> over an extended period of time have been made possible by the NASA SABER infrared sensor aboard the TIMED satellite which has been functioning since December of 2001. The orbital mission has continued over a significant portion of a solar cycle. Experimental data from SABER for several years have exhibited equatorial enhancements of the nighttime mesospheric OH (delta v = 2) <span class="hlt">airglow</span> layer consistent with the high average diurnal solar flux. The brightening of the OH <span class="hlt">airglow</span> typically means more H + O3 is being reacted. At both the spring and autumn seasonal equinoxes when the equatorial solar UV irradiance mean is greatest, the peak volume emission rate (VER) of the nighttime Meinel infrared <span class="hlt">airglow</span> typically appears to be both significantly brighter plus lower in altitude by several kilometres at low latitudes compared with midlatitude findings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA33A2423T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA33A2423T"><span>Long-term variation of horizontal phase velocity and propagation direction of mesospheric and thermospheric gravity waves by using <span class="hlt">airglow</span> <span class="hlt">images</span> obtained at Shigarkai, Japan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takeo, D.; Kazuo, S.; Hujinami, H.; Otsuka, Y.; Matsuda, T. S.; Ejiri, M. K.; Yamamoto, M.; Nakamura, T.</p> <p>2016-12-01</p> <p>Atmospheric gravity waves generated in the lower atmosphere transport momentum into the upper atmosphere and release it when they break. The released momentum drives the global-scale pole-to-pole circulation and causes global mass transport. Vertical propagation of the gravity waves and transportation of momentum depend on horizontal phase velocity of gravity waves according to equation about dispersion relation of waves. Horizontal structure of gravity waves including horizontal phase velocity can be seen in the <span class="hlt">airglow</span> <span class="hlt">images</span>, and there have been many studies about gravity waves by using <span class="hlt">airglow</span> <span class="hlt">images</span>. However, long-term variation of horizontal phase velocity spectrum of gravity waves have not been studied yet. In this study, we used 3-D FFT method developed by Matsuda et al., (2014) to analyze the horizontal phase velocity spectrum of gravity waves by using 557.7-nm (altitude of 90-100 km) and 630.0-nm (altitude of 200-300 km) <span class="hlt">airglow</span> <span class="hlt">images</span> obtained at Shigaraki MU Observatory (34.8 deg N, 136.1 deg E) over 16 years from October 1, 1998 to July 26, 2015. Results about 557.7-nm shows clear seasonal variation of propagation direction of gravity waves in the mesopause region. Between summer and winter, there are propagation direction anisotropies which probably caused by filtering due to zonal mesospheric jet and by difference of latitudinal location of wave sources relative to Shigaraki. Results about 630.0-nm shows clear negative correlation between the yearly power spectrum density of horizontal phase velocity and sunspot number. This negative correlation with solar activity is consistent with growth rate of the Perkins instability, which may play an important role in generating the nighttime medium-scale traveling ionospheric disturbances at middle latitudes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P31B2808B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P31B2808B"><span>Energy balance in Saturn's upper atmosphere: Joint Lyman-α <span class="hlt">airglow</span> <span class="hlt">observations</span> with HST and Cassini</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ben-Jaffel, L.; Baines, K. H.; Ballester, G.; Holberg, H. B.; Koskinen, T.; Moses, J. I.; West, R. A.; Yelle, R. V.</p> <p>2017-12-01</p> <p>We are conducting Hubble Space Telescope UV spectroscopy of Saturn's disk-reflected Lyman-α line (Ly-α) at the same time as Cassini <span class="hlt">airglow</span> measurements. Saturn's Ly-α emission is composed of solar and interplanetary (IPH) Ly-α photons scattered by its upper atmosphere. The H I Ly-a line probes different upper atmospheric layers down to the homopause, providing an independent way to investigate the H I abundance and energy balance. However, this is a degenerate, multi-parameter, radiative-transfer problem that depends on: H I column density, scattering process by thermal and superthermal hydrogen, time-variable solar and IPH sources, and instrument calibration. Our joint HST-Cassini campaign should break the degeneracy in the Saturn <span class="hlt">airglow</span> problem. First, line integrated fluxes simultaneously measured by HST/STIS (dayside) and Cassini/UVIS (nightside), avoiding solar variability, should resolve the solar and IPH sources. Second, high-resolution spectroscopy with STIS will reveal superthermal line broadening not accessible with a low-resolution spectrometer like UVIS. Third, a second visit <span class="hlt">observing</span> the same limb of Saturn will cross-calibrate the instruments and, with the STIS linewidth information, will yield the H I abundance, a key photochemical parameter not measured by Cassini. Finally, the STIS latitudinal mapping of the Ly-α linewidth will be correlated with Cassini's latitudinal temperature profile of the thermosphere, to provide an independent constraint on the thermospheric energy budget, a fundamental outstanding problem for giant planets. Here, we report the first results from the HST-Cassini campaign.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980007988','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980007988"><span>Consistency Between SC#21REF Solar XUV Energy Input and the 1973 Pioneer 10 <span class="hlt">Observations</span> of the Jovian Photoelectron Excited H2 <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gangopadhyay, P.; Ogawa, H. S.; Judge, D. L.</p> <p>1988-01-01</p> <p>It has been suggested in the literature that the F74113 solar spectrum for the solar minimum condition needs to be modified to explain the production of photoelectrons in the Earth's atmosphere. We have studied here the effect of another solar minimum spectrum, SC#21REF, on the Jovian upper atmosphere emissions and we have compared the predicted photoelectron excited H2 <span class="hlt">airglow</span> with the 1973 Pioneer 10 <span class="hlt">observations</span>, analyzed according to the methodology of Shemansky and Judge (1988). In this model calculation we find that in 1973, the Jovian H2 band emissions can be accounted for almost entirely by photoelectron excitation, if the preflight calibration of the Pioneer 10 ultraviolet photometer is adopted. If the SC#21REF flux shortward of 250 A is multiplied by 2 as proposed by Richards and Torr (1988) then the Pioneer 10 calibration and/or the <span class="hlt">airglow</span> model used must be modified in order to have a self consistent set of <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e007455.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e007455.html"><span>"Aurora Australis, <span class="hlt">Airglow</span>, Terminator view taken by the Expedition 29 crew"</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-18</p> <p>ISS029-E-007455 (18 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. It features Aurora Australis, <span class="hlt">airglow</span>, Earth?s Terminator and the southeastern Indian Ocean. Nadir coordinates are 51.78 degrees south latitude and 124.41 degrees east longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e007500.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e007500.html"><span>"Aurora Australis, <span class="hlt">Airglow</span>, Terminator view taken by the Expedition 29 crew"</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-18</p> <p>ISS029-E-007500 (18 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. It features the Aurora Australis, <span class="hlt">airglow</span> and parts of the southeastern Indian Ocean. Nadir coordinates are 50.66 degrees south latitude and 137.70 degrees east longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e007502.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e007502.html"><span>"Aurora Australis, <span class="hlt">Airglow</span>, Terminator view taken by the Expedition 29 crew"</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-18</p> <p>ISS029-E-007502 (18 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. It features Aurora Australis, <span class="hlt">airglow</span>, and parts of the southeast Indian Ocean. Nadir coordinates are 50.58 degrees south latitude and 138.28 degrees east longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e007473.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e007473.html"><span>"Aurora Australis, <span class="hlt">Airglow</span>, Terminator view taken by the Expedition 29 crew"</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-18</p> <p>ISS029-E-007473 (18 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. It features Aurora Australis, <span class="hlt">airglow</span>, Earth?s Terminator and parts of the southeast Indian Ocean. Nadir coordinates are 51.53 degrees south latitude and 129.80 degrees east longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940010211','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940010211"><span>Vacuum ultraviolet instrumentation for solar irradiance and thermospheric <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Woods, Thomas N.; Rottman, Gary J.; Bailey, Scott M.; Solomon, Stanley C.</p> <p>1993-01-01</p> <p>A NASA sounding rocket experiment was developed to study the solar extreme ultraviolet (EUV) spectral irradiance and its effect on the upper atmosphere. Both the solar flux and the terrestrial molecular nitrogen via the Lyman-Birge-Hopfield bands in the far ultraviolet (FUV) were measured remotely from a sounding rocket on October 27, 1992. The rocket experiment also includes EUV instruments from Boston University (Supriya Chakrabarti), but only the National Center for Atmospheric Research (NCAR)/University of Colorado (CU) four solar instruments and one <span class="hlt">airglow</span> instrument are discussed here. The primary solar EUV instrument is a 1/4 meter Rowland circle EUV spectrograph which has flown on three rockets since 1988 measuring the solar spectral irradiance from 30 to 110 nm with 0.2 nm resolution. Another solar irradiance instrument is an array of six silicon XUV photodiodes, each having different metallic filters coated directly on the photodiodes. This photodiode system provides a spectral coverage from 0.1 to 80 nm with about 15 nm resolution. The other solar irradiance instrument is a silicon avalanche photodiode coupled with pulse height analyzer electronics. This avalanche photodiode package measures the XUV photon energy providing a solar spectrum from 50 to 12,400 eV (25 to 0.1 nm) with an energy resolution of about 50 eV. The fourth solar instrument is an XUV <span class="hlt">imager</span> that <span class="hlt">images</span> the sun at 17.5 nm with a spatial resolution of 20 arc-seconds. The <span class="hlt">airglow</span> spectrograph measures the terrestrial FUV <span class="hlt">airglow</span> emissions along the horizon from 125 to 160 nm with 0.2 nm spectral resolution. The photon-counting CODACON detectors are used for three of these instruments and consist of coded arrays of anodes behind microchannel plates. The one-dimensional and two-dimensional CODACON detectors were developed at CU by Dr. George Lawrence. The pre-flight and post-flight photometric calibrations were performed at our calibration laboratory and at the Synchrotron Ultraviolet</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740051042&hterms=1079&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3D%2526%25231079','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740051042&hterms=1079&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3D%2526%25231079"><span>Correlation of 1.65 and 2.15 micron <span class="hlt">airglow</span> emissions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kieffaber, L. M.</p> <p>1974-01-01</p> <p>The intense infrared <span class="hlt">airglow</span> is due primarily to vibration-rotation bands of the OH molecule. This <span class="hlt">airglow</span> has been <span class="hlt">observed</span> with a 24-in. scanning photometer at two wavelengths. Narrow-band interference filters are used to limit <span class="hlt">observations</span> to the (9,7) band at 2.15 microns and the (4,2) and (5,3) bands at 1.65 microns. If OH emission results from creation of the excited OH molecule in the v = 9 vibrational state and subsequent cascading through lower vibrational levels, the 1.65 and 2.15 micron radiation will be well correlated in space and time. However, if several mechanisms are involved in producing OH in a variety of initial excitation levels, there is no reason to expect good correlation. Sky maps obtained simultaneously at 1.65 and 2.15 microns show strongly correlated intensity fluctuations. Quantitative analysis of these maps and other investigations of smaller areas of the sky yield correlation coefficients typically in excess of 0.8.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts099-356-026.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts099-356-026.html"><span>Earth limb views with greenish bands of <span class="hlt">airglow</span> during STS-99</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2000-04-06</p> <p>STS099-356-026 (11-22 February 2000) ---Because of its time exposure, this STS-99 35mm frame provides a view of several stars. The thin greenish band above the horizon is <span class="hlt">airglow</span>; radiation emitted by the atmosphere from a layer about 30 kilometers thick and about 100 kilometers altitude. The predominant emission in <span class="hlt">airglow</span> is the green 5577-Angstrom wavelength emission from atomic oxygen atoms. <span class="hlt">Airglow</span> is always and everywhere present in the atmosphere; it results from the recombination of molecules that have been broken apart by solar radiation during the day. But <span class="hlt">airglow</span> is so faint that it can only be seen at night by looking "edge on" at the emission layer, such as the view astronauts have in orbit.</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_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" 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_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <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 brightness 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 brightness wave (MBW) is the phenomenon that the OI (630-nm) <span class="hlt">airglow</span> enhancement propagates poleward once at local midnight. In this study, we first conducted geomagnetically conjugate <span class="hlt">observations</span> of 630nm <span class="hlt">airglow</span> for an MBW at conjugate stations. An <span class="hlt">airglow</span> enhancement which is considered to be an MBW was <span class="hlt">observed</span> in the 630-nm <span class="hlt">airglow</span> <span class="hlt">images</span> 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 <span class="hlt">airglow</span> <span class="hlt">images</span> at Chiang Mai, Thailand (MLAT: 8.9N), which is close to being conjugate point of Kototabang. This result indicates that the MBW does not have geomagnetic conjugacy. We simultaneously <span class="hlt">observed</span> thermospheric neutral winds <span class="hlt">observed</span> by a co-located Fabry-Perot interferometer at Kototabang. The <span class="hlt">observed</span> meridional winds turned from northward (geomagnetically equatorward) to southward (geomagnetically poleward) just before the MBW was <span class="hlt">observed</span>. The bottomside ionospheric heights <span class="hlt">observed</span> by ionosondes rapidly decreased at Kototabang and slightly increased at Chiang Mai simultaneously with the MBW passage. In the presentation, we discuss the MBW generation by the <span class="hlt">observed</span> poleward neutral winds at Kototabang, and the cause of the coinciding small height increase at Chiang Mai by the polarization electric field inside the <span class="hlt">observed</span> MBW at Kototabang.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014cosp...40E.968G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014cosp...40E.968G"><span>Mars dayside temperature from <span class="hlt">airglow</span> limb profiles : comparison with in situ measurements and models</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gérard, Jean-Claude; Bougher, Stephen; Montmessin, Franck; Bertaux, Jean-Loup; Stiepen, A.</p> <p></p> <p>The thermal structure of the Mars upper atmosphere is the result of the thermal balance between heating by EUV solar radiation, infrared heating and cooling, conduction and dynamic influences such as gravity waves, planetary waves, and tides. It has been derived from <span class="hlt">observations</span> performed from different spacecraft. These include in situ measurements of orbital drag whose strength depends on the local gas density. Atmospheric temperatures were determined from the altitude variation of the density measured in situ by the Viking landers and orbital drag measurements. Another method is based on remote sensing measurements of ultraviolet <span class="hlt">airglow</span> limb profiles obtained over 40 years ago with spectrometers during the Mariner 6 and 7 flybys and from the Mariner 9 orbiter. Comparisons with model calculations indicate that they both reflect the CO_2 scale height from which atmospheric temperatures have been deduced. Upper atmospheric temperatures varying over the wide range 270-445 K, with a mean value of 325 K were deduced from the topside scale height of the <span class="hlt">airglow</span> vertical profile. We present an analysis of limb profiles of the CO Cameron (a(3) Pi-X(1) Sigma(+) ) and CO_2(+) doublet (B(2) Sigma_u(+) - X(2) PiΠ_g) <span class="hlt">airglows</span> <span class="hlt">observed</span> with the SPICAM instrument on board Mars Express. We show that the temperature in the Mars thermosphere is very variable with a mean value of 270 K, but values ranging between 150 and 400 K have been <span class="hlt">observed</span>. These values are compared to earlier determinations and model predictions. No clear dependence on solar zenith angle, latitude or season is apparent. Similarly, exospheric variations with F10.7 in the SPICAM <span class="hlt">airglow</span> dataset are small over the solar minimum to moderate conditions sampled by Mars Express since 2005. We conclude that an unidentified process is the cause of the large <span class="hlt">observed</span> temperature variability, which dominates the other sources of temperature variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007AGUSMSA53A..01B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007AGUSMSA53A..01B"><span>Important Considerations When Using Hydroxyl <span class="hlt">Airglow</span> Measurements to Determine Climate Trends of the Mesopause Region.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burns, G.; French, J.</p> <p>2007-05-01</p> <p>Spectral calibrations, <span class="hlt">airglow</span> and possibly auroral contaminations, solar and telluric absorption features and the selection of transition probabilities can all influence rotational temperatures derived from measurements of hydroxyl <span class="hlt">airglow</span> intensities. Consideration and examples are given of these influences. Measurements and analyses are outlined for data checking that should be undertaken if a hydroxyl <span class="hlt">airglow</span> data set is to be used to determine climate trends. Multiple spectral calibrations should be conducted throughout the <span class="hlt">observing</span> period, with regular inter- comparisons to other calibration sources also required. Uncertainties in spectral calibrations should be expressed as a temperature equivalent. Sufficient spectral scans at maximum resolution should be obtained under all extreme <span class="hlt">observing</span> conditions (at the lowest solar depression angle operated both morning and night, moon and cloud both separately and combined, aurora and under conditions of enhanced atomic oxygen <span class="hlt">airglow</span>, and under clear sky conditions but with high atmospheric water vapour content) so that an uncertainty for the derived rotational temperatures can be determined for the established data selection criteria. Once the varying emission and absorption features for the hydroxyl region of interest at your site are understood for the <span class="hlt">observing</span> site, then the spectral resolution of the <span class="hlt">observing</span> instrument can be reduced to increase temporal resolution with reasonable confidence. This confidence should be tested by investigating the average rotational temperatures derived from all possible line intensity ratios under the extreme <span class="hlt">observing</span> conditions noted. If a spectral-fitting rotational temperature determination is used, the residuals from the fit should be summed and similarly examined. Hydroxyl measurements provide a cost effective means of monitoring the temperature of the climate-sensitive mesopause region on an almost nightly basis. If care is taken, they provide a valuable data set</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA33A2576A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA33A2576A"><span>Model simulations of line-of-sight effects in <span class="hlt">airglow</span> <span class="hlt">imaging</span> of acoustic and fast gravity waves from ground and space</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aguilar Guerrero, J.; Snively, J. B.</p> <p>2017-12-01</p> <p>Acoustic waves (AWs) have been predicted to be detectable by <span class="hlt">imaging</span> systems for the OH <span class="hlt">airglow</span> layer [Snively, GRL, 40, 2013], and have been identified in spectrometer data [Pilger et al., JASP, 104, 2013]. AWs are weak in the mesopause region, but can attain large amplitudes in the F region [Garcia et al., GRL, 40, 2013] and have local impacts on the thermosphere and ionosphere. Similarly, fast GWs, with phase speeds over 100 m/s, may propagate to the thermosphere and impart significant local body forcing [Vadas and Fritts, JASTP, 66, 2004]. Both have been clearly identified in ionospheric total electron content (TEC), such as following the 2013 Moore, OK, EF5 tornado [Nishioka et al., GRL, 40, 2013] and following the 2011 Tohoku-Oki tsunami [e.g., Galvan et al., RS, 47, 2012, and references therein], but AWs have yet to be unambiguously <span class="hlt">imaged</span> in MLT data and fast GWs have low amplitudes near the threshold of detection; nevertheless, recent <span class="hlt">imaging</span> systems have sufficient spatial and temporal resolution and sensitivity to detect both AWs and fast GWs with short periods [e.g., Pautet et al., AO, 53, 2014]. The associated detectability challenges are related to the transient nature of their signatures and to systematic challenges due to line-of-sight (LOS) effects such as enhancements and cancelations due to integration along aligned or oblique wavefronts and geometric intensity enhancements. We employ a simulated <span class="hlt">airglow</span> <span class="hlt">imager</span> framework that incorporates 2D and 3D emission rate data and performs the necessary LOS integrations for synthetic <span class="hlt">imaging</span> from ground- and space-based platforms to assess relative intensity and temperature perturbations. We simulate acoustic and fast gravity wave perturbations to the hydroxyl layer from a nonlinear, compressible model [e.g., Snively, 2013] for different idealized and realistic test cases. The results show clear signal enhancements when acoustic waves are <span class="hlt">imaged</span> off-zenith or off-nadir and the temporal evolution of these</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..473H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..473H"><span>Simultaneous 6300 Å <span class="hlt">airglow</span> and radar <span class="hlt">observations</span> of ionospheric irregularities and dynamics at the geomagnetic equator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hickey, Dustin A.; Martinis, Carlos R.; Mendillo, Michael; Baumgardner, Jeffrey; Wroten, Joei; Milla, Marco</p> <p>2018-03-01</p> <p>In March 2014 an all-sky <span class="hlt">imager</span> (ASI) was installed at the Jicamarca Radio Observatory (11.95° S, 76.87° W; 0.3° S MLAT). We present results of equatorial spread F (ESF) characteristics <span class="hlt">observed</span> at Jicamarca and at low latitudes. Optical 6300 and 7774 Å <span class="hlt">airglow</span> <span class="hlt">observations</span> from the Jicamarca ASI are compared with other collocated instruments and with ASIs at El Leoncito, Argentina (31.8° S, 69.3° W; 19.8° S MLAT), and Villa de Leyva, Colombia (5.6° N, 73.52° W; 16.4° N MLAT). We use Jicamarca radar data, in incoherent and coherent modes, to obtain plasma parameters and detect echoes from irregularities. We find that ESF depletions tend to appear in groups with a group-to-group separation around 400-500 km and within-group separation around 50-100 km. We combine data from the three ASIs to investigate the conditions at Jicamarca that could lead to the development of high-altitude, or topside, plumes. We compare zonal winds, obtained from a Fabry-Pérot interferometer, with plasma drifts inferred from the zonal motion of plasma depletions. In addition to the ESF studies we also investigate the midnight temperature maximum and its effects at higher latitudes, visible as a brightness wave at El Leoncito. The ASI at Jicamarca along with collocated and low-latitude instruments provide a clear two-dimensional view of spatial and temporal evolution of ionospheric phenomena at equatorial and low latitudes that helps to explain the dynamics and evolution of equatorial ionospheric/thermospheric processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.P54A..09S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.P54A..09S"><span>Pluto's Far Ultraviolet Spectrum and <span class="hlt">Airglow</span> Emissions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Steffl, A.; Schindhelm, E.; Kammer, J.; Gladstone, R.; Greathouse, T. K.; Parker, J. W.; Strobel, D. F.; Summers, M. E.; Versteeg, M. H.; Ennico Smith, K.; Hinson, D. P.; Linscott, I.; Olkin, C.; Parker, A. H.; Retherford, K. D.; Singer, K. N.; Tsang, C.; Tyler, G. L.; Weaver, H. A., Jr.; Woods, W. W.; Young, L. A.; Stern, A.</p> <p>2015-12-01</p> <p>The Alice far ultraviolet spectrograph on the New Horizons spacecraft is the second in a family of six instruments in flight on, or under development for, NASA and ESA missions. Here, we present initial results from the Alice <span class="hlt">observations</span> of Pluto during the historic flyby. Pluto's far ultraviolet spectrum is dominated by sunlight reflected from the surface with absorption by atmospehric constituents. We tentatively identify C2H2 and C2H4 in Pluto's atmosphere. We also present evidence for weak <span class="hlt">airglow</span> emissions.</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 brightness 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 <span class="hlt">airglow</span> for a midnight brightness 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 <span class="hlt">airglow</span> enhancement that was considered to be an MBW was <span class="hlt">observed</span> in OI (630-nm) <span class="hlt">airglow</span> <span class="hlt">images</span> 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 <span class="hlt">airglow</span> <span class="hlt">images</span> 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 <span class="hlt">airglow</span> 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://images.nasa.gov/#/details-S39-342-026.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-S39-342-026.html"><span>Aurora Australis, Spiked and Sinuous Red and Green <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1991-05-06</p> <p>STS039-342-026 (28 April-6 May 1991) --- This view of the Aurora Australis, or Southern Lights, shows a band of <span class="hlt">airglow</span> above the limb of Earth. Photo experts at NASA studying the mission photography identify the <span class="hlt">airglow</span> as being in the 80-120 kilometer altitude region and attribute its existence to atomic oxygen (wavelength of 5,577 Angstroms), although other atoms can also contribute. The atomic oxygen <span class="hlt">airglow</span> is usually most intense at altitudes around 65 degrees north and south latitude, and is most intense in the spring and fall of the year. The aurora phenomena is due to atmospheric oxygen and nitrogen being excited by the particles from the Van Allen Radiation belts which extend between the two geomagnetic poles. The red and green rays appear to extend upward to 200-300 kilometers, much higher than the usual upper limits of about 110 kilometers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Icar..300..386M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Icar..300..386M"><span>Temperature estimation from hydroxyl <span class="hlt">airglow</span> emission in the Venus night side mesosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Migliorini, A.; Snels, M.; Gérard, J.-C.; Soret, L.; Piccioni, G.; Drossart, P.</p> <p>2018-01-01</p> <p>The temperature of the night side of Venus at about 95 km has been determined by using spectral features of the hydroxyl <span class="hlt">airglow</span> emission around 3 μm, recorded from July 2006 to July 2008 by VIRTIS onboard Venus Express. The retrieved temperatures vary from 145.5 to about 198.1 K with an average value of 176.3 ± 14.3 K and are in good agreement with previous ground-based and space <span class="hlt">observations</span>. The variability with respect to latitude and local time has been studied, showing a minimum of temperature at equatorial latitudes, while temperature values increase toward mid latitudes with a local maximum at about 35°N. The present work provides an independent contribution to the temperature estimation in the transition region between the Venus upper mesosphere and the lower thermosphere, by using the OH emission as a thermometer, following the technique previously applied to the high-resolution O2(a1Δg) <span class="hlt">airglow</span> emissions <span class="hlt">observed</span> from ground.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSA51B2406K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSA51B2406K"><span><span class="hlt">Observations</span> of thunderstorm-related 630 nm <span class="hlt">airglow</span> depletions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kendall, E. A.; Bhatt, A.</p> <p>2015-12-01</p> <p>The Midlatitude All-sky <span class="hlt">imaging</span> Network for Geophysical <span class="hlt">Observations</span> (MANGO) is an NSF-funded network of 630 nm all-sky <span class="hlt">imagers</span> in the continental United States. MANGO will be used to <span class="hlt">observe</span> the generation, propagation, and dissipation of medium and large-scale wave activity in the subauroral, mid and low-latitude thermosphere. This network is actively being deployed and will ultimately consist of nine all-sky <span class="hlt">imagers</span>. These <span class="hlt">imagers</span> form a network providing continuous coverage over the western United States, including California, Oregon, Washington, Utah, Arizona and Texas extending south into Mexico. This network sees high levels of both medium and large scale wave activity. Apart from the widely reported northeast to southwest propagating wave fronts resulting from the so called Perkins mechanism, this network <span class="hlt">observes</span> wave fronts propagating to the west, north and northeast. At least three of these anomalous events have been associated with thunderstorm activity. <span class="hlt">Imager</span> data has been correlated with both GPS data and data from the AIRS (Atmospheric Infrared Sounder) instrument on board NASA's Earth <span class="hlt">Observing</span> System Aqua satellite. We will present a comprehensive analysis of these events and discuss the potential thunderstorm source mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JASTP.171..164H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JASTP.171..164H"><span>Influences of CO2 increase, solar cycle variation, and geomagnetic activity on <span class="hlt">airglow</span> from 1960 to 2015</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Tai-Yin</p> <p>2018-06-01</p> <p>Variations of <span class="hlt">airglow</span> intensity, Volume Emission Rate (VER), and VER peak height induced by the CO2 increase, and by the F10.7 solar cycle variation and geomagnetic activity were investigated to quantitatively assess their influences on <span class="hlt">airglow</span>. This study is an extension of a previous study by Huang (2016) covering a time period of 55 years from 1960 to 2015 and includes geomagnetic variability. Two <span class="hlt">airglow</span> models, OHCD-90 and MACD-90, are used to simulate the induced variations of O(1S) greenline, O2(0,1) atmospheric band, and OH(8,3) <span class="hlt">airglow</span> for this study. Overall, our results demonstrate that <span class="hlt">airglow</span> intensity and the peak VER variations of the three <span class="hlt">airglow</span> emissions are strongly correlated, and in phase, with the F10.7 solar cycle variation. In addition, there is a linear trend, be it increasing or decreasing, existing in the <span class="hlt">airglow</span> intensities and VERs due to the CO2 increase. On other hand, <span class="hlt">airglow</span> VER peak heights are strongly correlated, and out of phase, with the Ap index variation of geomagnetic activity. The CO2 increase acts to lower the VER peak heights of OH(8,3) <span class="hlt">airglow</span> and O(1S) greenline by 0.2 km in 55 years and it has no effect on the VER peak height of O2(0,1) atmospheric band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e005853.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e005853.html"><span><span class="hlt">Airglow</span> on the horizon against the starry sky view taken by the Expedition 29 crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-17</p> <p>ISS029-E-005853 (17 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. The <span class="hlt">image</span> features <span class="hlt">airglow</span> on the horizon against a starry sky with Russian spacecraft Soyuz and Progress in the foreground. Nadir coordinates are 27.8 degrees south latitude and 137.6 west longitude. The photo was taken at 11:32:37 GMT, Sept. 17, 2011.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.P53B2186H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.P53B2186H"><span>What generates Callisto's atmosphere? - Indications from calculations of ionospheric electron densities and <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hartkorn, O. A.; Saur, J.; Strobel, D. F.</p> <p>2016-12-01</p> <p>Callisto's atmosphere has been probed by the Galileo spacecraft and the Hubble Space Telescope (HST) and is expected to be composed of O2 and minor components CO2 and H2O. We use an ionosphere model coupled with a parametrized atmosphere model to calculate ionospheric electron densities and <span class="hlt">airglow</span>. By varying a prescribed neutral atmosphere and comparing the model results to Galileo radio occultation and HST-Cosmic Origin Spectrograph <span class="hlt">observations</span> we find that Callisto's atmosphere likely possesses a day/night asymmetry driven by solar illumination. We see two possible explanation for this asymmetry: 1) If sublimation dominates the atmosphere formation, a day/night asymmetry will be generated since the sublimation production rate is naturally much stronger at the day side than at the night side. 2) If surface sputtering dominates the atmosphere formation, a day/night asymmetry is likely generated as well since the sputtering yield increases with increasing surface temperature and, therefore, with decreasing solar zenith angle. The main difference between both processes is given by the fact that surface sputtering, in contrast to sublimation, is also a function of Callisto's orbital position since sputtering projectiles predominately co-rotate with the Jovian magnetosphere. On this basis, we develop a method that can discriminate between both explanations by comparing <span class="hlt">airglow</span> <span class="hlt">observations</span> at different orbital positions with <span class="hlt">airglow</span> predictions. Our predictions are based on our ionosphere model and an orbital position dependent atmosphere model originally developed for the O2 atmosphere of Europa by Plainaki et al. (2013).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-STS099-355-024.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-STS099-355-024.html"><span>Earth limb views with greenish bands of <span class="hlt">airglow</span> during STS-99</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2000-04-06</p> <p>STS099-355-024 (11-22 February 2000) -- Two separate atmospheric optical phenomena appear in this 35mm photograph captured from the Space Shuttle Endeavour. The thin greenish band above the horizon is <span class="hlt">airglow</span>; radiation emitted by the atmosphere from a layer about 30-kilometers thick and about 100-kilometers' altitude. The predominant emission in <span class="hlt">airglow</span> is the green 5577-Angstrom wavelength emission from atomic oxygen atoms, which is also the predominant emission from the aurora. A yellow-orange color is also seen in <span class="hlt">airglow</span>, which is the emission of the 5800-Angstrom wavelength from sodium atoms. <span class="hlt">Airglow</span> is always present in the atmosphere; it results from the recombination of molecules that have been broken apart by solar radiation during the day. But <span class="hlt">airglow</span> is so faint that it can only be seen at night by looking "edge on" at the emission layer, such as the view that astronauts have in Earth orbit. The other phenomenon in the photo appears to be a faint, diffuse red aurora. Red aurora occur from about 200 kilometers to as high as 500 kilometers altitude only in the auroral zones at polar latitudes. They are caused by the emission of 6300- Angstrom wavelength light from oxygen atoms that have been raised to a higher energy level (excited) by collisions with energetic electrons pouring down from the Earth's magnetosphere. The light is emitted when the atoms return to their original unexcited state. With the red light so faint in this picture, scientists are led to believe that the flux density of incoming electrons was small. Also, since there is no green aurora below the red, that indicates that the energy of the incoming electrons was low - higher energy electrons would penetrate deeper into the atmosphere where the green aurora is energized.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JASTP.171..260A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JASTP.171..260A"><span>On the importance of an atmospheric reference model: A case study on gravity wave-<span class="hlt">airglow</span> interactions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Amaro-Rivera, Yolián; Huang, Tai-Yin; Urbina, Julio</p> <p>2018-06-01</p> <p>The atmospheric reference model utilized in an <span class="hlt">airglow</span> numerical study is important since <span class="hlt">airglow</span> emissions depend on the number density of the light-emitting species. In this study, we employ 2-dimensional, nonlinear, time-dependent numerical models, Multiple <span class="hlt">Airglow</span> Chemistry Dynamics (MACD) and OH Chemistry Dynamics (OHCD), that use the MSISE-90, NRLMSISE-00, and Garcia and Solomon (GS) model data as atmospheric reference models, to investigate gravity wave-induced <span class="hlt">airglow</span> variations for the OH(8,3) <span class="hlt">airglow</span>, O2(0,1) atmospheric band, and O(1S) greenline emissions in the Mesosphere and Lower Thermosphere (MLT) region. Our results show that the OHCD-00 produces the largest wave-induced OH(8,3) <span class="hlt">airglow</span> intensity variation (∼34%), followed by the OHCD-90 (∼30%), then by the OHCD (∼22%). For O(1S) greenline, the MACD produces the largest wave-induced variation (∼33%), followed by the MACD-90 (∼28%), then by MACD-00 (∼26%). As for O2(0,1) atmospheric band, the MACD produces the largest wave-induced variation (∼31%), followed by the MACD-90 and MACD-00 (∼29%). Our study illustrates the importance and the need for a good atmospheric reference model that can accurately represent the atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.8423K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.8423K"><span>Investigating the Concept of Using <span class="hlt">Airglow</span> Measurements to Detect Seismicity on Venus</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kenda, Balthasar; Lognonné, Philippe; Komjathy, Attila; Banerdt, Bruce; Cutts, Jim; Soret, Lauriane; Jackson, Jennifer</p> <p>2017-04-01</p> <p>The internal structure and dynamics of Venus are poorly constrained by <span class="hlt">observations</span>. Seismology is among the best candidates for probing the interior of the planet, and it would also provide indispensable information about the present-day tectonic activity of Venus. However, due to the extreme surface temperatures, a long-duration seismic station seems to be beyond the technical capabilities achievable today. Nonetheless, the thick and dense atmosphere, which strongly couples with the ground, gives rise to the attractive option of detecting seismic waves from quakes within the atmosphere itself (Garcia et al., 2005, Lognonné and Johnson, 2007, 2015) using in-situ or remote-sensing measurements (Cutts et al., 2015). Here, we consider the bright <span class="hlt">airglow</span> emission of O2 at 1.27 μm on the nightside of Venus and we model the intensity fluctuations induced by Venus quakes. Synthetic seismograms in the <span class="hlt">airglow</span> layer, at 90-120 km altitude, are computed using normal-mode summation for a fully coupled solid planet-atmosphere system, including the effects of molecular relaxation of CO2 and a radiative boundary condition at the top of the atmosphere (Lognonné et al., 2016). The corresponding variations in the volumetric emission rate, calculated for realistic background intensities of the <span class="hlt">airglow</span> (Soret et al., 2012), are then vertically integrated to reproduce the signals that would be seen from orbit. The noise level of existing <span class="hlt">airglow</span> cameras suggests that the Rayleigh waves generated by quakes of magnitude 5 and above occurring on the nightside of the planet may be detectable up to about 60° in epicentral distance. A significant advantage of this technique is that a single orbiting camera may be sufficient to serve the role of a seismic network. By identifying and tracking the waves it is indeed possible to locate the source, estimate the magnitude and measure the horizontal surface-wave propagation velocities on Venus. In particular, it is expected that this would</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19720029859&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720029859&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dtwilight"><span>Vacuum ultraviolet spectra of the late twilight <span class="hlt">airglow</span>.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Buckley, J. L.; Moos, H. W.</p> <p>1971-01-01</p> <p>Evaluation of sounding rocket spectra of the late twilight (solar-zenith angle of 120 deg) ultraviolet <span class="hlt">airglow</span> between 1260 and 1900 A. The only <span class="hlt">observed</span> features are O I 1304 and 1356. When the instrument looked at an elevation of 17 deg above the western horizon, the brightnesses were 70 and 33 rayleighs, respectively. The upper limits on the total intensity of the Lyman-Birge-Hopfield and Vegard-Kaplan systems of N2 were 26 plus or minus 26 and 55 plus or minus 55 rayleighs, respectively. An estimate shows that a large part of the O I emissions may be due to excitation by conjugate-point electrons.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhDT.......206C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhDT.......206C"><span>Studies of the polar MLT region using SATI <span class="hlt">airglow</span> measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cho, Youngmin</p> <p></p> <p>To investigate atmospheric dynamics of the MLT (Mesosphere and Lower Thermosphere) region, a ground-based instrument called SATI (Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span>) was developed at York University. The rotational temperatures and emission rates of the OH (6-2) Meinel band and the O2 (0-1) Atmospheric band have been measured in the MLT region by the SATI instrument at Resolute Bay (74.68°N, 94.90°W) since November, 2001, and at the King Sejong station (62.22°S, 58.75°W) since February, 2002. The MLT measurements are examined for periodic oscillations in the ambient temperature and <span class="hlt">airglow</span> emission rate. A dominant and coherent 4-hr oscillation is seen in both the OH and O2 temperature and emission rate at Resolute Bay in November, 2001. Tidal variation with a 12 hour period is shown in hourly averaged temperatures of the season 2001--2002 and the season 2003--2004. In addition, planetary waves with periods of 3 and 4.5 days are also seen in a longer interval. The <span class="hlt">observations</span> at high latitudes have revealed that temperatures and emission rates are higher around the winter solstice. MLT cooling events were found at Resolute Bay in December, 2001 and February, 2002. They are compared with the UKMO (UK Meteorological Office) stratospheric assimilated data, and the MLT coolings coincide in time with the stratospheric warmings. A consistent inverse relationship of the OH temperatures and temperatures at 0.316 hPa is presented in the comparison. In previous studies of wave perturbations, the background (mean) values were normally subtracted from the instantaneous signal, but in the present investigation this was not done, allowing the long-term relationship to be examined. A positive relationship of the temperature and emission rate is seen from the SATI measurements for both short and long-term variations, suggesting that similar dynamical processes are responsible for both. This relationship is supported by satellite data from the SABER (Sounding of the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10562E..0WL','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10562E..0WL"><span>Atise: a miniature Fourier-transform spectro-<span class="hlt">imaging</span> concept for surveying auroras and <span class="hlt">airglow</span> monitoring from a 6/12u cubesat</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Le Courer, E.; Barthelemy, M.; Vialatte, A.; Prugniaux, M.; Bourdarot, G.; Sequies, T.; Monsinjon, P.; Puget, R.; Guerineau, N.</p> <p>2017-09-01</p> <p>The nanosatellite ATISE is a mission dedicated to the <span class="hlt">observation</span> of the emission spectra of the upper atmosphere (i.e. <span class="hlt">Airglow</span> and Auroras) mainly related to both the solar UV flux and the precipitation of suprathermal particles coming from the solar wind through the magnetosphere. ATISE will measure specifically the auroral emissions, and the <span class="hlt">airglow</span> (day- and night) in the spectral range between 380 and 900 nm at altitudes between 100 and 350 km. The exposure time will be 1 second in auroral region and 20 s at low latitude regions. The 5 year expected lifetime of this mission should cover almost a half of solar cycle (2 years nominal). This instrument concept is based on an innovative miniaturized Fourier-transform spectrometer (FTS) allowing simultaneous 1 Rayleigh sensitivity detection along six 1.5°x1° limb lines of sight. This 1-2kg payload instrument is hosted in a 12U cubeSat where 6U are allocated to the payload and 6U to the plateform subsystems. This represents a miniaturisation by a factor of 500 on weight and volume compared to previous Arizona-GLO instrument for equivalent performances in the visible. The instrument is based on microSPOC concept developed by ONERA and IPAG using one Fizeau interferometer per line of sight directly glued on top of the half of a very sensitive CMOS Pyxalis HDPYX detector. Three detectors are necessary with a total electrical consumption compatible with a 6U nanoSat. Each interferometer occupies a 1.4 M pixel part of detector, each is placed on an <span class="hlt">image</span> of the entrance pupil corresponding to a unique direction of the six lines of sight, this in order to have a uniform illumination permitting good spectral Fourier reconstruction from fringes created between the Fizeau plate and the detector itself. Despite a limited 8x6 cm telescope, this configuration takes advantage of FTS multiplex effect and permits us to maximize the throughput and to integrate very faint emission lines over a wide field of view even if the 1</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930009172','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930009172"><span>Visible <span class="hlt">Airglow</span> Experiment data analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Abreu, Vincent J.</p> <p>1990-01-01</p> <p>The Visible <span class="hlt">Airglow</span> Experiment (VAE) was designed to provide detailed profiles of the distribution of excited states of atoms and molecules in the upper atmosphere. The studies supported during the funding period (1983 - 1989) have made significant contributions in the area of thermospheric aeronomy, and the progress during the first four years of this period has been reviewed by Hays et al. (1988). The investigations carried out have resulted in more than 20 publications, and these are summarized.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P11C2518S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P11C2518S"><span>Pluto's Ultraviolet <span class="hlt">Airglow</span> and Detection of Ions in the Upper Atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Steffl, A.; Young, L. A.; Kammer, J.; Gladstone, R.; Hinson, D. P.; Summers, M. E.; Strobel, D. F.; Stern, S. A.; Weaver, H. A., Jr.; Olkin, C.; Ennico Smith, K.</p> <p>2017-12-01</p> <p>In July 2015, the Alice ultraviolet spectrograph aboard the New Horizons spacecraft made numerous <span class="hlt">observations</span> of Pluto and its atmosphere. We present here the far ultraviolet reflectance spectrum of Pluto and <span class="hlt">airglow</span> emissions from its atmosphere. At wavelengths greater than 1400Å, Pluto's spectrum is dominated by sunlight reflected from the surface of the planet. Various hydrocarbon species such as C2H4 are detected in absorption of the solar continuum. Below 1400Å, Pluto's atmosphere is opaque and the surface cannot be detected. However, after carefully removing various sources of background light, we see extremely faint <span class="hlt">airglow</span> emissions (<0.05 Rayleighs/Ångstrom) from Pluto's atmosphere. All of the emissions are produced by nitrogen in various forms: molecular, atomic, and singly ionized. The detection of N+ at 1086Å is the first, and thus far only, direct detection of ions in Pluto's atmosphere. This N+ emission line is produced primarily by dissociative photoionization of molecular N2 by solar EUV photons (energy > 34.7 eV; wavelength < 360Å). Notably absent from Pluto's spectrum are emission lines from argon at 1048 and 1067Å. We place upper limits on the amount of argon in Pluto's atmosphere above the tau=1 level (<span class="hlt">observed</span> to be at 750km tangent altitude) that are significantly lower than pre-encounter atmospheric models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940022877','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940022877"><span>Solar and <span class="hlt">airglow</span> measurements aboard the two suborbital flights NASA 36.098 and 36.107</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Woods, Thomas N.</p> <p>1994-01-01</p> <p>This suborbital program, involving the University of Colorado (CU), National Center for Atmospheric Research (NCAR), University of California at Berkeley (UCB), and Boston University (BU), has resulted in two rocket flights from the White Sands Missile Range, one in 1992 and one in 1993 as NASA 36.098 and 36.107 respectively. The rocket payload includes five solar instruments and one <span class="hlt">airglow</span> instrument from CU/NCAR and one solar instrument and two <span class="hlt">airglow</span> instruments from UCB/BU. This report discusses results on solar radiation measurements and the study of thermospheric <span class="hlt">airglow</span>, namely the photoelectron excited emissions from N2 and O, for the CU/NCAR program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUSMSA43A..03W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUSMSA43A..03W"><span>First Light from Triple-Etalon Fabry-Perot Interferometer for Atmospheric OI <span class="hlt">Airglow</span> (6300 A)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Watchorn, S.; Noto, J.; Pedersen, T.; Betremieux, Y.; Migliozzi, M.; Kerr, R. B.</p> <p>2006-05-01</p> <p>Scientific Solutions, Inc. (SSI) has developed a triple-etalon Fabry-Perot interferometer (FPI) to <span class="hlt">observe</span> neutral winds in the ionosphere by measuring neutral oxygen (O I) emission at 630.0 nm during the day. This instrument is to be deployed in the SSI <span class="hlt">airglow</span> building at the Cerro Tololo observatory (30.17S 70.81W) in Chile, in support of the Comm/Nav Outage Forecast System (C/NOFS) project. Post-deployment <span class="hlt">observation</span> will be made in conjunction with two other Clemson University Fabry-Perots in Peru, creating a longitudinal chain of interferometers for thermospheric <span class="hlt">observations</span>. These instruments will make autonomous day and night <span class="hlt">observations</span> of thermospheric dynamics. Instruments of this type can be constructed for a global chain of autonomous <span class="hlt">airglow</span> observatories. The FPI presented in this talk consists of three independently pressure-controlled etalons, fed collimated light by a front optical train headed by an all-sky lens with a 160-degree field of view. It can be controlled remotely via a web-based service which allows any internet-connected computer to mimic the control computer at the instrument site. In fall 2005, the SSI system was first assembled at the Millstone Hill Observatory in Westford, Massachusetts, and made day and evening <span class="hlt">observations</span>. It was then moved to the High-frequency Active Auroral Research Project (HAARP) site in Gakona, Alaska, to participate in joint optical/ionospheric heating campaigns. Additionally, natural <span class="hlt">airglow</span> <span class="hlt">observations</span> were made, both locally and remotely via the internet from Massachusetts. The Millstone and HAARP <span class="hlt">observations</span> with two etalons yielded strong 630-nm atmospheric Fraunhofer absorption lines, with some suggestion of the Ring effect. By modeling the atmospheric absorption line as the constant times the corresponding solar absorption -- itself modeled as a Gaussian plus a polynomial -- the absorption feature is subtracted, leaving only the emission feature. Software ring-summing tools developed at the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950029567&hterms=university+college&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Duniversity%2Bcollege','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950029567&hterms=university+college&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Duniversity%2Bcollege"><span>Interpretation of satellite <span class="hlt">airglow</span> <span class="hlt">observations</span> during the March 22, 1979, magnetic storm, using the coupled ionosphere-thermosphere model developed at University College, London</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Parish, H. F.; Gladstone, G. R.; Chakrabarti, S.</p> <p>1994-01-01</p> <p>The University of California, Berkeley, extreme ultraviolet spectrometer aboard the U.S. Air Force STP 78-1 satellite measured emission features in the Earth's dayglow due to neutral and ionized species in the atmosphere, in the 35 to 140-nm range. The spectrometer was operating between March 1979 and March 1980, including the period of the magnetic storm on March 22, 1979. Some of these measurements are interpreted using the predictions of the three-dimensional time-dependent coupled ionosphere-thermosphere model developed at University College, London. The <span class="hlt">observations</span> show a reduction in the atomic oxygen 130.4-nm <span class="hlt">airglow</span> emission at high northern latitudes following the storm. Model simulations show that this reduction in 130.4-nm emission is associated with an increase in the O2/O ratio. Analysis of model results using electron transport and radiative transport codes show that the brightness of 130.4-nm emission at high latitudes due to resonantly scattered sunlight is approximately twice that due to photoelectron impact excitation. However, the <span class="hlt">observed</span> decrease in the brightness at high northern latitudes is mainly due to a change in the photoelectron impact source, which contributes approximately 75% of the total, as well as its multiple scattering component; for the photoelectron impact source at 70 deg latitude and 200 km altitude, the reduction in multiple scattering is 1.5 times greater than the reduction in the initial excitation. The reduction in the <span class="hlt">airglow</span> emission is visible only in the norther n hemisphere because the south pole was not sunlit over the storm period. The comparison of model results with <span class="hlt">observations</span> suggests that 130.4-nm emission may be useful as a tracer for global changes in the concentration of atomic energy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JASTP..78...62T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JASTP..78...62T"><span>Simultaneous Rayleigh lidar and <span class="hlt">airglow</span> measurements of middle atmospheric waves over low latitudes in India</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Taori, A.; Kamalakar, V.; Raghunath, K.; Rao, S. V. B.; Russell, J. M.</p> <p>2012-04-01</p> <p>We utilize simultaneous Rayleigh lidar and mesospheric OH and O2 <span class="hlt">airglow</span> measurements to identify the dominant and propagating waves within 40-95 km altitude regions over a low latitude station Gadanki (13.8° N, 79.2 °E). It is found that waves with 0.4-0.6 h periodicity are common throughout the altitude range of 40-95 km with significant amplitudes. The ground based temperature measurements with lidar and <span class="hlt">airglow</span> monitoring are found to compare well with SABER data. With simultaneous Rayleigh lidar (temperature) and mesospheric <span class="hlt">airglow</span> (emission intensity and temperature) measurements, we estimate the amplitude growth and Krassovsky parameters to characterize the propagation and dissipation of these upward propagating waves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA33A2578H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA33A2578H"><span>Space-borne <span class="hlt">observation</span> of mesospheric bore by Visible and near Infrared Spectral <span class="hlt">Imager</span> onboard the International Space Station</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hozumi, Y.; Saito, A.; Sakanoi, T.; Yamazaki, A.; Hosokawa, K.</p> <p>2017-12-01</p> <p>Mesospheric bores were <span class="hlt">observed</span> by Visible and near Infrared Spectral <span class="hlt">Imager</span> (VISI) of the ISS-IMAP mission (Ionosphere, Mesosphere, upper Atmosphere and Plasmasphere mapping mission from the International Space Station) in O2 <span class="hlt">airglow</span> at 762 nm wavelength. The mesospheric bore is moving front of sharp jump followed by undulations or turbulence in the mesopause region. Since previous studies of mesospheric bore were mainly based on ground-based <span class="hlt">airglow</span> <span class="hlt">imaging</span> that is limited in field-of-view and <span class="hlt">observing</span> site, little is known about its horizontal extent and global behavior. Space-borne <span class="hlt">imaging</span> by ISS-IMAP/VISI provides an opportunity to study the mesospheric bore with a wide field-of-view and global coverage. A mesospheric bore was captured by VISI in two consecutive paths on 9 July 2015 over the south of African continent (48ºS - 54ºS and 15ºE). The wave front aligned with south-north direction and propagated to west. The phase velocity and wave length of the following undulation were estimated to 100 m/s and 30 km, respectively. Those parameters are similar to those reported by previous studies. 30º anti-clockwise rotation of the wave front was recognized in 100 min. Another mesospheric bore was captured on 9 May 2013 over the south Atlantic ocean (35ºS - 43ºS and 24ºW - 1ºE) with more than 2,200 km horizontal extent of wave front. The wave front aligned with southeast-northwest direction. Because the following undulation is recognized in the southwest side of the wave front, it is estimated to propagate to northeast direction. The wave front was modulated with 1,000 km wave length. This modulation implies inhomogeneity of the phase velocity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000038366','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000038366"><span>Gravity Wave Energetics Determined From Coincident Space-Based and Ground-Based <span class="hlt">Observations</span> of <span class="hlt">Airglow</span> Emissions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2000-01-01</p> <p>Significant progress was made toward the goals of this proposal in a number of areas during the covered period. Section 5.1 contains a copy of the originally proposed schedule. The tasks listed below have been accomplished: (1) Construction of space-based <span class="hlt">observing</span> geometry gravity wave model. This model has been described in detail in the paper accompanying this report (Section 5.2). It can simulate the <span class="hlt">observing</span> geometry of both ground-based, and orbital instruments allowing comparisons to be made between them. (2) Comparisons of relative emission intensity, temperatures, and Krassovsky's ratio for space- and ground-based <span class="hlt">observing</span> geometries. These quantities are used in gravity wave literature to describe the effects of the waves on the <span class="hlt">airglow</span>. (3) Rejection of Bates [1992], and Copeland [1994] chemistries for gravity wave modeling purposes. Excessive 02(A(sup 13)(Delta)) production led to overproduction of O2(b(sup 1)(Sigma)), the state responsible for the emission of O2. Atmospheric band. Attempts were made to correct for this behavior, but could not adequately compensate for this. (4) Rejection of MSX dataset due to lack of coincident data, and resolution necessary to characterize the waves. A careful search to identify coincident data revealed only four instances, with only one of those providing usable data. Two high latitude overpasses and were contaminated by auroral emissions. Of the remaining two mid-latitude coincidences, one overflight was obscured by cloud, leaving only one ten minute segment of usable data. Aside from the statistical difficulties involved in comparing measurements taken in this short period, the instrument lacks the necessary resolution to determine the vertical wavelength of the gravity wave. This means that the wave cannot be uniquely characterized from space with this dataset. Since no <span class="hlt">observed</span> wave can be uniquely identified, model comparisons are not possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770003789','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770003789"><span>Post sunset behavior of the 6300 A atomic oxygen <span class="hlt">airglow</span> emission</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smith, R. E.</p> <p>1976-01-01</p> <p>A theoretical model of the 6300 A OI <span class="hlt">airglow</span> emission was developed based on the assumptions that both the charged and neutral portions of the Earth's upper atmosphere are in steady state conditions of diffusive equilibrium. Intensities of 6300 A OI emission line were calculated using electron density true height profiles from a standard C-4 ionosonde and exospheric temperatures derived from Fabry-Perot interferometer measurements of the Doppler broadened 6300 A emission line shape as inputs to the model. Reaction rate coefficient values, production mechanism efficiencies, solar radiation fluxes, absorption cross sections, and models of the neutral atmosphere were varied parametrically to establish a set of acceptable inputs which will consistently predict 6300 A emission intensities that closely agree with intensities <span class="hlt">observed</span> during the post-sunset twilight period by an <span class="hlt">airglow</span> observatory consisting of a Fabry-Perot interferometer and a turret photometer. Emission intensities that can only result from the dissociative recombination of molecular oxygen ions were <span class="hlt">observed</span> during the latter portion of the <span class="hlt">observational</span> period. Theoretical calculations indicate that contamination of the 6300 A OI emission should be on the order of or less than 3 percent; however, these results are very sensitive to the wavelengths of the individual lines and their intensities relative to the 6300 A OI intensity. This combination of a model atmosphere, production mechanism efficiencies, and quenching coefficient values was used when the dissociative photoexcitation and direct impact excitation processes were contributing to the intensity to establish best estimates of solar radiation fluxes in the Schumann--Runge continuum and associated absorption cross sections. Results show that the Jacchia 1971 model of the upper atmosphere combined with the Ackerman recommended solar radiation fluxes and associated absorption cross sections produces theoretically calculated intensities that more</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20733778','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20733778"><span>Berkeley extreme-ultraviolet <span class="hlt">airglow</span> rocket spectrometer: BEARS.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cotton, D M; Chakrabarti, S</p> <p>1992-09-20</p> <p>We describe the Berkeley extreme-UV <span class="hlt">airglow</span> rocket spectrometer, which is a payload designed to test several thermospheric remote-sensing concepts by measuring the terrestrial O I far-UV and extreme-UV dayglow and the solar extreme-UV spectrum simultaneously. The instrument consisted of two near-normal Rowland mount spectrometers and a Lyman-alpha photometer. The dayglow spectrometer covered two spectral regions from 980 to 1040 A and from 1300 to 1360 A with 1.5-A resolution. The solar spectrometer had a bandpass of 250-1150 A with an ~ 10-A resolution. All three spectra were accumulated by using a icrochannel-plate-intensified, two-dimensional <span class="hlt">imaging</span> detector with three separate wedge-and strip anode readouts. The hydrogen Lyman-alpha photometer was included to monitor the solar Lyman-alpha irradiance and geocoronal Lyman-alpha emissions. The instrument was designed, fabricated, and calibrated at the University of California, Berkeley and was successfully launched on 30 September 1988 aboard the first test flight of a four-stage sounding rocket, Black Brant XII.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19810043952&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DAbreu%252C%2Bc.','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810043952&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DAbreu%252C%2Bc."><span>The O II /7320-7330 A/ <span class="hlt">airglow</span> - A morphological study</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yee, J. H.; Abreu, V. J.; Hays, P. B.</p> <p>1981-01-01</p> <p>A statistical study of the 7320-30 A <span class="hlt">airglow</span> arising from the metastable transition between aP and aD states of atomic oxygen ions was conducted by analyzing the data taken from the visible <span class="hlt">airglow</span> experiment on the Atmosphere Explorer satellites C and E during the time periods between 1974 and 1979. Averaged column emission rate profiles as a function of solar zenith angle and solar activity variation are presented. The galactic background has been carefully subtracted. The result shows that the rate of decreasing emission as a function of solar zenith angle agrees with the theoretical calculation based upon a neutral atmosphere model and the solar spectrum as measured by the EUV spectrometer on the Atmosphere Explorer satellite. Furthermore, an expected increase with solar activity also appeared in a plot of emission brightness versus solar 10.7-cm flux.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980236655','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980236655"><span>Continuing Studies in Support of Ultraviolet <span class="hlt">Observations</span> of Planetary Atmospheres</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Clark, John</p> <p>1997-01-01</p> <p>This program was a one-year extension of an earlier Planetary Atmospheres program grant, covering the period 1 August 1996 through 30 September 1997. The grant was for supporting work to complement an active program <span class="hlt">observing</span> planetary atmospheres with Earth-orbital telescopes, principally the Hubble Space Telescope (HST). The recent concentration of this work has been on HST <span class="hlt">observations</span> of Jupiter's upper atmosphere and aurora, but it has also included <span class="hlt">observations</span> of Io, serendipitous <span class="hlt">observations</span> of asteroids, and <span class="hlt">observations</span> of the velocity structure in the interplanetary medium. The <span class="hlt">observations</span> of Jupiter have been at vacuum ultraviolet wavelengths, including <span class="hlt">imaging</span> and spectroscopy of the auroral and <span class="hlt">airglow</span> emissions. The most recent HST <span class="hlt">observations</span> have been at the same time as in situ measurements made by the Galileo orbiter instruments, as reflected in the meeting presentations listed below. Concentrated efforts have been applied in this year to the following projects: The analysis of HST WFPC 2 <span class="hlt">images</span> of Jupiter's aurora, including the Io footprint emissions. We have performed a comparative analysis of the lo footprint locations with two magnetic field models, studied the statistical properties of the apparent dawn auroral storms on Jupiter, and found various other repeated patterns in Jupiter's aurora. Analysis and modeling of <span class="hlt">airglow</span> and auroral Ly alpha emission line profiles from Jupiter. This has included modeling the aurora] line profiles, including the energy degradation of precipitating charged particles and radiative transfer of the emerging emissions. Jupiter's auroral emission line profile is self-absorbed, since it is produced by an internal source, and the resulting emission with a deep central absorption from the overlying atmosphere permits modeling of the depth of the emissions, plus the motion of the emitting layer with respect to the overlying atmospheric column from the <span class="hlt">observed</span> Doppler shift of the central absorption. By contrast</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA51A2384L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA51A2384L"><span>A new method of derived equatorial plasma bubbles motion by tracing OI 630 nm emission all-sky <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, M.; Yu, T.; Chunliang, X.; Zuo, X.; Liu, Z.</p> <p>2017-12-01</p> <p>A new method for estimating the equatorial plasma bubbles (EPBs) motions from <span class="hlt">airglow</span> emission all-sky <span class="hlt">images</span> is presented in this paper. This method, which is called 'cloud-derived wind technology' and widely used in satellite <span class="hlt">observation</span> of wind, could reasonable derive zonal and meridional velocity vectors of EPBs drifts by tracking a series of successive <span class="hlt">airglow</span> 630.0 nm emission <span class="hlt">images</span>. <span class="hlt">Airglow</span> emission <span class="hlt">images</span> data are available from an all sky <span class="hlt">airglow</span> camera in Hainan Fuke (19.5°N, 109.2°E) supported by China Meridional Project, which can receive the 630.0nm emission from the ionosphere F region at low-latitudes to <span class="hlt">observe</span> plasma bubbles. A series of pretreatment technology, e.g. <span class="hlt">image</span> enhancement, orientation correction, <span class="hlt">image</span> projection are utilized to preprocess the raw <span class="hlt">observation</span>. Then the regions of plasma bubble extracted from the <span class="hlt">images</span> are divided into several small tracing windows and each tracing window can find a target window in the searching area in following <span class="hlt">image</span>, which is considered as the position tracing window moved to. According to this, velocities in each window are calculated by using the technology of cloud-derived wind. When applying the cloud-derived wind technology, the maximum correlation coefficient (MCC) and the histogram of gradient (HOG) methods to find the target window, which mean to find the maximum correlation and the minimum euclidean distance between two gradient histograms in respectively, are investigated and compared in detail. The maximum correlation method is fianlly adopted in this study to analyze the velocity of plasma bubbles because of its better performance than HOG. All-sky <span class="hlt">images</span> from Hainan Fuke, between August 2014 and October 2014, are analyzed to investigate the plasma bubble drift velocities using MCC method. The data at different local time at 9 nights are studied and find that zonal drift velocity in different latitude at different local time ranges from 50 m/s to 180 m/s and there is a peak value at</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JGRA..120.2222F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JGRA..120.2222F"><span>Geomagnetically conjugate <span class="hlt">observation</span> of plasma bubbles and thermospheric neutral winds 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.; Nishioka, M.; Kubota, M.; Tsugawa, T.; Nagatsuma, T.; Komonjinda, S.; Yatini, C. Y.</p> <p>2015-03-01</p> <p>This is the first paper that reports simultaneous <span class="hlt">observations</span> of zonal drift of plasma bubbles and the thermospheric neutral winds at geomagnetically conjugate points in both hemispheres. The plasma bubbles were <span class="hlt">observed</span> in the 630 nm nighttime <span class="hlt">airglow</span> <span class="hlt">images</span> taken by using highly sensitive all-sky <span class="hlt">airglow</span> <span class="hlt">imagers</span> at Kototabang, Indonesia (geomagnetic latitude (MLAT): 10.0°S), and Chiang Mai, Thailand (MLAT: 8.9°N), which are nearly geomagnetically conjugate stations, for 7 h from 13 to 20 UT (from 20 to 03 LT) on 5 April 2011. The bubbles continuously propagated eastward with velocities of 100-125 m/s. The 630 nm <span class="hlt">images</span> at Chiang Mai and those mapped to the conjugate point of Kototabang fit very well, which indicates that the <span class="hlt">observed</span> plasma bubbles were geomagnetically connected. The eastward thermospheric neutral winds measured by two Fabry-Perot interferometers were 70-130 m/s at Kototabang and 50-90 m/s at Chiang Mai. We compared the <span class="hlt">observed</span> plasma bubble drift velocity with the velocity calculated from the <span class="hlt">observed</span> neutral winds and the model conductivity, to investigate the F region dynamo contribution to the bubble drift velocity. The estimated drift velocities were 60-90% of the <span class="hlt">observed</span> velocities of the plasma bubbles, suggesting that most of the plasma bubble velocity can be explained by the F region dynamo effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730021619','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730021619"><span>Measurements of the Michigan <span class="hlt">Airglow</span> Observatory from 1971 to 1973 at Ester Dome Alaska</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mcwatters, K. D.; Meriwether, J. W.; Hays, P. B.; Nagy, A. F.</p> <p>1973-01-01</p> <p>The Michigan <span class="hlt">Airglow</span> Observatory (MAO) was located at Ester Dome Observatory, College, Alaska (latitude: 64 deg 53'N, longitude: 148 deg 03'W) since October, 1971. The MAO houses a 6-inch Fabry-Perot interferometer, a 2-channel monitoring photometer and a 4-channel tilting filter photometer. The Fabry-Perot interferometer was used extensively during the winter <span class="hlt">observing</span> seasons of 1971-72 and 1972-73 to measure temperature and mass motions of the neutral atmosphere above approximately 90 kilometers altitude. Neutral wind data from the 1971-72 <span class="hlt">observing</span> season as measured by <span class="hlt">observing</span> the Doppler shift of the gamma 6300 A atomic oxygen emission line are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..122..846H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..122..846H"><span>Numerical modeling of a multiscale gravity wave event and its <span class="hlt">airglow</span> signatures over Mount Cook, New Zealand, during the DEEPWAVE campaign</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Heale, C. J.; Bossert, K.; Snively, J. B.; Fritts, D. C.; Pautet, P.-D.; Taylor, M. J.</p> <p>2017-01-01</p> <p>A 2-D nonlinear compressible model is used to simulate a large-amplitude, multiscale mountain wave event over Mount Cook, NZ, <span class="hlt">observed</span> as part of the Deep Propagating Gravity Wave Experiment (DEEPWAVE) campaign and to investigate its <span class="hlt">observable</span> signatures in the hydroxyl (OH) layer. The campaign <span class="hlt">observed</span> the presence of a λx=200 km mountain wave as part of the 22nd research flight with amplitudes of >20 K in the upper stratosphere that decayed rapidly at <span class="hlt">airglow</span> heights. Advanced Mesospheric Temperature Mapper (AMTM) showed the presence of small-scale (25-28 km) waves within the warm phase of the large mountain wave. The simulation results show rapid breaking above 70 km altitude, with the preferential formation of almost-stationary vortical instabilities within the warm phase front of the mountain wave. An OH <span class="hlt">airglow</span> model is used to identify the presence of small-scale wave-like structures generated in situ by the breaking of the mountain wave that are consistent with those seen in the <span class="hlt">observations</span>. While it is easy to interpret these feature as waves in OH <span class="hlt">airglow</span> data, a considerable fraction of the features are in fact instabilities and vortex structures. Simulations suggest that a combination of a large westward perturbation velocity and shear, in combination with strong perturbation temperature gradients, causes both dynamic and convective instability conditions to be met particularly where the wave wind is maximized and the temperature gradient is simultaneously minimized. This leads to the inevitable breaking and subsequent generation of smaller-scale waves and instabilities which appear most prominent within the warm phase front of the mountain wave.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA21A2512K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA21A2512K"><span>Statistical study on the variations of OH and O2 rotational temperatures <span class="hlt">observed</span> by SATI at King Sejong Station (62.22S, 58.78W), Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, J.; Kim, J. H.; Jee, G.; Lee, C.; Kim, Y.</p> <p>2017-12-01</p> <p>Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span> (SATI) installed at King Sejong Station (62.22S, 58.78W), Antarctica, has been continuously measured the <span class="hlt">airglow</span> emissions from OH (6-2) Meinel and O2 (0-1) atmospheric bands since 2002, in order to investigate the dynamics of the polar MLT region. The measurements allow us to derive the rotational temperature at peak emission heights known as about 87 km and 94 km for OH and O2 <span class="hlt">airglows</span>, respectively. In this study, we briefly introduce improved analysis technique that modified original analysis code. The major change compared to original program is the improvement of the function to find the exact center position in the <span class="hlt">observed</span> <span class="hlt">image</span>. In addition to brief introduction of the improved technique, we also present the results statistically investigating the periodic variations on the temperatures of two layers during the period of 2002 through 2011 and compare our results with those from the temperatures measured by satellite.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.8129B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.8129B"><span>Network for the Detection of Mesopause Change (NDMC): What can we learn from <span class="hlt">airglow</span> measurements in terms of better understanding atmospheric dynamics?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bittner, Michael</p> <p>2013-04-01</p> <p>The international Network for the Detection of Mesopause Change (NDMC, http://wdc.dlr.de/ndmc) is a global program with the mission to promote international cooperation among research groups investigating the mesopause region (80-100 km) with the goal of early identification of changing climate signals. NDMC is contributing to the European Project "Atmospheric dynamics Research Infrastructure in Europe, ARISE". Measurements of the <span class="hlt">airglow</span> at the mesopause altitude region (80-100km) from most of the European NDMC stations including spectro-photometers and <span class="hlt">imagers</span> allow monitoring atmospheric variability at time scales comprising long-term trends, annual and seasonal variability, planetary and gravity waves and infrasonic signals. The measurements also allow validating satellite-based measurements such as from the TIMED-SABER instrument. Examples will be presented for <span class="hlt">airglow</span> measurements and for related atmospheric dynamics analysis on the abovementioned spatio-temporal scales and comparisons with satellite-based instruments as well as with LIDAR soundings in order to demonstrate the contribution of NDMC to the ARISE project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..12211794C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..12211794C"><span>The First Use of Coordinated Ionospheric Radio and Optical <span class="hlt">Observations</span> Over Italy: Convergence of High-and Low-Latitude Storm-Induced Effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cesaroni, C.; Alfonsi, L.; Pezzopane, M.; Martinis, C.; Baumgardner, J.; Wroten, J.; Mendillo, M.; Musicò, E.; Lazzarin, M.; Umbriaco, G.</p> <p>2017-11-01</p> <p>Ionospheric storm effects at midlatitudes were analyzed using different ground-based instruments distributed in Italy during the 13-15 November 2012 geomagnetic storm. These included an all-sky <span class="hlt">imager</span> (ASI) in Asiago (45.8°N, 11.5°E), a network of dual-frequeny Global Navigation Satellite Systems receivers (Rete Integrata Nazionale GPS network), and ionosondes in Rome (41.8°N, 12.5°E) and San Vito (40.6°N, 17.8°E). GPS measurements showed an unusual enhancement of total electron content (TEC) in southern Italy, during the nights of 14 and 15 November. The ASI <span class="hlt">observed</span> colocated enhancements of 630 nm <span class="hlt">airglow</span> at the same time, as did variations in NmF2 measured by the ionosondes. Moreover, wave-like perturbations were identified propagating from the north. The Ensemble Empirical Mode Decomposition, applied to TEC values revealed the presence of traveling ionospheric disturbances (TIDs) propagating southward between 01:30 UT and 03:00 UT on 15 November. These TIDs were characterized by weak TEC oscillations ( ±0.5 TEC unit), period of 45 min, and velocity of 500 m/s typical of large-scale TIDs. Optical <span class="hlt">images</span> showed enhanced <span class="hlt">airglow</span> entering the field of view of the ASI from the N-NE at 02:00 UT and propagating to the S-SW, reaching the region covered by the GPS stations after 03:00 UT, when TEC fluctuations are very small ( ±0.2 TEC unit). The enhancement of TEC and <span class="hlt">airglow</span> <span class="hlt">observed</span> in southern Italy could be a consequence of a poleward expansion of the northern crest of the equatorial ionization anomaly. The enhanced <span class="hlt">airglow</span> propagating from the north and the TEC waves resulted from energy injected at auroral latitudes as confirmed by magnetometer <span class="hlt">observations</span> in Scandinavia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850012171','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850012171"><span>Twilight Intensity Variation of the Infrared Hydroxyl <span class="hlt">Airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lowe, R. P.; Gilbert, K. L.; Niciejewski, R. J.</p> <p>1984-01-01</p> <p>The vibration rotation bands of the hydroxyl radical are the strongest features in the night <span class="hlt">airglow</span> and are exceeded in intensity in the dayglow only by the infrared atmospheric bands of oxygen. The variation of intensity during evening twilight is discussed. Using a ground-based Fourier Transform Spectrometer (FTS), hydroxyl intensity measurements as early as 3 deg solar depression were made. Models of the twilight behavior show that this should be sufficient to provide measurement of the main portion of the twilight intensity change. The instrument was equipped with a liquid nitrogen-cooled germanium detector whose high sensitivity combined with the efficiency of the FTS technique permits spectra of the region 1.1 to 1.6 microns at high signal-to-noise to be obtained in two minutes. The use of a polarizer at the entrance aperture of the instrument reduces the intensity of scattered sunlight by a factor of at least ten for zenith <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20165509','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20165509"><span>Silicon photodiode as a detector in the rocket-borne photometry of the near infrared <span class="hlt">airglow</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schaeffer, R C</p> <p>1976-11-01</p> <p>The application of a silicon P-I-N photodiode to the dc measurement of low levels of near ir radiation is described. It is shown that the threshold of signal detection is set by the current amplifier voltage noise, the effect of which at the output is determined by the value of source resistance of the photodiode. The photodiode was used as the detector in a compact interference filter photometer designed for rocket-borne studies of the <span class="hlt">airglow</span>. Flight results have proved the instrument's capability to provide measurements sufficiently precise to yield an accurate height profile of the (0-0) atmospheric band of O(2) night <span class="hlt">airglow</span> at lambda762 nm.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PASP..128i4504N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PASP..128i4504N"><span>Spatial and Temporal Stability of <span class="hlt">Airglow</span> Measured in the Meinel Band Window at 1191.3 nm</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nguyen, Hien T.; Zemcov, Michael; Battle, John; Bock, James J.; Hristov, Viktor; Korngut, Phillip; Meek, Andrew</p> <p>2016-09-01</p> <p>We report on the temporal and spatial fluctuations in the atmospheric brightness in the narrow band between Meinel emission lines at 1191.3 nm using a λ/Δλ = 320 near-infrared instrument. We present the instrument design and implementation, followed by a detailed analysis of data taken over the course of a night from Table Mountain Observatory. At low airmasses, the absolute sky brightness at this wavelength is found to be 5330 ± 30 nW m-2 sr-1, consistent with previous measurements of the inter-band <span class="hlt">airglow</span> at these wavelengths. This amplitude is larger than simple models of the continuum component of the <span class="hlt">airglow</span> emission at these wavelengths, confirming that an extra emissive or scattering component is required to explain the <span class="hlt">observations</span>. We perform a detailed investigation of the noise properties of the data and find no evidence for a noise component associated with temporal instability in the inter-line continuum. This result demonstrates that in several hours of ˜100 s integrations the noise performance of the instrument does not appear to significantly degrade from expectations, giving a proof of concept that near-infrared line intensity mapping may be feasible from ground-based sites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19740024671','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19740024671"><span>The global characteristics of atmosphere emissions in the lower thermosphere and their aeronomic implications. [OGO-4 <span class="hlt">airglow</span> photometric <span class="hlt">observations</span> of oxygen</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reed, E. I.; Chandra, S.</p> <p>1974-01-01</p> <p>The green line of atomic oxygen and the Herzberg bands of molecular oxygen as <span class="hlt">observed</span> from the OGO-4 <span class="hlt">airglow</span> photometer are discussed in terms of their spatial and temporal distributions and their relation to the atomic oxygen content in the lower thermosphere. Daily maps of the distribution of emissions show considerable structure (cells, patches, and bands) with appreciable daily changes. When data are averaged over periods of several days in length, the resulting patterns have occasional tendencies to follow geomagnetic parallels. The Seasonal variations are characterized by maxima in both the Northern and Southern Hemispheres in October, with the Northern Hemisphere having substantially higher emission rates. Formulae are derived relating the vertical column emission rates of the green line and the Herzberg bands to the atomic oxygen peak density. Global averages for the time period for these data (August 1967 to January 1968), when converted to maximum atomic oxygen densities near 95 km, have a range of 2.0 x 10 to the 11th power/cu cm 2.7 x 10 to the 11th power/cu cm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA12A..07C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA12A..07C"><span>Simulated GOLD <span class="hlt">Observations</span> of Atmospheric Waves</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Correira, J.; Evans, J. S.; Lumpe, J. D.; Rusch, D. W.; Chandran, A.; Eastes, R.; Codrescu, M.</p> <p>2016-12-01</p> <p>The Global-scale <span class="hlt">Observations</span> of the Limb and Disk (GOLD) mission will measure structures in the Earth's <span class="hlt">airglow</span> layer due to dynamical forcing by vertically and horizontally propagating waves. These measurements focus on global-scale structures, including compositional and temperature responses resulting from dynamical forcing. Daytime <span class="hlt">observations</span> of far-UV emissions by GOLD will be used to generate two-dimensional maps of the ratio of atomic oxygen and molecular nitrogen column densities (ΣO/N2 ) as well as neutral temperature that provide signatures of large-scale spatial structure. In this presentation, we use simulations to demonstrate GOLD's capability to deduce periodicities and spatial dimensions of large-scale waves from the spatial and temporal evolution <span class="hlt">observed</span> in composition and temperature maps. Our simulations include sophisticated forward modeling of the upper atmospheric <span class="hlt">airglow</span> that properly accounts for anisotropy in neutral and ion composition, temperature, and solar illumination. Neutral densities and temperatures used in the simulations are obtained from global circulation and climatology models that have been perturbed by propagating waves with a range of amplitudes, periods, and sources of excitation. Modeling of <span class="hlt">airglow</span> emission and predictions of ΣO/N2 and neutral temperatures are performed with the Atmospheric Ultraviolet Radiance Integrated Code (AURIC) and associated derived product algorithms. Predicted structure in ΣO/N2 and neutral temperature due to dynamical forcing by propagating waves is compared to existing <span class="hlt">observations</span>. Realistic GOLD Level 2 data products are generated from simulated <span class="hlt">airglow</span> emission using algorithm code that will be implemented operationally at the GOLD Science Data Center.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740051854&hterms=Dissociative&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DDissociative','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740051854&hterms=Dissociative&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DDissociative"><span>The 6300 A O/1-D/ <span class="hlt">airglow</span> and dissociative recombination</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wickwar, V. B.; Cogger, L. L.; Carlson, H. C.</p> <p>1974-01-01</p> <p>Measurements of night-time 6300 A <span class="hlt">airglow</span> intensities at the Arecibo Observatory have been compared with dissociative recombination calculations based on electron densities derived from simultaneous incoherent backscatter measurements. The agreement indicates that the nightglow can be fully accounted for by dissociative recombination. The comparisons are examined to determine the importance of quenching, heavy ions, ionization above the F-layer peak, and the temperature parameter of the model atmosphere. Comparable fits between the <span class="hlt">observed</span> and calculated intensities are found for several available model atmospheres. The least-squares fitting process, used to make the comparisons, produces comparable fits over a wide range of combinations of neutral densities and of reaction constants. Yet, the fitting places constraints upon the possible combinations; these constraints indicate that the latest laboratory chemical constants and densities extrapolated to a base altitude are mutually consistent.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRA..12111495S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRA..12111495S"><span>A statistical analysis of equatorial plasma bubble structures based on an all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span> network in China</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Longchang; Xu, Jiyao; Wang, Wenbin; Yuan, Wei; Li, Qinzeng; Jiang, Chaowei</p> <p>2016-11-01</p> <p>This paper investigates the statistical features of equatorial plasma bubbles (EPBs) using <span class="hlt">airglow</span> <span class="hlt">images</span> from 2012 to 2014 from a ground-based network of four <span class="hlt">imagers</span> in the equatorial region of China. It is found that (1) EPBs mainly occur during 21:00-00:00 local time (LT) in equinoxes. There is an asymmetry in occurrence rates between March (June) and September equinoxes (December solstices). (2) Most EPBs occur in groups of two to six depletions. The distance between adjacent EPB depletions is 100-700 km, and the average is 200-300 km. The zonal extension of an EPB group is usually less than 1500 km but can reach 3000 km. (3) EPBs usually have a maximum drift velocity near 100 m/s at 21:00-22:00 LT in 9.5° ± 1.5° geomagnetic latitude and then decrease to 50-70 m/s toward sunrise. (4) The averaged westward tilt angle of most EPBs (with respect to the geographic north-south) increased from 5°-10° to 23°-30° with LT between 20:00 and 03:00 LT, then decreasing to 10°-20° toward sunrise. (5) When 90 < F10.7 < 140, the maximum magnetic latitudinal extension (PMLE) is usually lower than 15.0° (apex height 725 km), but it can reach 23.0° (apex height 1330 km) when F10.7 > 140. The maximum PMLE increases by 3.4°-5.5° when F10.7 changes from 90 to 190. (6) The EPB occurrence patterns and zonal drift velocities are significantly different from those at Kolhapur, India, which locates west to our stations by 20.0°-32.0° in longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA31A2570B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA31A2570B"><span>Space Weather Now-Casting for Area-Denied Locations: Testing All-Sky-<span class="hlt">Imaging</span> Applications at Geomagnetic Conjugate Points.</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. L.; Mendillo, M.; Martinis, C. R.; Hickey, D. A.; Wroten, J.</p> <p>2017-12-01</p> <p>We explore the concept of using an all-sky-<span class="hlt">imager</span> (ASI) in one hemisphere to provide now-casting of ionospheric perturbations in the opposite hemisphere. The specific example deals with low-latitude plasma instabilities known as equatorial spread-F (ESF) that depend on geomagnetic field controlled electrodynamics. ASI <span class="hlt">observations</span> of 630.0 nm <span class="hlt">airglow</span> from 300 km exhibit regions of low emission ("<span class="hlt">airglow</span> depletions") that correlate highly with ESF patterns of radiowave disruptions, e.g., from GPS satellites. For both oceanographic and geopolitical reasons, there are vast regions of the globe that cannot be used for ground-based now-casting of local ESF effects. For such area-denied locations, it is possible for <span class="hlt">observations</span> of <span class="hlt">airglow</span> depletions from the opposite hemisphere to be used to specify both local and conjugate location environmental impacts. We use fifteen months of ASI <span class="hlt">observations</span> from the El Leoncito Observatory (Argentina) to predict simultaneous conditions at its trans-equatorial geomagnetic conjugate point in Villa de Leyva (Colombia)—validated by independent ASI <span class="hlt">observations</span> at that location. We find the success rate of conjugate point now-casting to be greater than 95% for large-scale ESF occurrence patterns. For a different pair of stations at higher magnetic latitudes, three years of <span class="hlt">observations</span> from the Arecibo Observatory (Puerto Rico) were used to make ESF now-casting at its conjugate point in Mercedes (Argentina) with a 85% success rate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870038348&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DAbreu%252C%2Bc.','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870038348&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DAbreu%252C%2Bc."><span>The quenching rate of O(1D) by O(3P). [with data from Visible <span class="hlt">Airglow</span> experiment on AE satellites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Abreu, V. J.; Yee, J. H.; Solomon, S. C.; Dalgarno, A.</p> <p>1986-01-01</p> <p>The rate coefficient for the quenching of O(1D) by O(3P) has recently been calculated by Yee et al. (1985). Their results indicate that quenching by atomic oxygen should not be ignored in the analysis of the 6300 A emission <span class="hlt">airglow</span>. Data obtained by the Visible <span class="hlt">Airglow</span> Experiment on board the AE satellites have been reanalyzed to determine the quenching rate of O(1D) by atomic oxygen. The results of this analysis are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JGRA..117.2317H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JGRA..117.2317H"><span>Estimating the electron energy distribution during ionospheric modification from spectrographic <span class="hlt">airglow</span> measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hysell, D. L.; Varney, R. H.; Vlasov, M. N.; Nossa, E.; Watkins, B.; Pedersen, T.; Huba, J. D.</p> <p>2012-02-01</p> <p>The electron energy distribution during an F region ionospheric modification experiment at the HAARP facility near Gakona, Alaska, is inferred from spectrographic <span class="hlt">airglow</span> emission data. Emission lines at 630.0, 557.7, and 844.6 nm are considered along with the absence of detectable emissions at 427.8 nm. Estimating the electron energy distribution function from the <span class="hlt">airglow</span> data is a problem in classical linear inverse theory. We describe an augmented version of the method of Backus and Gilbert which we use to invert the data. The method optimizes the model resolution, the precision of the mapping between the actual electron energy distribution and its estimate. Here, the method has also been augmented so as to limit the model prediction error. Model estimates of the suprathermal electron energy distribution versus energy and altitude are incorporated in the inverse problem formulation as representer functions. Our methodology indicates a heater-induced electron energy distribution with a broad peak near 5 eV that decreases approximately exponentially by 30 dB between 5-50 eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19730052459&hterms=patty&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpatty','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19730052459&hterms=patty&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dpatty"><span>Photometer for detection of sodium day <span class="hlt">airglow</span>.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mcmahon, D. J.; Manring, E. R.; Patty, R. R.</p> <p>1973-01-01</p> <p>Description of a photometer for daytime ground-based measurements of sodium <span class="hlt">airglow</span> emission. The photometer described can be characterized by the following principal features: (1) a narrow (4.5-A) interference filter for initial discrimination; (2) cooled photomultiplier detector to reduce noise from dark current fluctuations and chopping to eliminate the average dark current; (3) a sodium vapor resonance cell to provide an effective bandpass comparable to the Doppler line width; (4) separate detection of all light transmitted by the interference filter to evaluate the Rayleigh and Mie components within the Doppler width of the resonance cell; and (5) temperature quenching of the resonance cell to evaluate and account for instrumental imperfections.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss028e017123.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss028e017123.html"><span>Earth <span class="hlt">Observation</span></span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-07-15</p> <p>ISS028-E-017123 (16 July 2011) --- Separate atmospheric optical phenomena were captured in this electronic still photograph from the Inernational Space Station. The thin greenish band stretching along the Earth's horizon is <span class="hlt">airglow</span>; light emitted by the atmosphere from a layer about 30 kilometers thick and about 100 kilometers in altitude. The predominant emission in <span class="hlt">airglow</span> is the green 5577 Angstrom wavelength light from atomic oxygen atoms. <span class="hlt">Airglow</span> is always and everywhere present in the atmosphere; it results from the recombination of molecules that have been broken apart by solar radiation during the day. But <span class="hlt">airglow</span> is so faint that it can only be seen at night by looking "edge on" at the emission layer, such as the view astronauts and cosmonauts have in orbit. The second phenomenon is the appearnce of Aurora Australis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740044960&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740044960&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dtwilight"><span>Twilight <span class="hlt">airglow</span>. II - N2/+/ emission at 3914 A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sharp, W. E.</p> <p>1974-01-01</p> <p>One of the experiments aboard a rocket flight carrying instruments to measure the dawn <span class="hlt">airglow</span>, the ion and electron densities, and the photoelectron spectrum is reported. For a solar zenith angle of 90 deg the emission at 3914 A from N2(+) peaks at about 260 km. The integrated intensity from model calculations suggests that resonance scattering of 3914-A solar photons off N2(+) produces 90% of the emission, whereas simultaneous photoionization excitation of N2(+) produces less than 10% of the emission. Photoelectron impact excitation is found to contribute about 1%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMAE31B3408W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMAE31B3408W"><span>The D-Region Ledge at Nighttime: Why are Elves Collocated with the OH Meinel Band <span class="hlt">Airglow</span> Layer?</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Y. J.; Williams, E. R.; Hsu, R. R.</p> <p>2014-12-01</p> <p>The <span class="hlt">Imager</span> of Sprite and Upper Atmosphere Lightning (ISUAL) onboard the Taiwanese satellite Formosat-2 has continuously <span class="hlt">observed</span> transient luminous events (TLEs) within the +/-60 degree range of latitude for a decade since May 2004. The lightning electromagnetic pulse is responsible for Elves , the dominate TLE type which accounts for approximately 80% of the total TLE count according to the ISUAL global survey. By analyzing the limb-viewed <span class="hlt">images</span> with a wavelength filter of 622.8-754nm, 72% of elves are found to be 'glued' to the OH Meinel band (~630nm) nightglow layer within its thickness of 8km, with the OH layer normally at an altitude of 87 km (Huang et al., 2010).This collocation of elves and <span class="hlt">airglow</span> layer is frequently dismissed as coincidence, since the physical mechanisms for the formation of the two optical phenomena are macroscopically quite different. However, a common ingredient in the atmospheric chemistry is monatomic oxygen. O is needed to make O3 and ultimately hydroxyl OH, the main radiative species of the <span class="hlt">airglow</span> layer. O is also needed to form nitric oxide NO, the species with the lowest known ionization potential (9.26 eV) in the D-region. Thomas (1990) has documented steep increases in O concentration in the 83-85 km altitude range and Hale (1985) has found steep increases in electrical conductivity in the 84-85 km range, both with rocket measurements. A great simplification of the nighttime ionosphere is the presence of a single photon energy—10.2 eV—Lyman-α, originating in monatomic H in the Earth's geocorona. A simple Chapman layer calculation for the altitude of maximum photo-dissociation of O2, using the measured absorption cross-section of O2 at the Lyman-α energy, shows an altitude of maximum O production at 85 km. Elve emission in the nitrogen first positive band is enhanced by the presence of free electrons from ionized NO, but too large a conductivity will lead to the exclusion of the radiation field from the lightning return</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-s116e06796.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-s116e06796.html"><span>Earth <span class="hlt">Observations</span> taken by STS-116 Crewmember</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2006-12-18</p> <p>S116-E-06796 (17 Dec. 2006) --- A blanket of heavy cloud cover, <span class="hlt">airglow</span> and the blackness of space are featured in this <span class="hlt">image</span> photographed by a STS-116 crewmember on the International Space Station while Space Shuttle Discovery was docked with the station.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930061209&hterms=atmosphere+wind+profile&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Datmosphere%2Bwind%2Bprofile','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930061209&hterms=atmosphere+wind+profile&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Datmosphere%2Bwind%2Bprofile"><span>WINDII, the wind <span class="hlt">imaging</span> interferometer on the Upper Atmosphere Research Satellite</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shepherd, G. G.; Thuillier, G.; Gault, W. A.; Solheim, B. H.; Hersom, C.; Alunni, J. M.; Brun, J.-F.; Brune, S.; Charlot, P.; Cogger, L. L.</p> <p>1993-01-01</p> <p>The WIND <span class="hlt">imaging</span> interferometer (WINDII) was launched on the Upper Atmosphere Research Satellite (UARS) on September 12, 1991. This joint project, sponsored by the Canadian Space Agency and the French Centre National d'Etudes Spatiales, in collaboration with NASA, has the responsibility of measuring the global wind pattern at the top of the altitude range covered by UARS. WINDII measures wind, temperature, and emission rate over the altitude range 80 to 300 km by using the visible region <span class="hlt">airglow</span> emission from these altitudes as a target and employing optical Doppler interferometry to measure the small wavelength shifts of the narrow atomic and molecular <span class="hlt">airglow</span> emission lines induced by the bulk velocity of the atmosphere carrying the emitting species. The instrument used is an all-glass field-widened achromatically and thermally compensated phase-stepping Michelson interferometer, along with a bare CCD detector that <span class="hlt">images</span> the <span class="hlt">airglow</span> limb through the interferometer. A sequence of phase-stepped <span class="hlt">images</span> is processed to derive the wind velocity for two orthogonal view directions, yielding the vector horizontal wind. The process of data analysis, including the inversion of apparent quantities to vertical profiles, is described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA51B2390F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA51B2390F"><span>Ionospheric <span class="hlt">Observations</span> During a Geomagnetic Storm from LITES on the ISS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Finn, S. C.; Stephan, A. W.; Cook, T.; Budzien, S. A.; Chakrabarti, S.; Erickson, P. J.; Geddes, G.</p> <p>2017-12-01</p> <p>The Limb-<span class="hlt">Imaging</span> Ionospheric and Thermospheric Extreme-Ultraviolet Spectrograph (LITES) is an extreme-ultraviolet <span class="hlt">imaging</span> spectrograph that launched in February 2017 and was installed on the International Space Station (ISS). LITES is limb-viewing ( 150 - 350 km tangent altitude) and measures <span class="hlt">airglow</span> emissions from 60 - 140 nm with 0.2° angular and 1 nm spectral resolutions. We present early LITES results of <span class="hlt">observations</span> during a G2 geomagnetic storm in April 2017. In addition to LITES data, we will show complementary ground-based incoherent scatter radar (ISR) <span class="hlt">observations</span> from Millstone Hill during this storm. The combination of LITES EUV space-based <span class="hlt">observations</span> with the ground-based radio data is an example of the capability of campaign-style measurements of the ionosphere-thermosphere system using multiwavelength ground- and space-based instruments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA13B..09L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA13B..09L"><span>Comparison of rotational temperature derived from ground-based OH <span class="hlt">airglow</span> <span class="hlt">observations</span> with TIMED/SABER to evaluate the Einstein Coefficients</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, W.; Xu, J.; Smith, A. K.; Yuan, W.</p> <p>2017-12-01</p> <p>Ground-based <span class="hlt">observations</span> of the OH(9-4, 8-3, 6-2, 5-1, 3-0) band <span class="hlt">airglows</span> over Xinglong, China (40°24'N, 117°35'E) from December 2011 to 2014 are used to calculate rotational temperatures. The temperatures are calculated using five commonly used Einstein coefficient datasets. The kinetic temperature from TIMED/SABER is completely independent of the OH rotational temperature. SABER temperatures are weighted vertically by weighting functions calculated for each emitting vibrational state from two SABER OH volume emission rate profiles. By comparing the ground-based OH rotational temperature with SABER's, five Einstein coefficient datasets are evaluated. The results show that temporal variations of the rotational temperatures are well correlated with SABER's; the linear correlation coefficients are higher than 0.72, but the slopes of the fit between the SABER and rotational temperatures are not equal to 1. The rotational temperatures calculated using each set of Einstein coefficients produce a different bias with respect to SABER; these are evaluated over each of vibrational levels to assess the best match. It is concluded that rotational temperatures determined using any of the available Einstein coefficient datasets have systematic errors. However, of the five sets of coefficients, the rotational temperature derived with the Langhoff et al.'s (1986) set is most consistent with SABER. In order to get a set of optimal Einstein coefficients for rotational temperature derivation, we derive the relative values from ground-based OH spectra and SABER temperatures statistically using three year data. The use of a standard set of Einstein coefficients will be beneficial for comparing rotational temperatures <span class="hlt">observed</span> at different sites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-S100E5498.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-S100E5498.html"><span>A sunset Earth <span class="hlt">observation</span> <span class="hlt">image</span> taken during STS-100</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2001-04-26</p> <p>S100-E-5498 (26 April 2001) --- Earth's limb--the edge of the planet seen at twilight--was captured with a digital still camera by one of the STS-100 crew members aboard the Space Shuttle Endeavour. Near center frame the silhouette of cloud layers can be seen in the atmosphere, above which lies an <span class="hlt">airglow</span> layer (left).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EP%26S...70...88T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EP%26S...70...88T"><span>Initiation of a lightning search using the lightning and <span class="hlt">airglow</span> camera onboard the Venus orbiter Akatsuki</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takahashi, Yukihiro; Sato, Mitsuteru; Imai, Masataka; Lorenz, Ralph; Yair, Yoav; Aplin, Karen; Fischer, Georg; Nakamura, Masato; Ishii, Nobuaki; Abe, Takumi; Satoh, Takehiko; Imamura, Takeshi; Hirose, Chikako; Suzuki, Makoto; Hashimoto, George L.; Hirata, Naru; Yamazaki, Atsushi; Sato, Takao M.; Yamada, Manabu; Murakami, Shin-ya; Yamamoto, Yukio; Fukuhara, Tetsuya; Ogohara, Kazunori; Ando, Hiroki; Sugiyama, Ko-ichiro; Kashimura, Hiroki; Ohtsuki, Shoko</p> <p>2018-05-01</p> <p>The existence of lightning discharges in the Venus atmosphere has been controversial for more than 30 years, with many positive and negative reports published. The lightning and <span class="hlt">airglow</span> camera (LAC) onboard the Venus orbiter, Akatsuki, was designed to <span class="hlt">observe</span> the light curve of possible flashes at a sufficiently high sampling rate to discriminate lightning from other sources and can thereby perform a more definitive search for optical emissions. Akatsuki arrived at Venus during December 2016, 5 years following its launch. The initial operations of LAC through November 2016 have included a progressive increase in the high voltage applied to the avalanche photodiode detector. LAC began lightning survey <span class="hlt">observations</span> in December 2016. It was confirmed that the operational high voltage was achieved and that the triggering system functions correctly. LAC lightning search <span class="hlt">observations</span> are planned to continue for several years.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMSA41B1868R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMSA41B1868R"><span>Preliminary Analysis of <span class="hlt">Images</span> from the Thermospheric Temperature <span class="hlt">Imager</span> on Fast, Affordable, Science and Technology SATellite (FASTSAT)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rodriguez, M.; Jones, S.; Mentzell, E.; Gill, N.</p> <p>2011-12-01</p> <p>The Thermospheric Temperature <span class="hlt">Imager</span> (TTI) on Fast, Affordable, Science and Technology SATellite (FASTSAT) measures the upper atmospheric atomic oxygen emission at 135.6 nm and the molecular nitrogen LBH emission at 135.4 nm to determine the atmospheric O/N2 density ratio. <span class="hlt">Observations</span> of variations in this thermosheric ratio correspond to electron density variations in the ionosphere. The TTI design makes use of a Fabry-Perot interferometer to measure Doppler widened atmospheric emissions to determine neutral atmospheric temperature from low Earth orbit. FASTSAT launched November 10, 2010 and TTI is currently <span class="hlt">observing</span> geomagnetic signatures in the aurora and <span class="hlt">airglow</span>. This work is supported by NASA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110023311','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110023311"><span>Preliminary Analysis of <span class="hlt">Images</span> from the Thermospheric Temperature <span class="hlt">Image</span> on Fast, Affordable, Science and Technology Satellite (FASTSAT)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rodriquez, Marcello; Jones, Sarah; Mentzell, Eric; Gill, Nathaniel</p> <p>2011-01-01</p> <p>The Thermospheric Temperature <span class="hlt">Imager</span> (TTI) on Fast, Affordable, Science and Technology SATellite (FASTSAT) measures the upper atmospheric atomic oxygen emission at 135.6 nm and the molecular nitrogen LBH emission at 135.4 nm to determine the atmospheric O/N2 density ratio. <span class="hlt">Observations</span> of variations in this thermospheric ratio correspond to electron density variations in the ionosphere. The TTI design makes use of a Fabry-Perot interferometer to measure Doppler widened atmospheric emissions to determine neutral atmospheric temperature from low Earth orbit. FASTSAT launched November 10, 2010 and TTI is currently <span class="hlt">observing</span> geomagnetic signatures in the aurora and <span class="hlt">airglow</span>. This work is supported by NASA.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993IJRSP..22..197S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993IJRSP..22..197S"><span>All sky <span class="hlt">imaging</span> Fabry-Perot spectrometer for optical investigation of the upper atmosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sekar, R.; Gurubaran, S.; Sridharan, R.</p> <p>1993-06-01</p> <p>A simple optical design, keeping in view of the available components, has been worked out to develop the 'all sky <span class="hlt">imaging</span> Fabry-Perot spectrometer' to study the spatial structures in thermospheric winds and temperature. This system comprises three subsystems, namely, (1) field widening front-end optics, (2) high resolution Fabry-Perot spectrometer and (3) a two-dimensional detector. The design details of the above <span class="hlt">imaging</span> spectrometer that has been commissioned for routine <span class="hlt">observations</span> from Mt. Abu along with the first results on OI 6300 A <span class="hlt">airglow</span> emission are presented and discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..311F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..311F"><span>Case study of mesospheric front dissipation <span class="hlt">observed</span> over the northeast of Brazil</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fragoso Medeiros, Amauri; Paulino, Igo; Wrasse, Cristiano Max; Fechine, Joaquim; Takahashi, Hisao; Valentin Bageston, José; Paulino, Ana Roberta; Arlen Buriti, Ricardo</p> <p>2018-03-01</p> <p>On 3 October 2005 a mesospheric front was <span class="hlt">observed</span> over São João do Cariri (7.4° S, 36.5° W). This front propagated to the northeast and appeared in the <span class="hlt">airglow</span> <span class="hlt">images</span> on the west side of the observatory. By about 1.5 h later, it dissipated completely when the front crossed the local zenith. Ahead of the front, several ripple structures appeared during the dissipative process of the front. Using coincident temperature profile from the TIMED/SABER satellite and wind profiles from a meteor radar at São João do Cariri, the background of the atmosphere was investigated in detail. On the one hand, it was noted that a strong vertical wind shear in the propagation direction of the front produced by a semidiunal thermal tide was mainly responsible for the formation of duct (Doppler duct), in which the front propagated up to the zenith of the <span class="hlt">images</span>. On the other hand, the evolution of the Richardson number as well as the appearance of ripples ahead of the main front suggested that a presence of instability in the <span class="hlt">airglow</span> layer that did not allow the propagation of the front to the other side of the local zenith.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JASTP.130..151T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JASTP.130..151T"><span>Plasma bubble monitoring by TEC map and 630 nm <span class="hlt">airglow</span> <span class="hlt">image</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takahashi, H.; Wrasse, C. M.; Otsuka, Y.; Ivo, A.; Gomes, V.; Paulino, I.; Medeiros, A. F.; Denardini, C. M.; Sant'Anna, N.; Shiokawa, K.</p> <p>2015-08-01</p> <p>Equatorial ionosphere plasma bubbles over the South American continent were successfully <span class="hlt">observed</span> by mapping the total electron content (TECMAP) using data provided by ground-based GNSS receiver networks. The TECMAP could cover almost all of the continent within ~4000 km distance in longitude and latitude, monitoring TEC variability continuously with a time resolution of 10 min. Simultaneous <span class="hlt">observations</span> of OI 630 nm all-sky <span class="hlt">image</span> at Cachoeira Paulista (22.7°S, 45.0°W) and Cariri (7.4°S, 36.5°W) were used to compare the bubble structures. The spatial resolution of the TECMAP varied from 50 km to 1000 km, depending on the density of the <span class="hlt">observation</span> sites. On the other hand, optical <span class="hlt">imaging</span> has a spatial resolution better than 15 km, depicting the fine structure of the bubbles but covering a limited area (~1600 km diameter). TECMAP has an advantage in its spatial coverage and the continuous monitoring (day and night) form. The initial phase of plasma depletion in the post-sunset equatorial ionization anomaly (PS-EIA) trough region, followed by development of plasma bubbles in the crest region, could be monitored in a progressive way over the magnetic equator. In December 2013 to January 2014, periodically spaced bubble structures were frequently <span class="hlt">observed</span>. The longitudinal spacing between the bubbles was around 600-800 km depending on the day. The periodic form of plasma bubbles may suggest a seeding process related to the solar terminator passage in the ionosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EGUGA..1810082A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EGUGA..1810082A"><span>T he Analysis of the seasonal variations of equatorial plasma bubble, occurrence <span class="hlt">observed</span> from Oukaimeden Observatory, Morroco</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Amine, Lagheryeb; Zouhair, Benkhaldoun; Jonathan, Makela; Mohamed, Kaab; Aziza, Bounhir; Brian, Hardin; Dan, Fisher; Tmuthy, Duly</p> <p>2016-04-01</p> <p>T he Analysis of the seasonal variations of equatorial plasma bubble, occurrence using the 630.0 nm <span class="hlt">airglow</span> <span class="hlt">images</span> collected by the PICASSO <span class="hlt">imager</span> deployed at the Oukkaimden observatory in Morocco. Data have been taken since November 2013 to december 2015. We show the monthly average of appearance of EPBs. A maximum probability for bubble development is seen in the data in January and between late February and early March. We also <span class="hlt">observe</span> that there are a maximum period of appearance where the plasma is <span class="hlt">observed</span> (3-5 nights successivies) and we will discuss its connection with the solar activity in storm time. Future analysis will compare the probability of bubble occurrence in our site with the data raised in other <span class="hlt">observation</span> sites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018R%26QE..tmp...26S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018R%26QE..tmp...26S"><span>Spatial Characteristics of the 630-nm Artificial Ionospheric <span class="hlt">Airglow</span> Generation Region During the Sura Facility Pumping</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shindin, A. V.; Klimenko, V. V.; Kogogin, D. A.; Beletsky, A. B.; Grach, S. M.; Nasyrov, I. A.; Sergeev, E. N.</p> <p>2018-05-01</p> <p>We describe the method and the results of modeling and retrieval of the spatial distribution of excited oxygen atoms in the HF-pumped ionospheric region based on two-station records of artificial <span class="hlt">airglow</span> in the red line (λ = 630 nm). The HF ionospheric pumping was provided by the Sura facility. The red-line records of the night-sky portraits were obtained at two reception points—directly at the heating facility and 170 km east of it. The results were compared with the vertical ionospheric sounding data. It was found that in the course of the experiments the <span class="hlt">airglow</span> region was about 250 km high and did not depend on the altitude of the pump-wave resonance. The characteristic size of the region was 35 km, and the shape of the distribution isosurfaces was well described by oblique spheroids or a drop-shaped form. The average value of the maximum concentration of excited atoms during the experiment was about 1000 cm-3.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMSA51B2170F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMSA51B2170F"><span>Geomagnetic conjugate <span class="hlt">observations</span> of plasma bubbles and thermospheric neutral winds at equatorial 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.; Nishioka, M.; Kubota, M.; Tsugawa, T.; Nagatsuma, T.</p> <p>2012-12-01</p> <p>Plasma bubbles are plasma-density depletion which is developed by the Rayleigh-Taylor instability on the sunset terminator at equatorial latitudes. They usually propagate eastward after the sunset. The eastward propagation of the plasma bubbles is considered to be controlled by background eastward neutral winds in the thermosphere through the F-region dynamo effect. However, it is not clear how the F-region dynamo effect contributes to the propagation of the plasma bubbles, because plasma bubbles and background neutral winds have not been simultaneously <span class="hlt">observed</span> at geomagnetic conjugate points in the northern and southern hemispheres. In this study, geomagnetic conjugate <span class="hlt">observations</span> of the plasma bubbles at low latitudes with thermospheric neutral winds were reported. The plasma bubbles were <span class="hlt">observed</span> at Kototabang (0.2S, 100.3E, geomagnetic latitude (MLAT): 10.0S), Indonesia and at Chiang Mai (18.8N, 98.9E, MLAT: 8.9N), Thailand, which are geomagnetic conjugate stations, on 5 April, 2011 from 13 to 22 UT (from 20 to 05 LT). These plasma bubbles were <span class="hlt">observed</span> in the 630-nm <span class="hlt">airglow</span> <span class="hlt">images</span> taken by using highly-sensitive all-sky <span class="hlt">airglow</span> <span class="hlt">imagers</span> at both stations. They propagated eastward with horizontal velocities of about 100-125 m/s. Background thermospheric neutral winds were also <span class="hlt">observed</span> at both stations by using two Fabry-Perot interferometers (FPIs). The eastward wind velocities were about 70-130 m/s at Kototabang, and about 50-90 m/s at Chiang Mai. We estimated ion drift velocities by using these neutral winds <span class="hlt">observed</span> by FPIs and conductivities calculated from the IRI and MSIS models. The estimated velocities were about 60-90 % of the drift velocities of plasma bubbles. This result shows that most of the plasma bubble drift can be explained by the F-region dynamo effect, and additional electric field effect may come in to play.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMED33C0775B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMED33C0775B"><span>All Sky <span class="hlt">Imager</span> Network for Science and Education</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhatt, A.; Kendall, E. A.; Zalles, D. R.; Baumgardner, J. L.; Marshall, R. A.; Kaltenbacher, E.</p> <p>2012-12-01</p> <p>A new all sky <span class="hlt">imager</span> network for space weather monitoring and education outreach has been developed by SRI International. The goal of this program is to install sensitive, low-light all-sky <span class="hlt">imagers</span> across the continental United States to <span class="hlt">observe</span> upper atmospheric <span class="hlt">airglow</span> and aurora in near real time. While aurora borealis is often associated with the high latitudes, during intense geomagnetic storms it can extend well into the continental United States latitudes. <span class="hlt">Observing</span> auroral processes is instrumental in understanding the space weather, especially in the times of increasing societal dependence on space-based technologies. Under the THEMIS satellite program, Canada has installed a network of all-sky <span class="hlt">imagers</span> across their country to monitor aurora in real-time. However, no comparable effort exists in the United States. Knowledge of the aurora and <span class="hlt">airglow</span> across the entire United States in near real time would allow scientists to quickly assess the impact of a geomagnetic storm in concert with data from GPS networks, ionosondes, radars, and magnetometers. What makes this effort unique is that we intend to deploy these <span class="hlt">imagers</span> at high schools across the country. Selected high-schools will necessarily be in rural areas as the instrument requires dark night skies. At the commencement of the school year, we plan to give an introductory seminar on space weather at each of these schools. Science nuggets developed by SRI International in collaboration with the Center for GeoSpace Studies and the Center for Technology in Learning will be available for high school teachers to use during their science classes. Teachers can use these nuggets as desired within their own curricula. We intend to develop a comprehensive web-based interface that will be available for students and scientific community alike to <span class="hlt">observe</span> data across the network in near real time and also to guide students towards complementary space weather data sets. This interface will show the real time extent of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA33B..05G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA33B..05G"><span>Anticipated <span class="hlt">Observation</span> of Waves and Tides by the GOLD Mission Using a GCM and GLOW model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Greer, K.; Solomon, S. C.; Rusch, D. W.</p> <p>2017-12-01</p> <p>One of the major scientific objectives of the GOLD mission is to address the significance of atmospheric waves and tides propagating from below on the thermospheric temperature structure. Here we examine the modes of tides and spectrum of waves that will be <span class="hlt">observed</span> by GOLD in geostationary orbit. The GOLD instrument is an <span class="hlt">imaging</span> spectrograph that will measure the Earth's emissions from 132 to 162 nm. These measurements will be used to <span class="hlt">image</span> thermospheric temperature and composition near 160 km on the dayside disk at half-hour time scales. TIE-GCM is used to produce a realistic model atmosphere, where different wave and tidal components can be easily extracted, and GLobal <span class="hlt">AirglOW</span> (GLOW) model produces the emissions in the spectral bands <span class="hlt">observed</span> by GOLD.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080021265','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080021265"><span>A Multi-Instrument Measurement of a Mesospheric Bore at the Equator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shiokawa, K.; Suzuki, S.; Otsuka, Y.; Ogawa, T.; Nakamura, T.; Mlynczak, M. G.; Russell, J. M., III</p> <p>2005-01-01</p> <p>We have made a comprehensive measurement of mesospheric bore phenomenon at the equator at Kototabang, Indonesia (0.2 deg S, 100.3 deg E), using an <span class="hlt">airglow</span> <span class="hlt">imager</span>, an <span class="hlt">airglow</span> temperature photometer, a meteor radar, and the SABER instrument on board the TIMED satellite. The bore was detected in <span class="hlt">airglow</span> <span class="hlt">images</span> of both OH-band (peak emission altitude: 87 km) and 557.7-nm (96 km) emissions, as east-west front-like structure propagating northward with a velocity of 52-58 m/s. Wave trains with a horizontal wavelength of 30-70 km are <span class="hlt">observed</span> behind the bore front. The <span class="hlt">airglow</span> intensity decreases for all the mesospheric emissions of OI (557.7 nm), OH-band, O2-band (altitude: 94 km), and Na (589.3 nm) (90 km) after the bore passage. The rotational temperatures of both OH-band and O2-band also decrease approximately 10 K after the bore passage. An intense shear in northward wind velocity of 80m/s was <span class="hlt">observed</span> at altitudes of 84-90 km by the meteor radar. Kinetic temperature profile at altitudes of 20-120 km was <span class="hlt">observed</span> near Kototabang by TIMED/SABER. On the basis of these <span class="hlt">observations</span>, we discuss generation and ducting of the <span class="hlt">observed</span> mesospheric bore.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.4243D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.4243D"><span><span class="hlt">Observations</span> and Operational Products from the Special Sensor Ultraviolet Limb <span class="hlt">Imager</span> (SSULI)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dandenault, Patrick; Nicholas, Andrew C.; Coker, Clayton; Budzien, Scott A.; Chua, Damien H.; Finne, Ted T.; Metzler, Christopher A.; Dymond, Kenneth F.</p> <p></p> <p>The Naval Research Laboratory (NRL) has developed five ultraviolet remote sensing instru-ments for the Air Force Defense Meteorological Satellite Program (DMSP). These instruments known as SSULI (Special Sensor Ultraviolet Limb <span class="hlt">Imager</span>) are on the DMSP block of 5D3 satellites, which first launched in 2003. The DMSP satellites are launched in a near-polar, sun-synchronous orbit at an altitude of approximately 830 km. SSULI measures vertical profiles of the natural <span class="hlt">airglow</span> radiation from atoms, molecules and ions in the upper atmosphere and ionosphere by viewing the earth's limb at a tangent altitude of approximately 50 km to 750 km. Limb <span class="hlt">observations</span> are made from the extreme ultraviolet (EUV) to the far ultraviolet (FUV) over the wavelength range of 80 nm to 170 nm, with 1.8 nm resolution. An extensive operational data processing system, the SSULI Ground Data Analysis Software (GDAS), has been developed to generate environmental data products from SSULI spectral data in near-real time for use at the Air Force Weather Agency (AFWA). The operational software uses advanced science algorithms developed at NRL and was designed to calibrate data from USAF Raw Sensor Data Records (RSDR) and generate Environmental Data Records (EDRs). Data products from SSULI <span class="hlt">observations</span> include vertical profiles of electron (Ne) densities, N2, O2, O, O+, Temperature and also vertical Total Electron Content (TEC). On October 18, 2009, the third SSULI sensor launched from Vandenberg Air Force Base, aboard the DMSP F18 spacecraft. An overview of the SSULI operational program and the status of the F18 sensor will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000082009&hterms=solar+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dsolar%2Benergy','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000082009&hterms=solar+energy&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dsolar%2Benergy"><span>Solar Energy Deposition Rates in the Mesosphere Derived from <span class="hlt">Airglow</span> Measurements: Implications for the Ozone Model Deficit Problem</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mlynczak, Martin G.; Garcia, Rolando R.; Roble, Raymond G.; Hagan, Maura</p> <p>2000-01-01</p> <p>We derive rates of energy deposition in the mesosphere due to the absorption of solar ultraviolet radiation by ozone. The rates are derived directly from measurements of the 1.27-microns oxygen dayglow emission, independent of knowledge of the ozone abundance, the ozone absorption cross sections, and the ultraviolet solar irradiance in the ozone Hartley band. Fifty-six months of <span class="hlt">airglow</span> data taken between 1982 and 1986 by the near-infrared spectrometer on the Solar-Mesosphere Explorer satellite are analyzed. The energy deposition rates exhibit altitude-dependent annual and semi-annual variations. We also find a positive correlation between temperatures and energy deposition rates near 90 km at low latitudes. This correlation is largely due to the semiannual oscillation in temperature and ozone and is consistent with model calculations. There is also a suggestion of possible tidal enhancement of this correlation based on recent theoretical and <span class="hlt">observational</span> analyses. The <span class="hlt">airglow</span>-derived rates of energy deposition are then compared with those computed by multidimensional numerical models. The <span class="hlt">observed</span> and modeled deposition rates typically agree to within 20%. This agreement in energy deposition rates implies the same agreement exists between measured and modeled ozone volume mixing ratios in the mesosphere. Only in the upper mesosphere at midlatitudes during winter do we derive energy deposition rates (and hence ozone mixing ratios) consistently and significantly larger than the model calculations. This result is contrary to previous studies that have shown a large model deficit in the ozone abundance throughout the mesosphere. The climatology of solar energy deposition and heating presented in this paper is available to the community at the Middle Atmosphere Energy Budget Project web site at http://heat-budget.gats-inc.com.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19910019905','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19910019905"><span>Planetary instrument definition and development program: 'Miniature Monochromatic <span class="hlt">Imager</span>'</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Broadfoot, A. L.</p> <p>1991-01-01</p> <p>The miniature monochromatic <span class="hlt">imager</span> (MMI) development work became the basis for the preparation of several instruments which were built and flown on the shuttle STS-39 as well as being used in ground based experiments. The following subject areas are covered: (1) applications of the ICCD to <span class="hlt">airglow</span> and auroral measurements and (2) a panchromatic spectrograph with supporting monochromatic <span class="hlt">imagers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014cosp...40E3297T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014cosp...40E3297T"><span>Capability of simultaneous Rayleigh LiDAR and O2 <span class="hlt">airglow</span> measurements in exploring the short period wave characteristics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Taori, Alok; Raghunath, Karnam; Jayaraman, Achuthan</p> <p></p> <p>We use combination of simultaneous measurements made with Rayleigh lidar and O2 <span class="hlt">airglow</span> monitoring to improve lidar investigation capability to cover a higher altitude range. We feed instantaneous O2 <span class="hlt">airglow</span> temperatures instead the model values at the top altitude for subsequent integration method of temperature retrieval using Rayleigh lidar back scattered signals. Using this method, errors in the lidar temperature estimates converges at higher altitudes indicating better altitude coverage compared to regular methods where model temperatures are used instead of real-time measurements. This improvement enables the measurements of short period waves at upper mesospheric altitudes (~90 km). With two case studies, we show that above 60 km the few short period wave amplitude drastically increases while, some of the short period wave show either damping or saturation. We claim that by using such combined measurements, a significant and cost effective progress can be made in the understanding of short period wave processes which are important for the coupling across the different atmospheric regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012R%26QE...55...33G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012R%26QE...55...33G"><span><span class="hlt">Airglow</span> during ionospheric modifications by the sura facility radiation. experimental results obtained in 2010</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grach, S. M.; Klimenko, V. V.; Shindin, A. V.; Nasyrov, I. A.; Sergeev, E. N.; A. Yashnov, V.; A. Pogorelko, N.</p> <p>2012-06-01</p> <p>We present the results of studying the structure and dynamics of the HF-heated volume above the Sura facility obtained in 2010 by measurements of ionospheric <span class="hlt">airglow</span> in the red (λ = 630 nm) and green (λ = 557.7 nm) lines of atomic oxygen. Vertical sounding of the ionosphere (followed by modeling of the pump-wave propagation) and measurements of stimulated electromagnetic emission were used for additional diagnostics of ionospheric parameters and the processes occurring in the heated volume.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA32A..02S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA32A..02S"><span>Longitudinal Ionospheric Variability <span class="hlt">Observed</span> by LITES on the ISS</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stephan, A. W.; Finn, S. C.; Cook, T.; Geddes, G.; Chakrabarti, S.; Budzien, S. A.</p> <p>2017-12-01</p> <p>The Limb-<span class="hlt">Imaging</span> Ionospheric and Thermospheric Extreme-Ultraviolet Spectrograph (LITES) is an <span class="hlt">imaging</span> spectrograph designed to measure altitude profiles (150-350 km) of extreme- and far-ultraviolet <span class="hlt">airglow</span> emissions that originate from photochemical processes in the ionosphere and thermosphere. During the daytime, LITES <span class="hlt">observes</span> the bright O+ 83.4 nm emission from which the ionospheric profile can be inferred. At night, recombination emissions at 91.1 and 135.6 nm provide a direct measure of the electron content along the line of sight. LITES was launched and installed on the International Space Station (ISS) in late February 2017 where it has been operating along with the highly complementary GPS Radio Occultation and Ultraviolet Photometry - Colocated (GROUP-C) experiment. We will present some of the first <span class="hlt">observations</span> from LITES in April 2017 that show longitudinal patterns in ionospheric density and the daily variability in those patterns. LITES vertical <span class="hlt">imaging</span> from a vantage point near 410 km enables a particularly unique perspective on the altitude of the ionospheric peak density at night that can complement and inform other ground- and space-based measurements, and track the longitude-altitude variability that is reflective of changes in equatorial electrodynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSA31B2344B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSA31B2344B"><span>Auroral LSTIDs and SAR Arc Occurrences in Northern California During Geomagnetic Storms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhatt, A.; Kendall, E. A.</p> <p>2015-12-01</p> <p>A 630nm allsky <span class="hlt">imager</span> has been operated for two years in northern California at the Hat Creek Radio Observatory. F-region <span class="hlt">airglow</span> data captured by the <span class="hlt">imager</span> ranges from approximately L=1.7 -2.7. Since installation of the <span class="hlt">imager</span> several geomagnetic storms have occurred with varying intensities. Two main manifestations of the geomagnetic storms are <span class="hlt">observed</span> in the 630 nm <span class="hlt">airglow</span> data: large-scale traveling ionospheric disturbances that are launched from the auroral zone and Stable Auroral Red (SAR) arcs during more intense geomagnetic storms. We will present a statistical analysis of these storm-time phenomena in northern California for the past eighteen months. This <span class="hlt">imager</span> is part of a larger all-sky <span class="hlt">imaging</span> network across the continental United States, termed MANGO (Midlatitude All-sky-<span class="hlt">imaging</span> Network for Geophysical <span class="hlt">Observations</span>). Where available, we will add data from networked <span class="hlt">imagers</span> located at similar L-shell in other states as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10256E..4FR','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10256E..4FR"><span>Simulation of the fixed optical path difference of near infrared wind <span class="hlt">imaging</span> interferometer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rong, Piao; Zhang, Chunmin; Yan, Tingyu; Liu, Dongdong; Li, Yanfen</p> <p>2017-02-01</p> <p>As an important part of the earth, atmosphere plays a vital role in filtering the solar radiation, adjusting the temperature and organizing the water circulation and keeping human survival. The passive atmospheric wind measurement is based on the <span class="hlt">imaging</span> interferometer technology and Doppler effect of electromagnetic wave. By using the wind <span class="hlt">imaging</span> interferometer to get four interferograms of <span class="hlt">airglow</span> emission lines, the atmospheric wind velocity, temperature, pressure and emission rate can be derived. Exploring the multi-functional and integrated innovation of detecting wind temperature, wind velocity and trace gas has become a research focus in the field. In the present paper, the impact factors of the fixed optical path difference(OPD) of near infrared wind <span class="hlt">imaging</span> interferometer(NIWII) are analyzed and the optimum value of the fixed optical path difference is simulated, yielding the optimal results of the fixed optical path difference is 20 cm in near infrared wave band (the O2(a1Δg) <span class="hlt">airglow</span> emission at 1.27 microns). This study aims at providing theoretical basis and technical support for the detection of stratosphere near infrared wind field and giving guidance for the design and development of near infrared wind <span class="hlt">imaging</span> interferometer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19810054575&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810054575&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight"><span><span class="hlt">Observations</span> of neutral iron emission in twilight spectra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tepley, C. A.; Meriwether, J. W., Jr.; Walker, J. C. G.; Mathews, J. D.</p> <p>1981-01-01</p> <p>A method is presented for the analysis of twilight <span class="hlt">airglow</span> spectra that may be contaminated by atmospheric continuum emission of unknown brightness. The necessity of correcting for this continuum emission when measuring weak <span class="hlt">airglow</span> features in twilight is illustrated by application of the method to the neutral iron line at 3860 A.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P23D2773C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P23D2773C"><span>Improving MAVEN-IUVS Lyman-Alpha Apoapsis <span class="hlt">Images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chaffin, M.; AlMannaei, A. S.; Jain, S.; Chaufray, J. Y.; Deighan, J.; Schneider, N. M.; Thiemann, E.; Mayyasi, M.; Clarke, J. T.; Crismani, M. M. J.; Stiepen, A.; Montmessin, F.; Epavier, F.; McClintock, B.; Stewart, I. F.; Holsclaw, G.; Jakosky, B. M.</p> <p>2017-12-01</p> <p>In 2013, the Mars Atmosphere and Volatile EvolutioN (MAVEN) mission was launched to study the Martian upper atmosphere and ionosphere. MAVEN orbits through a very thin cloud of hydrogen gas, known as the hydrogen corona, that has been used to explore the planet's geologic evolution by detecting the loss of hydrogen from the atmosphere. Here we present various methods of extracting properties of the hydrogen corona from <span class="hlt">observations</span> using MAVEN's <span class="hlt">Imaging</span> Ultraviolet Spectograph (IUVS) instrument. The analysis presented here uses the IUVS Far Ultraviolet mode apoapase data. From apoapse, IUVS is able to obtain <span class="hlt">images</span> of the hydrogen corona by detecting the Lyman-alpha <span class="hlt">airglow</span> using a combination of instrument scan mirror and spacecraft motion. To complete one apoapse <span class="hlt">observation</span>, eight scan swaths are performed to collect the <span class="hlt">observations</span> and construct a coronal <span class="hlt">image</span>. However, these <span class="hlt">images</span> require further processing to account for the atmospheric MUV background that hinders the quality of the data. Here, we present new techniques for correcting instrument data. For the background subtraction, a multi-linear regression (MLR) routine of the first order MUV radiance was used to improve the <span class="hlt">images</span>. A flat field correction was also applied by fitting a polynomial to periapse radiance <span class="hlt">observations</span>. The apoapse data was re-binned using this fit.The results are presented as <span class="hlt">images</span> to demonstrate the improvements in the data reduction. Implementing these methods for more orbits will improve our understanding of seasonal variability and H loss. Asymmetries in the Martian hydrogen corona can also be assessed to improve current model estimates of coronal H in the Martian atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSM43A2474Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSM43A2474Z"><span>Mesoscale Magnetosphere-Ionosphere Coupling along Open Magnetic Field Lines Associated with <span class="hlt">Airglow</span> Patches: Field-aligned Currents and Precipitation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zou, Y.; Nishimura, Y.; Lyons, L. R.; Shiokawa, K.; Burchill, J. K.; Knudsen, D. J.; Buchert, S. C.; Chen, S.; Nicolls, M. J.; Ruohoniemi, J. M.; McWilliams, K. A.; Nishitani, N.</p> <p>2016-12-01</p> <p>Although <span class="hlt">airglow</span> patches are traditionally regarded as high-density plasma unrelated to local field-aligned currents (FACs) and precipitation, past <span class="hlt">observations</span> were limited to storm-time conditions. Recent non-storm time <span class="hlt">observations</span> show patches to be associated with azimuthally narrow ionospheric fast flow channels that substantially contribute to plasma transportation across the polar cap and connect dayside and nightside explosive disturbances. We examine whether non-storm time patches are related also to localized polar cap FACs and precipitation using Swarm- and FAST-<span class="hlt">imager</span>-radar conjunctions. In Swarm data, we commonly (66%) identify substantial magnetic perturbations indicating FAC enhancements around patches. These FACs have substantial densities (0.1-0.2 μA/m-2) and can be approximated as infinite current sheets (typically 75 km wide) orientated roughly parallel to patches. They usually exhibit a Region-1 sense, i.e. a downward FAC lying eastward of an upward FAC, and can close through Pedersen currents in the ionosphere, implying that the locally enhanced dawn-dusk electric field across the patch is imposed by processes in the magnetosphere. In FAST data, we identify localized precipitation that is enhanced within patches in comparison to weak polar rain outside patches. The precipitation consists of structured or diffuse soft electron fluxes. While the latter resembles polar rain only with higher fluxes, the former consists of discrete fluxes enhanced by 1-2 orders of magnitude from several to several hundred eV. Although the precipitation is not a major contributor to patch ionization, it implies that newly reconnected flux tubes that retain electrons of magnetosheath origin can rapidly traverse the polar cap from the dayside. Therefore non-storm time patches should be regarded as part of a localized magnetosphere-ionosphere coupling system along open magnetic field lines, and their transpolar evolution as a reflection of reconnected flux tubes</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('https://ntrs.nasa.gov/search.jsp?R=19950058911&hterms=twilight&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950058911&hterms=twilight&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtwilight"><span>Sensitivity of the 6300 A twilight <span class="hlt">airglow</span> to neutral composition</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Melendez-Alvira, D. J.; Torr, D. G.; Richards, P. G.; Swift, W. R.; Torr, M. R.; Baldridge, T.; Rassoul, H.</p> <p>1995-01-01</p> <p>The field line interhemispheric plasma (FLIP) model is used to study the 6300 A line intensity measured during three morning twilights from the McDonald Observatory in Texas. The <span class="hlt">Imaging</span> Spectrometric Observatory (ISO) measured the 6300 A intensity during the winter of 1987 and the spring and summer of 1988. The FLIP model reproduces the measured intensity and its variation through the twilight well on each day using neutral densities from the MSIS-86 empirical model. This is in spite of the fact that different component sources dominate the integrated volume emission rate on each of the days analyzed. The sensitivity of the intensity to neutral composition is computed by varying the N2, O2, and O densities in the FLIP model and comparing to the intensity computed with the unmodified MSIS-86 densities. The ion densities change self-consistently. Thus the change in neutral composition also changes the electron density. The F2 peak height is unchanged in the model runs for a given day. The intensity changes near 100 deg SZA are comparable to within 10% when either (O2), (N2), or (O) is changed, regardless of which component source is dominant. There is strong sensitivity to changes in (N2) when dissociative recombination is dominant, virtually no change in the nighttime (SZA greater than or equal to 108 deg) intensity with (O2) doubled, and sensitivity of over 50% to doubling or halving (O) at night. When excitation by conjugate photoelectrons is the dominant nighttime component source, the relative intensity change with (O) doubled or halved is very small. This study shows the strong need for simultaneous measurements of electron density and of emissions proportional to photoelectron fluxes if the 6300 A twilight <span class="hlt">airglow</span> is to be used to retrieve neutral densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001EP%26S...53..741K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001EP%26S...53..741K"><span>Characteristics of medium- and large-scale TIDs over Japan derived from OI 630-nm nightglow <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kubota, M.; Fukunishi, H.; Okano, S.</p> <p>2001-07-01</p> <p>A new optical instrument for studying upper atmospheric dynamics, called the Multicolor All-sky <span class="hlt">Imaging</span> System (MAIS), has been developed. The MAIS can obtain all-sky <span class="hlt">images</span> of <span class="hlt">airglow</span> emission at two different wavelengths simultaneously with a time resolution of several minutes. Since December 1991, <span class="hlt">imaging</span> <span class="hlt">observations</span> with the MAIS have been conducted at the Zao observatory (38.09°N, 140.56°E). From these <span class="hlt">observations</span>, two interesting events with wave structures have been detected in OI 630-nm nightglow <span class="hlt">images</span>. The first event was <span class="hlt">observed</span> on the night of June 2/3, 1992 during a geomagnetically quiet period. Simultaneous data of ionospheric parameters showed that they are caused by propagation of the medium-scale traveling ionospheric disturbance (TID). Phase velocity and horizontal wavelength determined from the <span class="hlt">image</span> data are 45-100 m/s and ~280 km, and the propagation direction is south-westward. The second event was <span class="hlt">observed</span> on the night of February 27/28, 1992 during a geomagnetic storm. It is found that a large enhancement of OI 630-nm emission is caused by a propagation of the large-scale TID. Meridional components of phase velocities and wavelengths determined from ionospheric data are 305-695 m/s (southward) and 930-5250 km. The source of this large-scale TID appears to be auroral processes at high latitudes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19920014282','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19920014282"><span>WAMDII: The Wide Angle Michelson Doppler <span class="hlt">Imaging</span> Interferometer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1992-01-01</p> <p>As part of an effort to learn more about the upper atmosphere and how it is linked to the weather experienced each day, NASA and NRCC are jointly sponsoring the Wide Angle Michelson Doppler <span class="hlt">Imaging</span> Interferometer (WAMDII) Mission. WAMDII will measure atmospheric temperature and wind speed in the upper atmosphere. In addition to providing data on the upper atmosphere, the wind speed and temperature readings WAMDII takes will also be highly useful in developing and updating computer simulated models of the upper atmosphere. These models are used in the design and testing of equipment and software for Shuttles, satellites, and reentry vehicles. In making its wind speed and temperature measurements, WAMDII examines the Earth's <span class="hlt">airglow</span>, a faint photochemical luminescence caused by the influx of solar ultraviolet energy into the upper atmosphere. During periods of high solar flare activity, the amount of this UV energy entering the upper atmosphere increases, and this increase may effect <span class="hlt">airglow</span> emissions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss030e064161.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss030e064161.html"><span>Earth <span class="hlt">Observations</span> taken by Expedition 30 crewmember</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2012-02-02</p> <p>ISS030-E-064161 (2 Feb. 2012) --- Parts of a number of European nations appear in this nighttime <span class="hlt">image</span> photographed from the International Space Station. The scene, captured by one of the Expedition 30 crew members, shows the British Isles (left, partially obstructed by one of the space station's solar array panels) with London just right of bottom center; the English Channel, which is dark; Paris (lower right corner); and the Netherlands (right side). The greenish <span class="hlt">airglow</span> is fairly uniform and minor until it transitions to daybreak on the right.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997JGR...10219949Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997JGR...10219949Y"><span>Global simulations and <span class="hlt">observations</span> of O(1S), O2(1Σ) and OH mesospheric nightglow emissions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yee, Jeng-Hwa; Crowley, G.; Roble, R. G.; Skinner, W. R.; Burrage, M. D.; Hays, P. B.</p> <p>1997-09-01</p> <p>Despite a large number of <span class="hlt">observations</span> of mesospheric nightglow emissions in the past, the quantitative comparison between theoretical and experimental brightnesses is rather poor, owing primarily to the short duration of the <span class="hlt">observations</span>, the strong variability of the tides, and the influence of short-timescale gravity waves. The high-resolution Doppler <span class="hlt">imager</span> (HRDI) instrument onboard the upper atmosphere research satellite (UARS) provides nearly simultaneous, near-global <span class="hlt">observations</span> of O(1S) green line, O2(0-1) atmospheric band, and OH Meinel band nightglow emissions. Three days of these <span class="hlt">observations</span> near the September equinox of 1993 are presented to show the general characteristics of the three emissions, including the emission brightness, peak emission altitude, and their temporal and spatial variabilities. The global distribution of these emissions is simulated on the basis of atmospheric parameters from the recently developed National Center for Atmospheric Research (NCAR) thermosphere-ionosphere-mesosphere-electrodynamics general circulation model (TIME-GCM). The most striking features revealed by the global simulation are the structuring of the mesospheric nightglow by the diurnal tides and enhancements of the <span class="hlt">airglow</span> at high latitudes. The model reproduces the inverse relationship <span class="hlt">observed</span> by HRDI between the nightglow brightness and peak emission altitude. Analysis of our model results shows that the large-scale latitudinal/tidal nightglow brightness variations are a direct result of a complex interplay between mesospheric and lower thermospheric diffusive and advective processes, acting mainly on the atomic oxygen concentrations. The inclination of the UARS spacecraft precluded <span class="hlt">observations</span> of high latitude nightglow emissions by HRDI. However, our predicted high-latitude brightness enhancements confirm previous limited groundbased <span class="hlt">observations</span> in the polar region. This work provides an initial validation of the NCAR-TIMEGCM using <span class="hlt">airglow</span> data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19740018790','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19740018790"><span>Laboratory studies on the excitation and collisional deactivation of metastable atoms and molecules in the aurora and <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zipf, E. C.</p> <p>1974-01-01</p> <p>The aeronomy group at the University of Pittsburgh is actively engaged in a series of coordinated satellite, sounding rocket, and laboratory studies designed to expand and clarify knowledge of the physics and chemistry of planetary atmospheres. Three major discoveries have been made that will lead ultimately to a complete and dramatic revision of our ideas on the ionospheres of Mars, Venus, and the Earth and on the origin of their vacuum ultraviolet <span class="hlt">airglows</span>. The results have already suggested a new generation of ionosphere studies which probably can be carried out best by laser heterodyning techniques. Laboratory studies have also identified, for the first time, the physical mechanism responsible for the remarkable nitric oxide buildup <span class="hlt">observed</span> in some auroral arcs. This development is an important break-through in auroral physics, and has military ramifications of considerable interest to the Department of Defense. This work may also shed some light on related NO and atomic nitrogen problems in the mesosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ACP....18.6691D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ACP....18.6691D"><span>The <span class="hlt">airglow</span> layer emission altitude cannot be determined unambiguously from temperature comparison with lidars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dunker, Tim</p> <p>2018-05-01</p> <p>I investigate the nightly mean emission height and width of the OH* (3-1) layer by comparing nightly mean temperatures measured by the ground-based spectrometer GRIPS 9 and the Na lidar at ALOMAR. The data set contains 42 coincident measurements taken between November 2010 and February 2014, when GRIPS 9 was in operation at the ALOMAR observatory (69.3° N, 16.0° E) in northern Norway. To closely resemble the mean temperature measured by GRIPS 9, I weight each nightly mean temperature profile measured by the lidar using Gaussian distributions with 40 different centre altitudes and 40 different full widths at half maximum. In principle, one can thus determine the altitude and width of an <span class="hlt">airglow</span> layer by finding the minimum temperature difference between the two instruments. On most nights, several combinations of centre altitude and width yield a temperature difference of ±2 K. The generally assumed altitude of 87 km and width of 8 km is never an unambiguous, good solution for any of the measurements. Even for a fixed width of ˜ 8.4 km, one can sometimes find several centre altitudes that yield equally good temperature agreement. Weighted temperatures measured by lidar are not suitable to unambiguously determine the emission height and width of an <span class="hlt">airglow</span> layer. However, when actual altitude and width data are lacking, a comparison with lidars can provide an estimate of how representative a measured rotational temperature is of an assumed altitude and width. I found the rotational temperature to represent the temperature at the commonly assumed altitude of 87.4 km and width of 8.4 km to within ±16 K, on average. This is not a measurement uncertainty.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AdSpR..38.2374C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AdSpR..38.2374C"><span><span class="hlt">Observation</span> of temperatures and emission rates from the OH and O 2 nightglow over a southern high latitude station</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chung, J.-K.; Kim, Y. H.; Won, Y.-I.; Moon, B. K.; Oh, T. H.</p> <p>2006-01-01</p> <p>A Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span> (SATI) was operated at King Sejong Station (62°13'S, 58°47'W), Korea Antarctic Research Station during the period of March, 2002-September, 2003. We analyze rotational temperatures and emission rates of the O 2 (0-1) and OH (6-2) nightglows obtained at 67 nights with clear sky lasting more than 4 h. A spectral analysis of the dataset shows two dominant oscillations with periods of 4 and 6 h. The 6-h oscillations have a nearly constant phase, whereas the 4-h oscillations have nearly random phases. Although the harmonic periods of both oscillations are suggestive of tidal origin, the 4-h oscillation may have interference by other sources such as gravity waves. The 6-h oscillations could be interpreted as zonally symmetric non-migrating tides because migrating tides except high order modes have very weak amplitudes at high latitudes according to the classical tidal theory. For most cases of the <span class="hlt">observed</span> oscillations the temperature peak leads the intensity peak, which is consistent with theoretical models for zonally symmetric tides, but contrary to other theoretical models for waves. It is needed to resolve among theoretical models whether or not zonally symmetric tide cause temperature variation prior to intensity variation in mesospheric <span class="hlt">airglows</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA31A2561S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA31A2561S"><span>Global-scale <span class="hlt">Observations</span> of the Limb and Disk (GOLD): Science Implementation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Solomon, S. C.; McClintock, W. E.; Eastes, R.; Anderson, D. N.; Andersson, L.; Burns, A. G.; Codrescu, M.; Daniell, R. E.; England, S.; Eparvier, F. G.; Evans, J. S.; Krywonos, A.; Lumpe, J. D.; Richmond, A. D.; Rusch, D. W.; Siegmund, O.; Woods, T. N.</p> <p>2017-12-01</p> <p>The Global-scale <span class="hlt">Observations</span> of the Limb and Disk (GOLD) is a NASA mission of opportunity that will <span class="hlt">image</span> the Earth's thermosphere and ionosphere from geostationary orbit. GOLD will investigate how the thermosphere-ionosphere (T-I) system responds to geomagnetic storms, solar radiation, and upward propagating tides and how the structure of the equatorial ionosphere influences the formation and evolution of equatorial plasma density irregularities. GOLD consists of a pair of identical <span class="hlt">imaging</span> spectrographs that will measure <span class="hlt">airglow</span> emissions at far-ultraviolet wavelengths from 132 to 162 nm. On the disk, temperature and composition will be determined during the day using emissions from molecular nitrogen Lyman-Birge-Hopfield (LBH) band and atomic oxygen 135.6 nm, and electron density will be derived at night from 135.6 nm emission. On the limb, exospheric temperature will be derived from LBH emission profiles, and molecular oxygen density will be measured using stellar occultations. This presentation describes the GOLD mission science implementation including the as-built instrument performance and the planned <span class="hlt">observing</span> scenario. It also describes the results of simulations performed by the GOLD team to validate that the measured instrument performance and <span class="hlt">observing</span> plan will return adequate data to address the science objectives of the mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AMT....10.4601S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AMT....10.4601S"><span>Tomographic reconstruction of atmospheric gravity wave parameters from <span class="hlt">airglow</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>Song, Rui; Kaufmann, Martin; Ungermann, Jörn; Ern, Manfred; Liu, Guang; Riese, Martin</p> <p>2017-11-01</p> <p>Gravity waves (GWs) play an important role in the dynamics of the mesosphere and lower thermosphere (MLT). Therefore, global <span class="hlt">observations</span> of GWs in the MLT region are of particular interest. The small scales of GWs, however, pose a major problem for the <span class="hlt">observation</span> of GWs from space. We propose a new <span class="hlt">observation</span> strategy for GWs in the mesopause region by combining limb and sub-limb satellite-borne remote sensing measurements for improving the spatial resolution of temperatures that are retrieved from atmospheric soundings. In our study, we simulate satellite <span class="hlt">observations</span> of the rotational structure of the O2 A-band nightglow. A key element of the new method is the ability of the instrument or the satellite to operate in so-called <q>target mode</q>, i.e. to point at a particular point in the atmosphere and collect radiances at different viewing angles. These multi-angle measurements of a selected region allow for tomographic 2-D reconstruction of the atmospheric state, in particular of GW structures. The feasibility of this tomographic retrieval approach is assessed using simulated measurements. It shows that one major advantage of this <span class="hlt">observation</span> strategy is that GWs can be <span class="hlt">observed</span> on a much smaller scale than conventional <span class="hlt">observations</span>. We derive a GW sensitivity function, and it is shown that <q>target mode</q> <span class="hlt">observations</span> are able to capture GWs with horizontal wavelengths as short as ˜ 50 km for a large range of vertical wavelengths. This is far better than the horizontal wavelength limit of 100-200 km obtained from conventional limb sounding.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT........63V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT........63V"><span>A study of the feasibility and performance of an active/passive <span class="hlt">imager</span> using silicon focal plane arrays and incoherent continuous wave laser diodes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vollmerhausen, Richard H.</p> <p></p> <p>This dissertation describes an active/passive <span class="hlt">imager</span> (API) that provides reliable, nighttime, target acquisition in a man-portable package with effective visual range of about 4 kilometers. The reflective imagery is easier to interpret than currently used thermal imagery. Also, in the active mode, the API provides performance equivalent to the big-aperture, thermal systems used on weapons platforms like tanks and attack helicopters. This dissertation describes the research needed to demonstrate both the feasibility and utility of the API. Part of the research describes implementation of a silicon focal plane array (SFPA) capable of both active and passive <span class="hlt">imaging</span>. The passive <span class="hlt">imaging</span> mode exceeds the nighttime performance of currently fielded, man-portable sensors. Further, when scene illumination is insufficient for passive <span class="hlt">imaging</span>, the low dark current of SFPA makes it possible to use continuous wave laser diodes (CWLD) to add an active <span class="hlt">imaging</span> mode. CWLD have advantages of size, efficiency, and improved eye safety when compared to high peak-power diodes. Because of the improved eye safety, the API provides user-demanded features like video output and extended range gates in the active as well as passive <span class="hlt">imaging</span> modes. Like any other night vision device, the API depends on natural illumination of the scene for passive operation. Although it has been known for decades that "starlight" illumination is actually from diffuse <span class="hlt">airglow</span> emissions, the research described in this dissertation provides the first estimates of the global and temporal variation of ground illumination due to <span class="hlt">airglow</span>. A third related element of the current research establishes the impact of atmospheric aerosols on API performance. We know from day experience that atmospheric scattering of sunlight into the <span class="hlt">imager</span> line-of-sight can blind the <span class="hlt">imager</span> and drastically degrade performance. Atmospheric scattering of sunlight is extensively covered in the literature. However, previous literature did not</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004cosp...35.3258C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004cosp...35.3258C"><span><span class="hlt">Observation</span> of upper mesospheric temperatures and emission rates from the OH and O2 nightglow over King Sejong Station, Antartica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chung, J.-K.; Kim, Y. H.; Moon, B.-G.; Oh, T.-H.; Won, Y.-I.</p> <p></p> <p>A spectral <span class="hlt">airglow</span> temperature <span class="hlt">imager</span> (SATI) was operated at King Sejong Station (62.22^oS, 301.2^oE), Korea Antarctic Research Station during the period March, 2002 through October, 2003. We analyze data obtained at 24 and 22 nights with clear sky condition lasting more than 6 hours in 2002 and 2003, respectively. A dominant and coherent 4-hr oscillation was seen in both the OH(6-2) and O_2(0-1) band <span class="hlt">airglow</span> rotational temperatures at two nights, and similar weak features appeared at several nights . The data also show fluctuations of long period that seem to relate to tides, and short period oscillations that could be due to propagating gravity waves. Detailed harmonic analysis will be performed to seasonal data sets to identify any variation in the major atmospheric oscillations over season.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19800026554&hterms=day+night&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dday%2Bnight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19800026554&hterms=day+night&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dday%2Bnight"><span>O2/1 Delta/ emission in the day and night <span class="hlt">airglow</span> of Venus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Connes, P.; Noxon, J. F.; Traub, W. A.; Carleton, N. P.</p> <p>1979-01-01</p> <p>An intense <span class="hlt">airglow</span> from O2(1 Delta) at 1.27 microns on both the light and the dark sides of Venus has been detected by using a ground-based high-resolution Fourier-transform spectrometer. Both dayglow and nightglow are roughly 1,000 times brighter than the visible O2 nightglow found by Veneras 9 and 10 in 1975. The column emission rate of O2(1 Delta) from Venus is close to the rate at which fresh O atoms are produced from photolysis of CO2 on the day side. Formation of O2(1 Delta) is thus a major step in the removal of O atoms from the atmosphere, and dynamical processes must carry these atoms to the night side fast enough to yield a maximum density near 90 km, which is almost constant over the planet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870007337','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870007337"><span>In search of stratospheric bromine oxide</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lestrade, John Patrick</p> <p>1986-01-01</p> <p>The <span class="hlt">Imaging</span> Spectrometric Observatory (ISO) is capable of recording spectra in the wavelength range of 200 to 12000 Angstroms. Data from a recent Spacelab 1 ATLAS mission has <span class="hlt">imaged</span> the terrestrial <span class="hlt">airglow</span> at tangent ray heights of 90 and 150 km. These data contain information about trace atmospheric constituents such as bromine oxide (BrO), hydroxyl (OH), and chlorine dioxide (OClO). The abundances of these species are critical to stratospheric models of catalytic ozone destruction. Heretofore, very few <span class="hlt">observations</span> were made especially for BrO. Software was developed to purge unwanted solar features from the <span class="hlt">airglow</span> spectra. The next step is a measure of the strength of the emission features for BrO. The final analysis will yield the scale height of this important compound.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JASTP.155...86K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JASTP.155...86K"><span>Statistical analysis of mesospheric gravity waves over King Sejong Station, Antarctica (62.2°S, 58.8°W)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kam, Hosik; Jee, Geonhwa; Kim, Yong; Ham, Young-bae; Song, In-Sun</p> <p>2017-03-01</p> <p>We have investigated the characteristics of mesospheric short period (<1 h) gravity waves which were <span class="hlt">observed</span> with all-sky <span class="hlt">images</span> of OH Meinel band and OI 557 nm <span class="hlt">airglows</span> over King Sejong Station (KSS) (62.22°S, 58.78°W) during a period of 2008-2015. By applying 2-dimensional FFT to time differenced <span class="hlt">images</span>, we derived horizontal wavelengths, phase speeds, and propagating directions (188 and 173 quasi-monochromatic waves from OH and OI <span class="hlt">airglow</span> <span class="hlt">images</span>, respectively). The majority of the <span class="hlt">observed</span> waves propagated predominantly westward, implying that eastward waves were filtered out by strong eastward stratospheric winds. In order to obtain the intrinsic properties of the <span class="hlt">observed</span> waves, we utilized winds simultaneously measured by KSS Meteor Radar and temperatures from Aura Microwave Limb Sounder (MLS). More than half the waves propagated horizontally, as waves were in Doppler duct or evanescent in the vertical direction. This might be due to strong eastward background wind field in the mesosphere over KSS. For freely propagating waves, the vertical wavelengths were in the interquartile range of 9-33 km with a median value of 15 km. The vertical wavelengths are shorter than those <span class="hlt">observed</span> at Halley station (76°S, 27°W) where the majority of the <span class="hlt">observed</span> waves were freely propagating. The difference in the wave propagating characteristics between KSS and Halley station suggests that gravity waves may affect mesospheric dynamics in this part of the Antarctic Peninsula more strongly than over the Antarctic continent. Furthermore, strong wind shear over KSS played an important role in changing the vertical wavenumbers as the waves propagated upward between two <span class="hlt">airglow</span> layers (87 and 96 km).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JASTP.135..192W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JASTP.135..192W"><span>Spatial gravity wave characteristics obtained from multiple OH(3-1) <span class="hlt">airglow</span> temperature time series</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wachter, Paul; Schmidt, Carsten; Wüst, Sabine; Bittner, Michael</p> <p>2015-12-01</p> <p>We present a new approach for the detection of gravity waves in OH-<span class="hlt">airglow</span> <span class="hlt">observations</span> at the measurement site Oberpfaffenhofen (11.27°E, 48.08°N), Germany. The measurements were performed at the German Remote Sensing Data Center (DFD) of the German Aerospace Center (DLR) during the period from February 4th, 2011 to July 6th, 2011. In this case study the <span class="hlt">observations</span> were carried out by three identical Ground-based Infrared P-branch Spectrometers (GRIPS). These instruments provide OH(3-1) rotational temperature time series, which enable spatio-temporal investigations of gravity wave characteristics in the mesopause region. The instruments were aligned in such a way that their fields of view (FOV) formed an equilateral triangle in the OH-emission layer at a height of 87 km. The Harmonic Analysis is applied in order to identify joint temperature oscillations in the three individual datasets. Dependent on the specific gravity wave activity in a single night, it is possible to detect up to four different wave patterns with this method. The values obtained for the waves' periods and phases are then used to derive further parameters, such as horizontal wavelength, phase velocity and the direction of propagation. We identify systematic relationships between periods and amplitudes as well as between periods and horizontal wavelengths. A predominant propagation direction towards the East and North-North-East characterizes the waves during the <span class="hlt">observation</span> period. There are also indications of seasonal effects in the temporal development of the horizontal wavelength and the phase velocity. During late winter and early spring the derived horizontal wavelengths and the phase velocities are smaller than in the subsequent period from early April to July 2011.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/227151-analysis-lyman-alpha-he-angstrom-airglow-measurements-using-spherical-radiative-transfer-model','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/227151-analysis-lyman-alpha-he-angstrom-airglow-measurements-using-spherical-radiative-transfer-model"><span>Analysis of Lyman {alpha} and He I 584-{Angstrom} <span class="hlt">airglow</span> measurements using a spherical radiative transfer 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>Bush, B.C.; Chakrabarti, S.</p> <p>1995-10-01</p> <p>The authors report on the scattering and excitation mechanisms of the terrestrial exospheric H I 1216-{Angstrom} <span class="hlt">airglow</span> emissions by comparing simulations from a radiative transfer model with spectroscopic measurements from an Earth-orbiting satellite. The purpose of these comparisons are twofold: to assess the sensitivity of the input parameters to the model results and to test the applicability of the model to <span class="hlt">airglow</span> analysis. The model incorporates a spherically oriented atmosphere to account for the extended scale heights of the exospheric scatterers as well as to properly mimic scattering across the terminator region from the dayside to the nightside hemispheres. Spectroscopicmore » Lyman {alpha} and He I 584 {Angstrom} data were obtained by the STP78-1 satellite that circumnavigated the Earth in a noon/midnight orbit at an altitude of 600 km. The {open_quotes}best fit{close_quotes} analysis of the Lyman {alpha} data acquired on March 25, 1979, requires scaling the hydrogen density distribution obtained from the MSIS-90 (Hedin) atmospheric model by 45-50%, the exospheric temperature by 90-100%, and the Lyman {alpha} solar flux predicted by EUV91 model (Tobiska) by 1.9-2.0. Similar analysis of the He I 584 {Angstrom} data acquired on March 5, 1979, requires scaling the helium density distribution obtained from the MSIS-90 (Hedin) atmospheric model by 60-80% and the exospheric temperature by 105-115% while using a line center 584-{Angstrom} solar flux of 1.44x10{sup 10} photons cm{sup {minus}2}s{sup {minus}1} {Angstrom}{sup {minus}1}. 46 refs., 22 figs., 5 tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014cosp...40E3040S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014cosp...40E3040S"><span><span class="hlt">Airglow</span> at 630 and 557.7 nm during HF pumping of the Ionosphere near the 4th Gyroharmonic at the ``Sura'' Facility in September 2012</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shindin, Alexey; Nasyrov, Igor; Grach, Savely; Sergeev, Evgeny; Klimenko, Vladimir; Beletsky, Alexandr</p> <p></p> <p>We present results of artificial optical emission <span class="hlt">observations</span> in the red (630 nm) and green (557.7 nm) lines of the atomic oxygen during ionosphere HF pumping at the Sura facility (56.1°N, 46.1°E, magnetic field dip angle 71.5°) in Sep. 2012. Pump wave (PW) of O-polarization at frequencies f0 = 4.74 - 5.64 MHz was used in the experiment according to ionospheric conditions after sunset. Two CCD cameras (S1C/079-FP(FU) and KEO Sentinel with fields of view 20° and 145°, respectively, and 3 photometers were used for the emission registration. For estimation of a relation between the PW frequency f0 and 4th electron gyroharmonic 4fce Stimulated Electromagnetic Emission (SEE) registration was applied (for details see [1]). On September 11 the pump beam was inclined by 12° to the South, the PW frequencies f0 = 5.40 and 5.42 MHz were slightly above 4fce. On September 13, for vertical pumping, f0 was 5.64 MHz (well above 4fce), 5.32 - 5.42 MHz (around 4fce) and 4.74 MHz (well below 4fce). On September 14 the vertical pumping at f0 = 5.30 - 5.36 MHz and 4.74 MHz was used. In the latter day due to natural motion of the ionosphere and concurrent SEE measurements we were able to obtain a fine dependence of the optical brightness on the proximity f0 and 4fce. For the red line no essential dependence, as well of the shape and position of the <span class="hlt">airglow</span> spot on the proximity was obtained with one exception: on Sep. 14 when, according to the SEE spectra, f0 was just below 4fce (by 15-20 kHz), the brightness essentially increased, by 1.25-1.5 times. For the green line, the brightest emission occurred when f0 was passing through 4fce (Sep. 14) and when f0 = 5.64 MHz (Sep. 13, well above 4fce). Also, on Sep. 14 the <span class="hlt">airglow</span> enhancement in the red line during the pumping was replaced by the suppression of the background emission when the ionosphere critical frequency approached to f0 by less than 500 kHz. Similar effect was obtained on Sep. 11 and in [2] for south-inclined pump beam</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMSH21A4094P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSH21A4094P"><span>Bayesian Analysis Of HMI Solar <span class="hlt">Image</span> <span class="hlt">Observables</span> And Comparison To TSI Variations And MWO <span class="hlt">Image</span> <span class="hlt">Observables</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parker, D. G.; Ulrich, R. K.; Beck, J.</p> <p>2014-12-01</p> <p>We have previously applied the Bayesian automatic classification system AutoClass to solar magnetogram and intensity <span class="hlt">images</span> from the 150 Foot Solar Tower at Mount Wilson to identify classes of solar surface features associated with variations in total solar irradiance (TSI) and, using those identifications, modeled TSI time series with improved accuracy (r > 0.96). (Ulrich, et al, 2010) AutoClass identifies classes by a two-step process in which it: (1) finds, without human supervision, a set of class definitions based on specified attributes of a sample of the <span class="hlt">image</span> data pixels, such as magnetic field and intensity in the case of MWO <span class="hlt">images</span>, and (2) applies the class definitions thus found to new data sets to identify automatically in them the classes found in the sample set. HMI high resolution <span class="hlt">images</span> capture four <span class="hlt">observables</span>-magnetic field, continuum intensity, line depth and line width-in contrast to MWO's two <span class="hlt">observables</span>-magnetic field and intensity. In this study, we apply AutoClass to the HMI <span class="hlt">observables</span> for <span class="hlt">images</span> from May, 2010 to June, 2014 to identify solar surface feature classes. We use contemporaneous TSI measurements to determine whether and how variations in the HMI classes are related to TSI variations and compare the characteristic statistics of the HMI classes to those found from MWO <span class="hlt">images</span>. We also attempt to derive scale factors between the HMI and MWO magnetic and intensity <span class="hlt">observables</span>. The ability to categorize automatically surface features in the HMI <span class="hlt">images</span> holds out the promise of consistent, relatively quick and manageable analysis of the large quantity of data available in these <span class="hlt">images</span>. Given that the classes found in MWO <span class="hlt">images</span> using AutoClass have been found to improve modeling of TSI, application of AutoClass to the more complex HMI <span class="hlt">images</span> should enhance understanding of the physical processes at work in solar surface features and their implications for the solar-terrestrial environment. Ulrich, R.K., Parker, D, Bertello, L. and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMNH23A0255O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMNH23A0255O"><span>Ionospheric detection of tsunami earthquakes: <span class="hlt">observation</span>, modeling and ideas for future early warning</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Occhipinti, G.; Manta, F.; Rolland, L.; Watada, S.; Makela, J. J.; Hill, E.; Astafieva, E.; Lognonne, P. H.</p> <p>2017-12-01</p> <p>Detection of ionospheric anomalies following the Sumatra and Tohoku earthquakes (e.g., Occhipinti 2015) demonstrated that ionosphere is sensitive to earthquake and tsunami propagation: ground and oceanic vertical displacement induces acoustic-gravity waves propagating within the neutral atmosphere and detectable in the ionosphere. <span class="hlt">Observations</span> supported by modelling proved that ionospheric anomalies related to tsunamis are deterministic and reproducible by numerical modeling via the ocean/neutral-atmosphere/ionosphere coupling mechanism (Occhipinti et al., 2008). To prove that the tsunami signature in the ionosphere is routinely detected we show here perturbations of total electron content (TEC) measured by GPS and following tsunamigenic earthquakes from 2004 to 2011 (Rolland et al. 2010, Occhipinti et al., 2013), nominally, Sumatra (26 December, 2004 and 12 September, 2007), Chile (14 November, 2007), Samoa (29 September, 2009) and the recent Tohoku-Oki (11 Mars, 2011). Based on the <span class="hlt">observations</span> close to the epicenter, mainly performed by GPS networks located in Sumatra, Chile and Japan, we highlight the TEC perturbation <span class="hlt">observed</span> within the first 8 min after the seismic rupture. This perturbation contains information about the ground displacement, as well as the consequent sea surface displacement resulting in the tsunami. In addition to GNSS-TEC <span class="hlt">observations</span> close to the epicenter, new exciting measurements in the far-field were performed by <span class="hlt">airglow</span> measurement in Hawaii show the propagation of the internal gravity waves induced by the Tohoku tsunami (Occhipinti et al., 2011). This revolutionary <span class="hlt">imaging</span> technique is today supported by two new <span class="hlt">observations</span> of moderate tsunamis: Queen Charlotte (M: 7.7, 27 October, 2013) and Chile (M: 8.2, 16 September 2015). We finally detail here our recent work (Manta et al., 2017) on the case of tsunami alert failure following the Mw7.8 Mentawai event (25 October, 2010), and its twin tsunami alert response following the Mw7</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('https://ntrs.nasa.gov/search.jsp?R=20170003170&hterms=Situ&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DIn%2BSitu','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003170&hterms=Situ&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DIn%2BSitu"><span>Simultaneous <span class="hlt">Observations</span> of Atmospheric Tides from Combined in Situ and Remote <span class="hlt">Observations</span> at Mars from the MAVEN Spacecraft</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>England, Scott L.; Liu, Guiping; Withers, Paul; Yigit, Erdal; Lo, Daniel; Jain, Sonal; Schneider, Nicholas M. (Inventor); Deighan, Justin; McClintock, William E.; Mahaffy, Paul R.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20170003170'); toggleEditAbsImage('author_20170003170_show'); toggleEditAbsImage('author_20170003170_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20170003170_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20170003170_hide"></p> <p>2016-01-01</p> <p>We report the <span class="hlt">observations</span> of longitudinal variations in the Martian thermosphere associated with nonmigrating tides. Using the Neutral Gas Ion Mass Spectrometer (NGIMS) and the <span class="hlt">Imaging</span> Ultraviolet Spectrograph (IUVS) on NASA's Mars Atmosphere and Volatile EvolutioN Mission (MAVEN) spacecraft, this study presents the first combined analysis of in situ and remote <span class="hlt">observations</span> of atmospheric tides at Mars for overlapping volumes, local times, and overlapping date ranges. From the IUVS <span class="hlt">observations</span>, we determine the altitude and latitudinal variation of the amplitude of the nonmigrating tidal signatures, which is combined with the NGIMS, providing information on the compositional impact of these waves. Both the <span class="hlt">observations</span> of <span class="hlt">airglow</span> from IUVS and the CO2 density <span class="hlt">observations</span> from NGIMS reveal a strong wave number 2 signature in a fixed local time frame. The IUVS <span class="hlt">observations</span> reveal a strong latitudinal dependence in the amplitude of the wave number 2 signature. Combining this with the accurate CO2 density <span class="hlt">observations</span> from NGIMS, this would suggest that the CO2 density variation is as high as 27% at 0-10 deg latitude. The IUVS <span class="hlt">observations</span> reveal little altitudinal dependence in the amplitude of the wave number 2 signature, varying by only 20% from 160 to 200 km. <span class="hlt">Observations</span> of five different species with NGIMS show that the amplitude of the wave number 2 signature varies in proportion to the inverse of the species scale height, giving rise to variation in composition as a function of longitude. The analysis and discussion here provide a roadmap for further analysis as additional coincident data from these two instruments become available.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss039e009160.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss039e009160.html"><span>Earth <span class="hlt">Observations</span> taken by the Expedition 39 Crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-04-02</p> <p>ISS039-E-009160 (2 April 2014) --- This nighttime view featuring the aurora borealis, the moon and Moscow was photographed by an Expedition 39 crew member on the International Space Station. A thin green line of the aurora borealis crosses the top of this <span class="hlt">image</span>. The moon appears as a white disc just above the aurora. <span class="hlt">Airglow</span> appears as a blue-white cusp on Earth's limb. Russia's capital city Moscow makes a splash of yellow (lower left), with its easily recognized radial pattern of highways. Other cities are Nizhni Novgorod (lower center) 400 kilometers from Moscow, St. Petersburg (left) 625 kilometers from Moscow, and Finland?s capital city Helsinki.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006cosp...36.2681S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006cosp...36.2681S"><span><span class="hlt">Observation</span> sequences and onboard data processing of Planet-C</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suzuki, M.; Imamura, T.; Nakamura, M.; Ishi, N.; Ueno, M.; Hihara, H.; Abe, T.; Yamada, T.</p> <p></p> <p>Planet-C or VCO Venus Climate Orbiter will carry 5 cameras IR1 IR 1micrometer camera IR2 IR 2micrometer camera UVI UV <span class="hlt">Imager</span> LIR Long-IR camera and LAC Lightning and <span class="hlt">Airglow</span> Camera in the UV-IR region to investigate atmospheric dynamics of Venus During 30 hr orbiting designed to quasi-synchronize to the super rotation of the Venus atmosphere 3 groups of scientific <span class="hlt">observations</span> will be carried out i <span class="hlt">image</span> acquisition of 4 cameras IR1 IR2 UVI LIR 20 min in 2 hrs ii LAC operation only when VCO is within Venus shadow and iii radio occultation These <span class="hlt">observation</span> sequences will define the scientific outputs of VCO program but the sequences must be compromised with command telemetry downlink and thermal power conditions For maximizing science data downlink it must be well compressed and the compression efficiency and <span class="hlt">image</span> quality have the significant scientific importance in the VCO program <span class="hlt">Images</span> of 4 cameras IR1 2 and UVI 1Kx1K and LIR 240x240 will be compressed using JPEG2000 J2K standard J2K is selected because of a no block noise b efficiency c both reversible and irreversible d patent loyalty free and e already implemented as academic commercial software ICs and ASIC logic designs Data compression efficiencies of J2K are about 0 3 reversible and 0 1 sim 0 01 irreversible The DE Digital Electronics unit which controls 4 cameras and handles onboard data processing compression is under concept design stage It is concluded that the J2K data compression logics circuits using space</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMSA21B..01M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMSA21B..01M"><span><span class="hlt">Imaging</span>, radio, and modeling results pertaining to the ionospheric signature of the 11 March 2011 tsunami over the Pacific Ocean</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Makela, J. J.; Lognonne, P.; Occhipinti, G.; Hebert, H.; Gehrels, T.; Coisson, P.; Rolland, L. M.; Allgeyer, S.; Kherani, A.</p> <p>2011-12-01</p> <p>The Mw=9.0 earthquake that occurred off the east coast of Honshu, Japan on 11 March 2011 launched a tsunami that traveled across the Pacific Ocean, in turn launching vertically propagating atmospheric gravity waves. Upon reaching 250-350 km in altitude, these waves impressed their signature on the thermosphere/ionosphere system. We present <span class="hlt">observations</span> of this signature obtained using a variety of radio instruments and an <span class="hlt">imaging</span> system located on the islands of Hawaii. These measurements represent the first optical <span class="hlt">images</span> recorded of the <span class="hlt">airglow</span> signature resulting from the passage of a tsunami. Results from these instruments clearly show wave structure propagating in the upper atmosphere with the same velocity as the ocean tsunami, emphasizing the coupled nature of the ocean, atmosphere, and ionosphere. Modeling results are also presented to highlight current understandings of this coupling process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018Ap%26SS.363...83S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018Ap%26SS.363...83S"><span>Study of equatorial plasma bubbles using all sky <span class="hlt">imager</span> and scintillation technique from Kolhapur station: a case study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharma, A. K.; Gurav, O. B.; Gaikwad, H. P.; Chavan, G. A.; Nade, D. P.; Nikte, S. S.; Ghodpage, R. N.; Patil, P. T.</p> <p>2018-04-01</p> <p>The nightglow <span class="hlt">observations</span> of OI 630.0 nm emission carried out from low latitude station Kolhapur using All Sky <span class="hlt">Imager</span> (ASI) with 140° field of view (FOV) for the month of April 2011 are used. The <span class="hlt">images</span> were processed to study the field aligned irregularities often called as equatorial plasma bubbles (EPBs). The present study focuses on the occurrence of scintillation during the traversal of EPBs over ionospheric pierce point (IPP). Here we dealt with the depletion level (depth) of the EPB structures and its effect on VHF signals. We compared VHF scintillation data with <span class="hlt">airglow</span> intensities at Ionospheric pierce point (IPP) from the same location and found that the largely depleted EPBs make stronger scintillation. From previous literature, it is believed that the small scale structures are present near the steeper walls of EPBs which often degrades the communication, the analysis presented in this paper confirms this belief.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990009871&hterms=monographs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dmonographs','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990009871&hterms=monographs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dmonographs"><span>Auroral <span class="hlt">Observations</span> from the POLAR Ultraviolet <span class="hlt">Imager</span> (UVI)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Germany, G. A.; Spann, J. F.; Parks, G. K.; Brittnacher, M. J.; Elsen, R.; Chen, L.; Lummerzheim, D.; Rees, M. H.</p> <p>1998-01-01</p> <p>Because of the importance of the auroral regions as a remote diagnostic of near-Earth plasma processes and magnetospheric structure, spacebased instrumentation for <span class="hlt">imaging</span> the auroral regions have been designed and operated for the last twenty-five years. The latest generation of <span class="hlt">imagers</span>, including those flown on the POLAR satellite, extends this quest for multispectral resolution by providing three separate <span class="hlt">imagers</span> for the visible, ultraviolet, and X ray <span class="hlt">images</span> of the aurora. The ability to <span class="hlt">observe</span> extended regions allows <span class="hlt">imaging</span> missions to significantly extend the <span class="hlt">observations</span> available from in situ or groundbased instrumentation. The complementary nature of <span class="hlt">imaging</span> and other <span class="hlt">observations</span> is illustrated below using results from tile GGS Ultraviolet <span class="hlt">Imager</span> (UVI). Details of the requisite energy and intensity analysis are also presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930017596','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930017596"><span>Ultraviolet <span class="hlt">Imaging</span> Telescope (UIT) <span class="hlt">observations</span> of galaxies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Neff, S. G.</p> <p>1993-01-01</p> <p>Ultraviolet <span class="hlt">images</span> of several galaxies were obtained during the ASTRO-1 shuttle mission in December, 1990. The <span class="hlt">images</span> have a FWHM angular resolution of approximately 3 arcsecond and are of circular fields approximately 40 arcminutes in diameter. Most galaxies were <span class="hlt">observed</span> in at least two and sometimes as many as four broad bands. A very few fields were <span class="hlt">observed</span> with narrower band filters. The most basic result of these <span class="hlt">observations</span> is that most systems look dramatically different in the UV from their well-known optical appearances. Preliminary results of these studies will be presented. Information will be available on fields <span class="hlt">observed</span> by the UTI during the ASTRO 1 mission; when that data becomes public it can be obtained from the NSSDC. The ASTRO observatory is expected to fly again in 1994 with approximately half of the <span class="hlt">observing</span> time from that mission devoted to guest <span class="hlt">observers</span>. The Ultraviolet <span class="hlt">Imaging</span> telescope is extremely well suited for galaxy studies, and the UIT term is interested in encouraging a wide range of scientific studies by guest <span class="hlt">observers</span>. Ultraviolet <span class="hlt">Imaging</span> telescope is extremely well suited for galaxy studies, and the UIT team is interested in encouraging a wide range of scientific studies by guest <span class="hlt">observers</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA13B..05Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA13B..05Y"><span>Interhemispheric Asymmetry in the Mesosphere and Lower Thermosphere <span class="hlt">Observed</span> by SABER/TIMED</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yee, J. H.</p> <p>2017-12-01</p> <p>In this paper we analyze nearly 15 years of satellite <span class="hlt">observations</span> of temperature, <span class="hlt">airglow</span>, and composition in the Mesosphere and Lower Thermosphere (MLT) to quantify their interhemispheric asymmetries ao one can provide quantitative links between <span class="hlt">observed</span> asymmetries and the spatial and temporal variations of the gravity wave activity. Two processes are believed to be responsible for <span class="hlt">observed</span> interhemispheric differences in the MLT. The first is the direct radiation effect from the eccentricity of the Earth orbit amd the other is the difference in gravity wave source distribution and filtering due to asymmetries in mean winds of the lower atmosphere. Both processes have been theoretically investigated to explain the <span class="hlt">observed</span> asymmetry in some of the atmospheric parameters, but not self-consistently in all <span class="hlt">observed</span> parameters together. In this paper we will show the asymmetry in the time-varying zonal-mean latitudinal structures of temperature, <span class="hlt">airglow</span> emission rate, and composition <span class="hlt">observed</span> by TIMED/SABER. We will quantify their interhemispheric asymmetries for different seasons under different solar activity conditions. In addition, temperature measurements will also be used to obtain temporal and spatial morphology of gravity wave potential energies. We will interpret the asymmetry in the <span class="hlt">observed</span> fields and examine qualitatively their consistency with the two responsible processes, especially the one due to gravity wave filtering process. Our goal is to introduce and to share the spatial and temporal morphologies of all the <span class="hlt">observed</span> fields to the modeling community so, together self-consistently, they be can be used to gain physical insights into the relative importance of various drivers responsible for the <span class="hlt">observed</span> asymmetry, especially the role of gravity wave induced eddy drag and mixing, a critical, but least quantitatively understood process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22069179-ray-phase-imaging-from-static-observation-dynamic-observation','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22069179-ray-phase-imaging-from-static-observation-dynamic-observation"><span>X-ray phase <span class="hlt">imaging</span>-From static <span class="hlt">observation</span> to dynamic <span class="hlt">observation</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>Momose, A.; Yashiro, W.; Olbinado, M. P.</p> <p>2012-07-31</p> <p>We are attempting to expand the technology of X-ray grating phase <span class="hlt">imaging</span>/tomography to enable dynamic <span class="hlt">observation</span>. X-ray phase <span class="hlt">imaging</span> has been performed mainly for static cases, and this challenge is significant since properties of materials (and hopefully their functions) would be understood by <span class="hlt">observing</span> their dynamics in addition to their structure, which is an inherent advantage of X-ray <span class="hlt">imaging</span>. Our recent activities in combination with white synchrotron radiation for this purpose are described. Taking advantage of the fact that an X-ray grating interferometer functions with X-rays of a broad energy bandwidth (and therefore high flux), movies of differential phase imagesmore » and visibility <span class="hlt">images</span> are obtained with a time resolution of a millisecond. The time resolution of X-ray phase tomography can therefore be a second. This study is performed as a part of a project to explore X-ray grating interferometry, and our other current activities are also briefly outlined.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JASS...31..169J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JASS...31..169J"><span>Ground-based <span class="hlt">Observations</span> for the Upper Atmosphere at King Sejong Station, Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jee, Geonhwa; Kim, Jeong-Han; Lee, Changsup; Kim, Yong Ha</p> <p>2014-06-01</p> <p>Since the operation of the King Sejong Station (KSS) started in Antarctic Peninsula in 1989, there have been continuous efforts to perform the <span class="hlt">observation</span> for the upper atmosphere. The <span class="hlt">observations</span> during the initial period of the station include Fabry-Perot Interferometer (FPI) and Michelson Interferometer for the mesosphere and thermosphere, which are no longer in operation. In 2002, in collaboration with York University, Canada, the Spectral <span class="hlt">Airglow</span> Temperature <span class="hlt">Imager</span> (SATI) was installed to <span class="hlt">observe</span> the temperature in the mesosphere and lower thermosphere (MLT) region and it has still been producing the mesopause temperature data until present. The <span class="hlt">observation</span> was extended by installing the meteor radar in 2007 to <span class="hlt">observe</span> the neutral winds and temperature in the MLT region during the day and night in collaboration with Chungnam National University. We also installed the all sky camera in 2008 to <span class="hlt">observe</span> the wave structures in the MLT region. All these <span class="hlt">observations</span> are utilized to study on the physical characteristics of the MLT region and also on the wave phenomena such as the tide and gravity wave in the upper atmosphere over KSS that is well known for the strong gravity wave activity. In this article, brief introductions for the currently operating instruments at KSS will be presented with their applications for the study of the upper atmosphere</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA23A2379C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA23A2379C"><span>Analysis of TIMED/GUVI Dayglow Utraviolet Oxygen <span class="hlt">Images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Christensen, A. B.; Crowley, G.; Meier, R.</p> <p>2016-12-01</p> <p>Analysis of the atomic oxygen resonance transition at 130.4 nm and the inter-combination transition at 135.6 nm measured by the TIMED/GUVI mission demonstrates the state of knowledge of these important dayglow emission features and the degree to which current models can simulate their global properties. The complete modeling framework comprises several models, including the Thermosphere ionosphere Mesosphere Electrodynamics General Circulation Model (TIME-GCM), Assimilative Mapping of Ionospheric Electrodynamics (AMIE), a partial frequency redistribution resonance scattering model usually called REDISTER needed to compute the optically thick radiative transfer of the 130.4 nm emission, <span class="hlt">airglow</span> emission models, GLOW and AURIC and other procedures. <span class="hlt">Observations</span> for four different days, collected under different geophysical conditions of magnetic activity and solar cycle, show very good agreement with the calculated emission brightness and geographic distribution for both emissions. The differences between the <span class="hlt">airglow</span> codes for the 135.6 nm emission will be discussed in connection to the photoelectron energy loss cross sections, as well as the excitation cross sections used in the various models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15839266','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15839266"><span>Multifacet structure of <span class="hlt">observed</span> reconstructed integral <span class="hlt">images</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Martínez-Corral, Manuel; Javidi, Bahram; Martínez-Cuenca, Raúl; Saavedra, Genaro</p> <p>2005-04-01</p> <p>Three-dimensional <span class="hlt">images</span> generated by an integral <span class="hlt">imaging</span> system suffer from degradations in the form of grid of multiple facets. This multifacet structure breaks the continuity of the <span class="hlt">observed</span> <span class="hlt">image</span> and therefore reduces its visual quality. We perform an analysis of this effect and present the guidelines in the design of lenslet <span class="hlt">imaging</span> parameters for optimization of viewing conditions with respect to the multifacet degradation. We consider the optimization of the system in terms of field of view, <span class="hlt">observer</span> position and pupil function, lenslet parameters, and type of reconstruction. Numerical tests are presented to verify the theoretical analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRA..121.8134F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRA..121.8134F"><span>The <span class="hlt">Imager</span> for Sprites and Upper Atmospheric Lightning (ISUAL)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Frey, H. U.; Mende, S. B.; Harris, S. E.; Heetderks, H.; Takahashi, Y.; Su, H.-T.; Hsu, R.-R.; Chen, A. B.; Fukunishi, H.; Chang, Y.-S.; Lee, L.-C.</p> <p>2016-08-01</p> <p>The <span class="hlt">Imager</span> for Sprites and Upper Atmospheric Lightning (ISUAL) was the first specifically dedicated instrument to <span class="hlt">observe</span> lightning-induced transient luminous events (TLE): sprites, elves, halos, and gigantic jets from space. The <span class="hlt">Imager</span> is an intensified CCD system operating in the visible wavelength region with a filter wheel to select from six positions with filters. The <span class="hlt">Imager</span> has a 5° × 20° (vertical times horizontal) field of view. The spectrophotometer (SP) is populated with six photometers with individual filters for emissions from the far ultraviolet to the near infrared. An array photometer with two channels operating in the blue and red provides altitude profiles of the emission over 16 altitude bins each. The Associated Electronics Package (AEP) controls instrument functions and interfaces with the spacecraft. ISUAL was launched 21 May 2004 into a Sun-synchronous 890 km orbit on the Formosat-2 satellite and has successfully been collecting data ever since. ISUAL is running on the nightside of the orbit and is pointed to the east of the orbit down toward the limb. The instrument runs continuously and writes data to a circular buffer. Whenever the SP detects a sudden signal increase above a preset threshold, a trigger signal is generated that commands the system to keep the data for about 400 ms starting from ~50 ms before the trigger. Over its lifetime of ~11 years the system recorded thousands of TLE and also successfully <span class="hlt">observed</span> aurora and <span class="hlt">airglow</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016DPS....4851511R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016DPS....4851511R"><span>Chapman Solar Zenith Angle variations at Titan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Royer, Emilie M.; Ajello, Joseph; Holsclaw, Gregory; West, Robert; Esposito, Larry W.; Bradley, Eric Todd</p> <p>2016-10-01</p> <p>Solar XUV photons and magnetospheric particles are the two main sources contributing to the <span class="hlt">airglow</span> in the Titan's upper atmosphere. We are focusing here on the solar XUV photons and how they influence the <span class="hlt">airglow</span> intensity. The Cassini-UVIS <span class="hlt">observations</span> analyzed in this study consist each in a partial scan of Titan, while the center of the detector stays approximately at the same location on Titan's disk. We used <span class="hlt">observations</span> from 2008 to 2012, which allow for a wide range of Solar Zenith Angle (SZA). Spectra from 800 km to 1200 km of altitude have been corrected from the solar spectrum using TIMED/SEE data. We <span class="hlt">observe</span> that the <span class="hlt">airglow</span> intensity varies as a function of the SZA and follows a Chapman curve. Three SZA regions are identified: the sunlit region ranging from 0 to 50 degrees. In this region, the intensity of the <span class="hlt">airglow</span> increases, while the SZA decreases. Between SZA 50 and 100 degrees, the <span class="hlt">airglow</span> intensity decreases from it maximum to its minimum. In this transition region the upper atmosphere of Titan changes from being totally sunlit to being in the shadow of the moon. For SZA 100 to 180 degrees, we <span class="hlt">observe</span> a constant <span class="hlt">airglow</span> intensity close to zero. The behavior of the <span class="hlt">airglow</span> is also similar to the behavior of the electron density as a function of the SZA as <span class="hlt">observed</span> by Ågren at al (2009). Both variables exhibit a decrease intensity with increasing SZA. The goal of this study is to understand such correlation. We demonstrate the importance of the solar XUV photons contribution to the Titan <span class="hlt">airglow</span> and prove that the strongest contribution to the Titan dayglow occurs by solar fluorescence rather than the particle impact that predominates at night.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.4177U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.4177U"><span>Measuring FeO variation using astronomical spectroscopic <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>Unterguggenberger, Stefanie; Noll, Stefan; Feng, Wuhu; Plane, John M. C.; Kausch, Wolfgang; Kimeswenger, Stefan; Jones, Amy; Moehler, Sabine</p> <p>2017-03-01</p> <p><span class="hlt">Airglow</span> emission lines of OH, O2, O and Na are commonly used to probe the MLT (mesosphere-lower thermosphere) region of the atmosphere. Furthermore, molecules like electronically excited NO, NiO and FeO emit a (pseudo-) continuum. These continua are harder to investigate than atomic emission lines. So far, limb-sounding from space and a small number of ground-based low-to-medium resolution spectra have been used to measure FeO emission in the MLT. In this study the medium-to-high resolution echelle spectrograph X-shooter at the Very Large Telescope (VLT) in the Chilean Atacama Desert (24°37' S, 70°24' W; 2635 m) is used to study the FeO pseudo-continuum in the range from 0.5 to 0.72 µm based on 3662 spectra. Variations of the FeO spectrum itself, as well as the diurnal and seasonal behaviour of the FeO and Na emission intensities, are reported. These <span class="hlt">airglow</span> emissions are linked by their common origin, meteoric ablation, and they share O3 as a common reactant. Major differences are found in the main emission peak of the FeO <span class="hlt">airglow</span> spectrum between 0.58 and 0.61 µm, compared with a theoretical spectrum. The FeO and Na <span class="hlt">airglow</span> intensities exhibit a similar nocturnal variation and a semi-annual seasonal variation with equinoctial maxima. This is satisfactorily reproduced by a whole atmosphere chemistry climate model, if the quantum yields for the reactions of Fe and Na with O3 are 13 ± 3 and 11 ± 2 % respectively. However, a comparison between the modelled O3 in the upper mesosphere and measurements of O3 made with the SABER satellite instrument suggests that these quantum yields may be a factor of ˜ 2 smaller.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018GeoRL..45...15G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018GeoRL..45...15G"><span>Predicting Electron Population Characteristics in 2-D Using Multispectral Ground-Based <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grubbs, Guy; Michell, Robert; Samara, Marilia; Hampton, Donald; Jahn, Jorg-Micha</p> <p>2018-01-01</p> <p>Ground-based <span class="hlt">imaging</span> and in situ sounding rocket data are compared to electron transport modeling for an active inverted-V type auroral event. The Ground-to-Rocket Electrodynamics-Electrons Correlative Experiment (GREECE) mission successfully launched from Poker Flat, Alaska, on 3 March 2014 at 11:09:50 UT and reached an apogee of approximately 335 km over the aurora. Multiple ground-based electron-multiplying charge-coupled device (EMCCD) <span class="hlt">imagers</span> were positioned at Venetie, Alaska, and aimed toward magnetic zenith. The <span class="hlt">imagers</span> <span class="hlt">observed</span> the intensity of different auroral emission lines (427.8, 557.7, and 844.6 nm) at the magnetic foot point of the rocket payload. Emission line intensity data are correlated with electron characteristics measured by the GREECE onboard electron spectrometer. A modified version of the GLobal <span class="hlt">airglOW</span> (GLOW) model is used to estimate precipitating electron characteristics based on optical emissions. GLOW predicted the electron population characteristics with 20% error given the <span class="hlt">observed</span> spectral intensities within 10° of magnetic zenith. Predictions are within 30% of the actual values within 20° of magnetic zenith for inverted-V-type aurora. Therefore, it is argued that this technique can be used, at least in certain types of aurora, such as the inverted-V type presented here, to derive 2-D maps of electron characteristics. These can then be used to further derive 2-D maps of ionospheric parameters as a function of time, based solely on multispectral optical <span class="hlt">imaging</span> data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930047925&hterms=iso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Diso','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930047925&hterms=iso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Diso"><span>Mesospheric nightglow spectral survey taken by the ISO spectral spatial <span class="hlt">imager</span> on ATLAS 1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Owens, J. K.; Torr, D. G.; Torr, M. R.; Chang, T.; Fennelly, J. A.; Richards, P. G.; Morgan, M. F.; Baldridge, T. W.; Fellows, C. W.; Dougani, H.</p> <p>1993-01-01</p> <p>This paper reports the first comprehensive spectral survey of the mesospheric <span class="hlt">airglow</span> between 260 and 832 nm taken by the <span class="hlt">Imaging</span> Spectrometric Observatory on the ATLAS 1 mission. We select data taken in the spectral window between 275 and 300 nm to determine the variation with altitude of the Herzberg I bands originating from the vibrational levels v-prime = 3 to 8. These data provide the first spatially resolved spectral measurements of the system. The data are used to demonstrate that to within an uncertainty of +/- 10 percent, the vibrational distribution remains invariant with altitude. The deficit reported previously for the v-prime = 5 level is not <span class="hlt">observed</span> although there is a suggestion of depletion in v-prime = 6. The data could be used to place tight constraints on the vibrational dependence of quenching rate coefficients, and on the abundance of atomic oxygen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19960029055&hterms=iso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Diso','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19960029055&hterms=iso&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Diso"><span>Mesospheric nightglow spectral survey taken by the ISO spectral spatial <span class="hlt">imager</span> on Atlas 1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Owens, J. K.; Torr, D. G.; Torr, M. R.; Chang, T.; Fennelly, J. A.; Richards, P. G.; Morgan, M. F.; Baldridge, T. W.; Dougani, H.; Swift, W.</p> <p>1993-01-01</p> <p>This paper reports the first comprehensive spectral survey of the mesospheric <span class="hlt">airglow</span> between 260 and 832 nm taken by the <span class="hlt">Imaging</span> Spectrometric Observatory (ISO) on the ATLAS I mission. We select data taken in the spectral window between 275 and 300 nm to determine the variation with altitude of the Herzberg I bands originating from the vibrational levels v' = 3 to 8. These data provide the first spatially resolved spectral measurements of the system. The data are used to demonstrate that to within an uncertainty of + 10%, the vibrational distribution remains invariant with altitude. The deficit reported previously for the v' = 5 level is not <span class="hlt">observed</span> although there is a suggestion of depletion in v' = 6. The data could be used to place tight constraints on the vibrational dependence of quenching rate coefficients, and on the abundance of atomic oxygen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SPIE.7710E..0NB','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SPIE.7710E..0NB"><span>Dynamic <span class="hlt">image</span> fusion and general <span class="hlt">observer</span> preference</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burks, Stephen D.; Doe, Joshua M.</p> <p>2010-04-01</p> <p>Recent developments in <span class="hlt">image</span> fusion give the user community many options for ways of presenting the imagery to an end-user. Individuals at the US Army RDECOM CERDEC Night Vision and Electronic Sensors Directorate have developed an electronic system that allows users to quickly and efficiently determine optimal <span class="hlt">image</span> fusion algorithms and color parameters based upon collected imagery and videos from environments that are typical to <span class="hlt">observers</span> in a military environment. After performing multiple multi-band data collections in a variety of military-like scenarios, different waveband, fusion algorithm, <span class="hlt">image</span> post-processing, and color choices are presented to <span class="hlt">observers</span> as an output of the fusion system. The <span class="hlt">observer</span> preferences can give guidelines as to how specific scenarios should affect the presentation of fused imagery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1611013J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1611013J"><span>A new technique for measuring aerosols with moonlight <span class="hlt">observations</span> and a sky background model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jones, Amy; Noll, Stefan; Kausch, Wolfgang; Kimeswenger, Stefan; Szyszka, Ceszary; Unterguggenberger, Stefanie</p> <p>2014-05-01</p> <p>There have been an ample number of studies on aerosols in urban, daylight conditions, but few for remote, nocturnal aerosols. We have developed a new technique for investigating such aerosols using our sky background model and astronomical <span class="hlt">observations</span>. With a dedicated <span class="hlt">observing</span> proposal we have successfully tested this technique for nocturnal, remote aerosol studies. This technique relies on three requirements: (a) sky background model, (b) <span class="hlt">observations</span> taken with scattered moonlight, and (c) spectrophotometric standard star <span class="hlt">observations</span> for flux calibrations. The sky background model was developed for the European Southern Observatory and is optimized for the Very Large Telescope at Cerro Paranal in the Atacama desert in Chile. This is a remote location with almost no urban aerosols. It is well suited for studying remote background aerosols that are normally difficult to detect. Our sky background model has an uncertainty of around 20 percent and the scattered moonlight portion is even more accurate. The last two requirements are having astronomical <span class="hlt">observations</span> with moonlight and of standard stars at different airmasses, all during the same night. We had a dedicated <span class="hlt">observing</span> proposal at Cerro Paranal with the instrument X-Shooter to use as a case study for this method. X-Shooter is a medium resolution, echelle spectrograph which covers the wavelengths from 0.3 to 2.5 micrometers. We <span class="hlt">observed</span> plain sky at six different distances (7, 13, 20, 45, 90, and 110 degrees) to the Moon for three different Moon phases (between full and half). Also direct <span class="hlt">observations</span> of spectrophotometric standard stars were taken at two different airmasses for each night to measure the extinction curve via the Langley method. This is an ideal data set for testing this technique. The underlying assumption is that all components, other than the atmospheric conditions (specifically aerosols and <span class="hlt">airglow</span>), can be calculated with the model for the given <span class="hlt">observing</span> parameters. The scattered</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/2016cosp...41E2022V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016cosp...41E2022V"><span>First retrievals of MLT sodium profiles based on satellite sodium nightglow <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>Von Savigny, Christian; Zilker, Bianca; Langowski, Martin</p> <p>2016-07-01</p> <p>The Na D lines are a well known feature of the terrestrial <span class="hlt">airglow</span> and have been identified for the first time in 1929. During the daytime the Na <span class="hlt">airglow</span> emission is caused by resonance fluorescence, while during the night the excitation occurs by chemiluminescent reactions. Knowledge of Na in the mesopause region is of interest, because the Na layer is thought to be maintained by meteoric ablation and Na measurements allow constraining the meteoric mass influx into the Earth system. In this contribution we employ SCIAMACHY/Envisat nighttime limb measurements of the Na D-line <span class="hlt">airglow</span> from fall 2002 to spring 2012 - in combination with photochemical models - in order to retrieve Na concentration profiles in the 75 - 100 km altitude range. The Na profiles show realistic peak altitudes, number densities and seasonal variations. The retrieval scheme, sample results and comparisons to ground-based LIDAR measurements of Na as well as SCIAMACHY daytime retrievals will be presented. Moreover, uncertainties in the assumed photochemical scheme and their impact on the Na retrievals will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA43B2658S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA43B2658S"><span>Improved Background Removal in Sounding Rocket Neutral Atom <span class="hlt">Imaging</span> Data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Smith, M. R.; Rowland, D. E.</p> <p>2017-12-01</p> <p>The VISIONS sounding rocket, launched into a substorm on Feb 7, 2013 from Poker Flat, Alaska had a novel miniaturized energetic neutral atom (ENA) <span class="hlt">imager</span> onboard. We present further analysis of the ENA data from this rocket flight, including improved removal of ultraviolet and electron contamination. In particular, the relative error source contributions due to geocoronal, auroral, and <span class="hlt">airglow</span> UV, as well as energetic electrons from 10 eV to 3 keV were assessed. The resulting data provide a more clear understanding of the spatial and temporal variations of the ion populations that are energized to tens or hundreds of eV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080015493','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080015493"><span>SABER <span class="hlt">Observations</span> of the OH Meinel <span class="hlt">Airglow</span> Variability Near the Mesopause</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Marsh, Daniel R.; Smith, Anne K.; Mlynczak, Martin G.</p> <p>2005-01-01</p> <p>The Sounding of the Atmosphere using Broadband Emission Radiometry (SABER) instrument, one of four on board the TIMED satellite, <span class="hlt">observes</span> the OH Meinel emission at 2.0 m that peaks near the mesopause. The emission results from reactions between members of the oxygen and hydrogen chemical families that can be significantly affected by mesopause dynamics. In this study we compare SABER measurements of OH Meinel emission rates and temperatures with predictions from a 3-dimensional chemical dynamical model. In general, the model is capable of reproducing both the <span class="hlt">observed</span> diurnal and seasonal OH Meinel emission variability. The results indicate that the diurnal tide has a large effect on the overall magnitude and temporal variation of the emission in low latitudes. This tidal variability is so dominant that the seasonal cycle in the nighttime emission depends very strongly on the local time of the analysis. At higher latitudes, the emission has an annual cycle that is due mainly to transport of oxygen by the seasonally reversing mean circulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030068415&hterms=Plasma+Ring&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPlasma%2BRing','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030068415&hterms=Plasma+Ring&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPlasma%2BRing"><span><span class="hlt">IMAGE</span> <span class="hlt">Observations</span> of Plasmasphere/Ring Current Interactions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gallagher, D. L.; Adrian, M. L.; Perez, J.; Sandel, B. R.</p> <p>2003-01-01</p> <p>Evidence has been found in <span class="hlt">IMAGE</span> <span class="hlt">observations</span> that overlap of the plasmasphere and the ring current may lead to enhanced loss of plasma into the ionosphere. It has long been anticipated that this mixing of plasma leads to coupling and resulting consequences on both populations. Wave generation, pitch angle scattering, and heating are some of the consequences that are anticipated. <span class="hlt">IMAGE</span> plasmasphere ring current, and auroral <span class="hlt">observations</span> will be presented and used to explore these interactions and their effects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..1910580R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..1910580R"><span>Magnetospheric particle precipitation at Titan</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Royer, Emilie; Esposito, Larry; Crary, Frank; Wahlund, Jan-Erik</p> <p>2017-04-01</p> <p>Although solar XUV radiation is known to be the main source of ionization in Titan's upper atmosphere around 1100 km of altitude, magnetospheric particle precipitation can also account for about 10% of the ionization process. Magnetospheric particle precipitation is expected to be the most intense on the nightside of the satelllite and when Titan's orbital position around Saturn is the closest to Noon Saturn Local Time (SLT). In addition, on several occasion throughout the Cassini mission, Titan has been <span class="hlt">observed</span> while in the magnetosheath. We are reporting here Ultraviolet (UV) <span class="hlt">observations</span> of Titan <span class="hlt">airglow</span> enhancements correlated to these magnetospheric changing conditions occurring while the spacecraft, and thus Titan, are known to have crossed Saturn's magnetopause and have been exposed to the magnetosheath environnment. Using Cassini-Ultraviolet <span class="hlt">Imaging</span> Spectrograph (UVIS) <span class="hlt">observations</span> of Titan around 12PM SLT as our primary set of data, we present evidence of Titan's upper atmosphere response to a fluctuating magnetospheric environment. Pattern recognition software based on 2D UVIS detector <span class="hlt">images</span> has been used to retrieve <span class="hlt">observations</span> of interest, looking for <span class="hlt">airglow</span> enhancement of a factor of 2. A 2D UVIS detector <span class="hlt">image</span>, created for each UVIS <span class="hlt">observation</span> of Titan, displays the spatial dimension of the UVIS slit on the x-axis and the time on the y-axis. In addition, data from the T32 flyby and from April 17, 2005 from in-situ Cassini instruments are used. Correlations with data from simultaneous <span class="hlt">observations</span> of in-situ Cassini instruments (CAPS, RPWS and MIMI) has been possible on few occasions and events such as electron burst and reconnections can be associated with unusual behaviors of the Titan <span class="hlt">airglow</span>. CAPS in-situ measurements acquired during the T32 flyby are consistent with an electron burst <span class="hlt">observed</span> at the spacecraft as the cause of the UV emission. Moreover, on April 17, 2005 the UVIS <span class="hlt">observation</span> displays feature similar to what could be a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23453000','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23453000"><span>Quantification of heterogeneity <span class="hlt">observed</span> in medical <span class="hlt">images</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brooks, Frank J; Grigsby, Perry W</p> <p>2013-03-02</p> <p>There has been much recent interest in the quantification of visually evident heterogeneity within functional grayscale medical <span class="hlt">images</span>, such as those obtained via magnetic resonance or positron emission tomography. In the case of <span class="hlt">images</span> of cancerous tumors, variations in grayscale intensity imply variations in crucial tumor biology. Despite these considerable clinical implications, there is as yet no standardized method for measuring the heterogeneity <span class="hlt">observed</span> via these <span class="hlt">imaging</span> modalities. In this work, we motivate and derive a statistical measure of <span class="hlt">image</span> heterogeneity. This statistic measures the distance-dependent average deviation from the smoothest intensity gradation feasible. We show how this statistic may be used to automatically rank <span class="hlt">images</span> of in vivo human tumors in order of increasing heterogeneity. We test this method against the current practice of ranking <span class="hlt">images</span> via expert visual inspection. We find that this statistic provides a means of heterogeneity quantification beyond that given by other statistics traditionally used for the same purpose. We demonstrate the effect of tumor shape upon our ranking method and find the method applicable to a wide variety of clinically relevant tumor <span class="hlt">images</span>. We find that the automated heterogeneity rankings agree very closely with those performed visually by experts. These results indicate that our automated method may be used reliably to rank, in order of increasing heterogeneity, tumor <span class="hlt">images</span> whether or not object shape is considered to contribute to that heterogeneity. Automated heterogeneity ranking yields objective results which are more consistent than visual rankings. Reducing variability in <span class="hlt">image</span> interpretation will enable more researchers to better study potential clinical implications of <span class="hlt">observed</span> tumor heterogeneity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2859675','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2859675"><span>Spatio-temporal Hotelling <span class="hlt">observer</span> for signal detection from <span class="hlt">image</span> sequences</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Caucci, Luca; Barrett, Harrison H.; Rodríguez, Jeffrey J.</p> <p>2010-01-01</p> <p>Detection of signals in noisy <span class="hlt">images</span> is necessary in many applications, including astronomy and medical <span class="hlt">imaging</span>. The optimal linear <span class="hlt">observer</span> for performing a detection task, called the Hotelling <span class="hlt">observer</span> in the medical literature, can be regarded as a generalization of the familiar prewhitening matched filter. Performance on the detection task is limited by randomness in the <span class="hlt">image</span> data, which stems from randomness in the object, randomness in the <span class="hlt">imaging</span> system, and randomness in the detector outputs due to photon and readout noise, and the Hotelling <span class="hlt">observer</span> accounts for all of these effects in an optimal way. If multiple temporal frames of <span class="hlt">images</span> are acquired, the resulting data set is a spatio-temporal random process, and the Hotelling <span class="hlt">observer</span> becomes a spatio-temporal linear operator. This paper discusses the theory of the spatio-temporal Hotelling <span class="hlt">observer</span> and estimation of the required spatio-temporal covariance matrices. It also presents a parallel implementation of the <span class="hlt">observer</span> on a cluster of Sony PLAYSTATION 3 gaming consoles. As an example, we consider the use of the spatio-temporal Hotelling <span class="hlt">observer</span> for exoplanet detection. PMID:19550494</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19550494','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19550494"><span>Spatio-temporal Hotelling <span class="hlt">observer</span> for signal detection from <span class="hlt">image</span> sequences.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Caucci, Luca; Barrett, Harrison H; Rodriguez, Jeffrey J</p> <p>2009-06-22</p> <p>Detection of signals in noisy <span class="hlt">images</span> is necessary in many applications, including astronomy and medical <span class="hlt">imaging</span>. The optimal linear <span class="hlt">observer</span> for performing a detection task, called the Hotelling <span class="hlt">observer</span> in the medical literature, can be regarded as a generalization of the familiar prewhitening matched filter. Performance on the detection task is limited by randomness in the <span class="hlt">image</span> data, which stems from randomness in the object, randomness in the <span class="hlt">imaging</span> system, and randomness in the detector outputs due to photon and readout noise, and the Hotelling <span class="hlt">observer</span> accounts for all of these effects in an optimal way. If multiple temporal frames of <span class="hlt">images</span> are acquired, the resulting data set is a spatio-temporal random process, and the Hotelling <span class="hlt">observer</span> becomes a spatio-temporal linear operator. This paper discusses the theory of the spatio-temporal Hotelling <span class="hlt">observer</span> and estimation of the required spatio-temporal covariance matrices. It also presents a parallel implementation of the <span class="hlt">observer</span> on a cluster of Sony PLAYSTATION 3 gaming consoles. As an example, we consider the use of the spatio-temporal Hotelling <span class="hlt">observer</span> for exoplanet detection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080037618&hterms=negev+radiation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnegev%2Bradiation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080037618&hterms=negev+radiation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnegev%2Bradiation"><span>Plasmaspheric Plumes <span class="hlt">Observed</span> by the CLUSTER and <span class="hlt">IMAGE</span> Spacecraft</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fung, S. F.; Benson, R. F.; Garcia, L. N.; Adrian, M. L.; Sandel, B.; Goldstein, M. L.</p> <p>2008-01-01</p> <p>Global <span class="hlt">IMAGE</span>/EUV <span class="hlt">observations</span> have revealed complex changes in plasmaspheric structures as the plasmasphere responds to geomagnetic activity while remaining under varying degrees of influence by co-rotation, depending on the radial distance. The complex plasmaspheric dynamics, with different scales of variability, is clearly far from being well understood. There is now renewed interest in the plasmasphere due to its apparent connections with the development of the ring current and radiation belt, and loss of ionospheric plasmas. Early in the mission, the Cluster spacecraft only crossed the plasmapause (L - 4) occasionally and made measurements of the outer plasmasphere and plasmaspheric drainage plumes. The study by Darrouzet et al. [2006] provided detailed analyses of in situ Cluster <span class="hlt">observations</span> and <span class="hlt">IMAGE</span> EUV <span class="hlt">observations</span> of three plasmaspheric plumes detected in April-June, 2002. Within the next couple of years, Cluster orbit will change, causing perigee to migrate to lower altitudes, and thus providing excellent opportunities to obtain more detailed measurements of the plasmasphere. In this paper, we report our analyses of the earlier Cluster-<span class="hlt">IMAGE</span> events by incorporating the different perspectives provided by the <span class="hlt">IMAGE</span> Radio Plasma <span class="hlt">Imager</span> (RPI) <span class="hlt">observations</span>. We will discuss our new understanding of the structure and dynamics of the Cluster-<span class="hlt">IMAGE</span> events.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17730507','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17730507"><span>One Mars year: viking lander <span class="hlt">imaging</span> <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>Jones, K L; Arvidson, R E; Guinness, E A; Bragg, S L; Wall, S D; Carlston, C E; Pidek, D G</p> <p>1979-05-25</p> <p>Throughout the complete Mars year during which they have been on the planet, the <span class="hlt">imaging</span> systems aboard the two Viking landers have documented a variety of surface changes. Surface condensates, consisting of both solid H(2)O and CO(2), formed at the Viking 2 lander site during the winter. Additional <span class="hlt">observations</span> suggest that surface erosion rates due to dust redistribution may be substantially less than those predicted on the basis of pre-Viking <span class="hlt">observations</span>. The Viking 1 lander will continue to acquire and transmit a predetermined sequence of <span class="hlt">imaging</span> and meteorology data as long as it is operative.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22220517-correlation-between-model-observer-human-observer-performance-ct-imaging-when-lesion-location-uncertain','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22220517-correlation-between-model-observer-human-observer-performance-ct-imaging-when-lesion-location-uncertain"><span>Correlation between model <span class="hlt">observer</span> and human <span class="hlt">observer</span> performance in CT <span class="hlt">imaging</span> when lesion location is uncertain</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>Leng, Shuai; Yu, Lifeng; Zhang, Yi</p> <p>2013-08-15</p> <p>Purpose: The purpose of this study was to investigate the correlation between model <span class="hlt">observer</span> and human <span class="hlt">observer</span> performance in CT <span class="hlt">imaging</span> for the task of lesion detection and localization when the lesion location is uncertain.Methods: Two cylindrical rods (3-mm and 5-mm diameters) were placed in a 35 × 26 cm torso-shaped water phantom to simulate lesions with −15 HU contrast at 120 kV. The phantom was scanned 100 times on a 128-slice CT scanner at each of four dose levels (CTDIvol = 5.7, 11.4, 17.1, and 22.8 mGy). Regions of interest (ROIs) around each lesion were extracted to generate imagesmore » with signal-present, with each ROI containing 128 × 128 pixels. Corresponding ROIs of signal-absent <span class="hlt">images</span> were generated from <span class="hlt">images</span> without lesion mimicking rods. The location of the lesion (rod) in each ROI was randomly distributed by moving the ROIs around each lesion. Human <span class="hlt">observer</span> studies were performed by having three trained <span class="hlt">observers</span> identify the presence or absence of lesions, indicating the lesion location in each <span class="hlt">image</span> and scoring confidence for the detection task on a 6-point scale. The same <span class="hlt">image</span> data were analyzed using a channelized Hotelling model <span class="hlt">observer</span> (CHO) with Gabor channels. Internal noise was added to the decision variables for the model <span class="hlt">observer</span> study. Area under the curve (AUC) of ROC and localization ROC (LROC) curves were calculated using a nonparametric approach. The Spearman's rank order correlation between the average performance of the human <span class="hlt">observers</span> and the model <span class="hlt">observer</span> performance was calculated for the AUC of both ROC and LROC curves for both the 3- and 5-mm diameter lesions.Results: In both ROC and LROC analyses, AUC values for the model <span class="hlt">observer</span> agreed well with the average values across the three human <span class="hlt">observers</span>. The Spearman's rank order correlation values for both ROC and LROC analyses for both the 3- and 5-mm diameter lesions were all 1.0, indicating perfect rank ordering agreement of the figures of merit (AUC) between</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5606269','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5606269"><span>A new method to evaluate <span class="hlt">image</span> quality of CBCT <span class="hlt">images</span> quantitatively without <span class="hlt">observers</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>Shimizu, Mayumi; Okamura, Kazutoshi; Yoshida, Shoko; Weerawanich, Warangkana; Tokumori, Kenji; Jasa, Gainer R; Yoshiura, Kazunori</p> <p>2017-01-01</p> <p>Objectives: To develop an <span class="hlt">observer</span>-free method for quantitatively evaluating the <span class="hlt">image</span> quality of CBCT <span class="hlt">images</span> by applying just-noticeable difference (JND). Methods: We used two test objects: (1) a Teflon (polytetrafluoroethylene) plate phantom attached to a dry human mandible; and (2) a block phantom consisting of a Teflon step phantom and an aluminium step phantom. These phantoms had holes with different depths. They were immersed in water and scanned with a CB MercuRay (Hitachi Medical Corporation, Tokyo, Japan) at tube voltages of 120 kV, 100 kV, 80 kV and 60 kV. Superimposed <span class="hlt">images</span> of the phantoms with holes were used for evaluation. The number of detectable holes was used as an index of <span class="hlt">image</span> quality. In detecting holes quantitatively, the threshold grey value (ΔG), which differentiated holes from the background, was calculated using a specific threshold (the JND), and we extracted the holes with grey values above ΔG. The indices obtained by this quantitative method (the extracted hole values) were compared with the <span class="hlt">observer</span> evaluations (the <span class="hlt">observed</span> hole values). In addition, the contrast-to-noise ratio (CNR) of the shallowest detectable holes and the deepest undetectable holes were measured to evaluate the contribution of CNR to detectability. Results: The results of this evaluation method corresponded almost exactly with the evaluations made by <span class="hlt">observers</span>. The extracted hole values reflected the influence of different tube voltages. All extracted holes had an area with a CNR of ≥1.5. Conclusions: This quantitative method of evaluating CBCT <span class="hlt">image</span> quality may be more useful and less time-consuming than evaluation by <span class="hlt">observation</span>. PMID:28045343</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA00739&hterms=Dark+web&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DDark%2Bweb','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA00739&hterms=Dark+web&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DDark%2Bweb"><span>Eclipse <span class="hlt">Images</span> of Io (3 views)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1997-01-01</p> <p>These three <span class="hlt">images</span> of Io in eclipse (top) show volcanic hot spots and <span class="hlt">airglow</span> associated with volcanic plumes and Io's atmosphere. They were acquired by NASA's Galileo spacecraft during three separate orbits of Jupiter when the moon was in Jupiter's shadow. Brightnesses are color-coded from red which displays the highest intensity to dark blue which displays zero intensity (no light).<p/>Below them are the corresponding views of Io in reflected sunlight, reprojected from a global mosaic of <span class="hlt">images</span> obtained during Galileo's first and second orbits of Jupiter. These lit views help to identify the locations of the hot spots seen in the eclipse <span class="hlt">images</span>. The grid marks are at 15 degree intervals of latitude and longitude. North is to the top.<p/>In the eclipse <span class="hlt">images</span> (top) small red ovals and perhaps some small green areas are due to thermal emission from volcanic hot spots with temperatures hotter than about 700 kelvin (about 1000 degrees Fahrenheit). Diffuse greenish areas seen near the limb or edge of the moon are probably the result of auroral and/or <span class="hlt">airglow</span> emissions of neutral species of oxygen or sulfur in volcanic plumes and in Io's patchy atmosphere.<p/>All <span class="hlt">images</span> were acquired by the solid state <span class="hlt">imaging</span> (CCD) system on NASA's Galileo spacecraft. The top left <span class="hlt">image</span> was obtained during the spacecraft's fourth orbit (E4) on December 17, 1996, the top middle <span class="hlt">image</span> during the sixth orbit (E6) on February 21, 1997, and the top right <span class="hlt">image</span> during the first orbit (G1) on June 29th, 1996. The relatively long exposures used to obtain these eclipse <span class="hlt">images</span> lead to some smearing of the picture elements which reduces the actual resolution. Unsmeared they would have resolutions of 17.6, 9.1, and 10.5 kilometers per picture element respectively (left to right).<p/>The Jet Propulsion Laboratory, Pasadena, CA manages the Galileo mission for NASA's Office of Space Science, Washington, DC. JPL is an operating division of California Institute of Technology (Caltech</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSH23B2442P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSH23B2442P"><span>Bayesian Analysis of Hmi <span class="hlt">Images</span> and Comparison to Tsi Variations and MWO <span class="hlt">Image</span> <span class="hlt">Observables</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Parker, D. G.; Ulrich, R. K.; Beck, J.; Tran, T. V.</p> <p>2015-12-01</p> <p>We have previously applied the Bayesian automatic classification system AutoClass to solar magnetogram and intensity <span class="hlt">images</span> from the 150 Foot Solar Tower at Mount Wilson to identify classes of solar surface features associated with variations in total solar irradiance (TSI) and, using those identifications, modeled TSI time series with improved accuracy (r > 0.96). (Ulrich, et al, 2010) AutoClass identifies classes by a two-step process in which it: (1) finds, without human supervision, a set of class definitions based on specified attributes of a sample of the <span class="hlt">image</span> data pixels, such as magnetic field and intensity in the case of MWO <span class="hlt">images</span>, and (2) applies the class definitions thus found to new data sets to identify automatically in them the classes found in the sample set. HMI high resolution <span class="hlt">images</span> capture four <span class="hlt">observables</span>-magnetic field, continuum intensity, line depth and line width-in contrast to MWO's two <span class="hlt">observables</span>-magnetic field and intensity. In this study, we apply AutoClass to the HMI <span class="hlt">observables</span> for <span class="hlt">images</span> from June, 2010 to December, 2014 to identify solar surface feature classes. We use contemporaneous TSI measurements to determine whether and how variations in the HMI classes are related to TSI variations and compare the characteristic statistics of the HMI classes to those found from MWO <span class="hlt">images</span>. We also attempt to derive scale factors between the HMI and MWO magnetic and intensity <span class="hlt">observables</span>.The ability to categorize automatically surface features in the HMI <span class="hlt">images</span> holds out the promise of consistent, relatively quick and manageable analysis of the large quantity of data available in these <span class="hlt">images</span>. Given that the classes found in MWO <span class="hlt">images</span> using AutoClass have been found to improve modeling of TSI, application of AutoClass to the more complex HMI <span class="hlt">images</span> should enhance understanding of the physical processes at work in solar surface features and their implications for the solar-terrestrial environment.Ulrich, R.K., Parker, D, Bertello, L. and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988SPIE..914...60L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988SPIE..914...60L"><span>A Computational <span class="hlt">Observer</span> For Performing Contrast-Detail Analysis Of Ultrasound <span class="hlt">Images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lopez, H.; Loew, M. H.</p> <p>1988-06-01</p> <p>Contrast-Detail (C/D) analysis allows the quantitative determination of an <span class="hlt">imaging</span> system's ability to display a range of varying-size targets as a function of contrast. Using this technique, a contrast-detail plot is obtained which can, in theory, be used to compare <span class="hlt">image</span> quality from one <span class="hlt">imaging</span> system to another. The C/D plot, however, is usually obtained by using data from human <span class="hlt">observer</span> readings. We have shown earlier(7) that the performance of human <span class="hlt">observers</span> in the task of threshold detection of simulated lesions embedded in random ultrasound noise is highly inaccurate and non-reproducible for untrained <span class="hlt">observers</span>. We present an objective, computational method for the determination of the C/D curve for ultrasound <span class="hlt">images</span>. This method utilizes digital <span class="hlt">images</span> of the C/D phantom developed at CDRH, and lesion-detection algorithms that simulate the Bayesian approach using the likelihood function for an ideal <span class="hlt">observer</span>. We present the results of this method, and discuss the relationship to the human <span class="hlt">observer</span> and to the comparability of <span class="hlt">image</span> quality between systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/1246636','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/1246636"><span>Pinhole <span class="hlt">imaging</span> in Legg-Perthes disease: further <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>Danigelis, J A</p> <p>1976-01-01</p> <p>Fifty-nine patients with Legg-Perthes disease and 12 others were studied using 99mTc-polyphosphate and the pinhole collimator <span class="hlt">imaging</span> technique. Radiographs of both hips were correlated with <span class="hlt">images</span> in each patient. In the Legg-Perthes patients, a radionuclide uptake deficiency of variable size was <span class="hlt">observed</span> in the proximal femoral epiphysis (EOC), which we believe is related to varying degrees of impaired blood supply. During later disease stages, adjacent zones of increased radionuclide activity of revascularization were <span class="hlt">observed</span> that would replace the uptake defect eventually. Unless radiographic evidence of new bone formation was <span class="hlt">observed</span> in the EOC, it was impossible to predict either the presence or extent of revascularization until bone <span class="hlt">imaging</span> was done. Those patients with revascularization activity in the EOC exhibited a relatively short time interval (average, 3.2 months) before evidence of new bone formation radiographically. Others with increased radionuclide concentration limited to the growth plate and/or metaphysis averaged a much longer 7.8 months. In two patients there was a reversal of the initially increased activity in the growth plate, suggesting another vascular insult. There were no false-negative bone-<span class="hlt">image</span> findings in the 12 cases that clinically and/or radiologically simulated Legg-Perthes disease. Our <span class="hlt">image</span> studies correlate well with published histopathologic investigations, indicating to us that assessment of extent of pathologic involvement and of the disease course is facilitated by this technique. Subsequently, this could influence treatment selection and provide a more objective baseline from which to judge treatment results. Continued experience suggests pinhole bone <span class="hlt">imaging</span> has useful clinical application in Legg-Perthes disease and other childhood hip disorders.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRA..122.7504D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.7504D"><span>Coordinated <span class="hlt">observations</span> of postmidnight irregularities and thermospheric neutral winds and temperatures 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>Dao, Tam; Otsuka, Yuichi; Shiokawa, Kazuo; Nishioka, Michi; Yamamoto, Mamoru; Buhari, Suhaila M.; Abdullah, Mardina; Husin, Asnawi</p> <p>2017-07-01</p> <p>We investigated a postmidnight field-aligned irregularity (FAI) event <span class="hlt">observed</span> with the Equatorial Atmosphere Radar at Kototabang (0.2°S, 100.3°E, dip latitude 10.4°S) in Indonesia on the night of 9 July 2010 using a comprehensive data set of both neutral and plasma parameters. We examined the rate of total electron content change index (ROTI) obtained from GPS receivers in Southeast Asia, <span class="hlt">airglow</span> <span class="hlt">images</span> detected by an all-sky <span class="hlt">imager</span>, and thermospheric neutral winds and temperatures obtained by a Fabry-Perot interferometer at Kototabang. Altitudes of the F layer (h'F) <span class="hlt">observed</span> by ionosondes at Kototabang, Chiang Mai, and Chumphon were also surveyed. We found that the postmidnight FAIs occurred within plasma bubbles and coincided with kilometer-scale plasma density irregularities. We also <span class="hlt">observed</span> an enhancement of the magnetically equatorward thermospheric neutral wind at the same time as the increase of h'F at low-latitude stations, but h'F at a station near the magnetic equator remained invariant. Simultaneously, a magnetically equatorward gradient of thermospheric temperature was identified at Kototabang. The convergence of equatorward neutral winds from the Northern and Southern Hemispheres could be associated with a midnight temperature maximum occurring around the magnetic equator. Equatorward neutral winds can uplift the F layer at low latitudes and increase the growth rate of Rayleigh-Taylor instabilities, causing more rapid extension of plasma bubbles. The equatorward winds in both hemispheres also intensify the eastward Pedersen current, so a large polarization electric field generated in the plasma bubble might play an important role in the generation of postmidnight FAIs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10566E..26B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10566E..26B"><span>ALISEO on MIOSat: an <span class="hlt">imaging</span> interferometer for earth <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barducci, A.; Castagnoli, F.; Castellini, G.; Guzzi, D.; Marcoionni, P.; Pippi, I.</p> <p>2017-11-01</p> <p>The Italian Space Agency (ASI) decided to perform an low cost Earth <span class="hlt">observation</span> mission based on a new mini satellite named MIOsat which will carry various technological payloads. Among them an <span class="hlt">imaging</span> interferometer designed and now ready to be assembled and tested by our Institute. The instrument, named ALISEO (Aerospace Leap-frog <span class="hlt">Imaging</span> Stationary interferometer for Earth <span class="hlt">Observation</span>), operates in the common-path Sagnac configuration, and it does not utilize any moving part to scan the phase delays between the two interfering beams. The sensor acquires target <span class="hlt">images</span> modulated by a pattern of autocorrelation functions of the energy coming from each scene pixel, and the resulting fringe pattern remains spatially fixed with respect to the instrument's field-of-view. The complete interferogram of each target location is retrieved by introducing a relative source-<span class="hlt">observer</span> motion, which allows any <span class="hlt">image</span> pixels to be <span class="hlt">observed</span> under different viewing-angles and experience discrete path differences. The paper describes the main characteristics of the <span class="hlt">imaging</span> interferometer as well as the overall optical configuration and the electronics layout. Moreover some theoretical issues concerning sampling theory in "common path" <span class="hlt">imaging</span> interferometry are investigated. The experimental activity performed in laboratory is presented and its outcomes are analysed. Particularly, a set of measurements has been carried out using both standard (certificate) reflectance tiles and natural samples of different volcanic rocks. An algorithm for raw data pre-processing aimed at retrieving the at-sensor radiance spectrum is introduced and its performance is addressed by taking into account various issues such as dark signal subtraction, spectral instrument response compensation, effects of vignetting, and Fourier backtransform. Finally, examples of retrieved absolute reflectance of several samples are sketched at different wavelengths.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950017430','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950017430"><span>Field <span class="hlt">observations</span> using an AOTF polarimetric <span class="hlt">imaging</span> spectrometer</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cheng, Li-Jen; Hamilton, Mike; Mahoney, Colin; Reyes, George</p> <p>1993-01-01</p> <p>This paper reports preliminary results of recent field <span class="hlt">observations</span> using a prototype acousto-optic tunable filter (AOTF) polarimetric <span class="hlt">imaging</span> spectrometer. The data illustrate application potentials for geoscience. The operation principle of this instrument is different from that of current airborne multispectral <span class="hlt">imaging</span> instruments, such as AVIRIS. The AOTF instrument takes two orthogonally polarized <span class="hlt">images</span> at a desired wavelength at one time, whereas AVIRIS takes a spectrum over a predetermined wavelength range at one pixel at a time and the <span class="hlt">image</span> is constructed later. AVIRIS does not have any polarization measuring capability. The AOTF instrument could be a complement tool to AVIRIS. Polarization measurement is a desired capability for many applications in remote sensing. It is well know that natural light is often polarized due to various scattering phenomena in the atmosphere. Also, scattered light from canopies is reported to have a polarized component. To characterize objects of interest correctly requires a remote sensing <span class="hlt">imaging</span> spectrometer capable of measuring object signal and background radiation in both intensity and polarization so that the characteristics of the object can be determined. The AORF instrument has the capability to do so. The AOTF instrument has other unique properties. For example, it can provide spectral <span class="hlt">images</span> immediately after the <span class="hlt">observation</span>. The instrument can also allow <span class="hlt">observations</span> to be tailored in real time to perform the desired experiments and to collect only required data. Consequently, the performance in each mission can be increased with minimal resources. The prototype instrument was completed in the beginning of this year. A number of outdoor field experiments were performed with the objective to evaluate the capability of this new technology for remote sensing applications and to determine issues for further improvements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5526444','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5526444"><span>NOVEL <span class="hlt">OBSERVATIONS</span> AND POTENTIAL APPLICATIONS USING DIGITAL INFRARED IRIS <span class="hlt">IMAGING</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>Roberts, Daniel K.; Lukic, Ana; Yang, Yongyi; Moroi, Sayoko E.; Wilensky, Jacob T.; Wernick, Miles N.</p> <p>2017-01-01</p> <p>Digital infrared (IR) iris photography using a modified digital camera system was carried out on about 300 subjects seen during routine clinical care and research at one facility. Since this <span class="hlt">image</span> database offered opportunity to gain new insight into the potential utility of IR iris <span class="hlt">imaging</span>, it was surveyed for unique <span class="hlt">image</span> patterns. Then, a selection of photos was compiled that would illustrate the spectrum of this <span class="hlt">imaging</span> experience. Potentially informative <span class="hlt">image</span> patterns were <span class="hlt">observed</span> in subjects with cataracts, diabetic retinopathy, Posner-Schlossman syndrome, iridociliary cysts, long anterior lens zonules, nevi, oculocutaneous albinism, pigment dispersion syndrome, pseudophakia, suspected vascular anomaly, and trauma. <span class="hlt">Image</span> patterns were often unanticipated regardless of pre-existing information and suggest that IR iris <span class="hlt">imaging</span> may have numerous potential clinical and research applications, some of which may still not be recognized. These <span class="hlt">observations</span> suggest further development and study of this technology. PMID:19320317</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/2017AGUFMSA41B2629C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA41B2629C"><span>Greenland Network (GNET) <span class="hlt">observations</span> of Polar cap Patches and Arcs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cesar, V. E.; Pradipta, R.; Pedersen, T.</p> <p>2017-12-01</p> <p>TEC values collected with the Greenland Network (GNET) of GPS/GNSS receivers and 630.0 nm <span class="hlt">airglow</span> emissions recorded with an all-sky <span class="hlt">imager</span> located at Qaanaaq in Greenland are used to investigate the relationship between the appearance and evolution of polar cap patches (PCP) and Sun-aligned arcs (S-AA) and the characteristics of the solar wind. Both PCP and S-AA produce TEC enhancements, but the PCP velocity is 10 times larger than the S-AA's drift. In addition, PCP move anti-sunwardly and the S-AA move in the dawn-dusk direction. We use these properties of PCPs and S-AAs and calculate the velocity of the TEC enhancements to identify and discriminate between patches and arcs. The physical location of the boundary of the polar cap is based on DMSP <span class="hlt">observations</span> of particle precipitation. The IMF and other solar wind parameters are gathered with the ACE satellite that is positioned at the L1 point. Our <span class="hlt">observations</span> indicate that during December 2009, TEC enhancements occur in the polar cap almost every day, but only when the solar wind velocity exceeds 290 km/s. PCPs appear almost immediately after the Bz turns southward; however, the S-AAs develop a few hours after Bz points northward. These conclusions demonstrate the ability of GNET continuous measurements over Greenland to conduct investigations of the formation and evolution of polar cap patches and arcs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/AD1013931','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/AD1013931"><span>Dual Hemisphere Investigations of Ionospheric Irregularities that Disrupt Radio Communications and Navigation</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2016-07-25</p> <p>ignatures of waves that can be used to study both upward and horizonta l ocean-atmosphere co upling. Our primary observat ional technique is optical-one...the first time. 2. Studies of Earthquab and Tsunami-induced Waves in the Ionosphere. One ofthe more spectac.dar uses of all-sky <span class="hlt">airglow</span> <span class="hlt">imaging</span> was...the recent discovery of waves in the ionospheric <span class="hlt">airglow</span> layer caused by the great earthquake and tsunami of 11 March 2011 (Makela et al. , 2011</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910068677&hterms=australian+copyright&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Daustralian%2Bcopyright','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910068677&hterms=australian+copyright&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Daustralian%2Bcopyright"><span>The dark side of Venus - Near-infrared <span class="hlt">images</span> and spectra from the Anglo-Australian Observatory</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Crisp, D.; Allen, D. A.; Grinspoon, D. H.; Pollack, J. B.</p> <p>1991-01-01</p> <p>Near-IR <span class="hlt">images</span> and spectra of the night side of Venus taken at the Anglo-Australian Telescope during February 1990 reveal four new thermal emission windows at 1.10, 1.18, 1.27, and 1.31 microns, in addition to the previously discovered windows at 1.74 and 2.3 microns. <span class="hlt">Images</span> of the Venus night side show similar bright and dark markings in all windows, but their contrast is much lower at short wavelengths. The 1.27-micron window includes a bright, high-altitude O2 <span class="hlt">airglow</span> feature in addition to a thermal contribution from the deep atmosphere. Simulations of the 1.27- and 2.3-micron spectra indicate water vapor mixing ratios near 40 + or - 20 ppm by volume between the surface and the cloud base.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850028663&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DAbreu%252C%2Bc.','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850028663&hterms=Abreu&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DAbreu%252C%2Bc."><span>Tomographic inversion of satellite photometry</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Solomon, S. C.; Hays, P. B.; Abreu, V. J.</p> <p>1984-01-01</p> <p>An inversion algorithm capable of reconstructing the volume emission rate of thermospheric <span class="hlt">airglow</span> features from satellite photometry has been developed. The accuracy and resolution of this technique are investigated using simulated data, and the inversions of several sets of <span class="hlt">observations</span> taken by the Visible <span class="hlt">Airglow</span> Experiment are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040003964&hterms=ghosts&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dghosts','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040003964&hterms=ghosts&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dghosts"><span>Earth <span class="hlt">Observing</span>-1 Advanced Land <span class="hlt">Imager</span>: <span class="hlt">Imaging</span> Performance On-Orbit</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hearn, D. R.</p> <p>2002-01-01</p> <p>This report analyzes the on-orbit <span class="hlt">imaging</span> performance of the Advanced Land <span class="hlt">Imager</span> (ALI) on the Earth <span class="hlt">Observing</span>-1 satellite. The pre-flight calibrations are first summarized. The methods used to reconstruct and geometrically correct the <span class="hlt">image</span> data from this push-broom sensor are described. The method used here does not refer to the position and attitude telemetry from the spacecraft. Rather, it is assumed that the <span class="hlt">image</span> of the scene moves across the focal plane with a constant velocity, which can be ascertained from the <span class="hlt">image</span> data itself. Next, an assortment of the <span class="hlt">images</span> so reconstructed is presented. Color <span class="hlt">images</span> sharpened with the 10-m panchromatic band data are shown, and the algorithm for producing them from the 30-m multispectral data is described. The approach taken for assessing spatial resolution is to compare the sharpness of features in the on-orbit <span class="hlt">image</span> data with profiles predicted on the basis of the pre-flight calibrations. A large assortment of bridge profiles is analyzed, and very good fits to the predicted shapes are obtained. Lunar calibration scans are analyzed to examine the sharpness of the edge-spread function at the limb of the moon. The darkness of the space beyond the limb is better for this purpose than anything that could be simulated on the ground. From these scans, we find clear evidence of scattering in the optical system, as well as some weak ghost <span class="hlt">images</span>. Scans of planets and stars are also analyzed. Stars are useful point sources of light at all wavelengths, and delineate the point-spread functions of the system. From a quarter-speed scan over the Pleiades, we find that the ALI can detect 6th magnitude stars. The quality of the reconstructed <span class="hlt">images</span> verifies the capability of the ALI to produce Landsat-type multi spectral data. The signal-to-noise and panchromatic spatial resolution are considerably superior to those of the existing Landsat sensors. The spatial resolution is confirmed to be as good as it was designed to be.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1986sao..reptQ....N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986sao..reptQ....N"><span>High resolution astrophysical <span class="hlt">observations</span> using speckle <span class="hlt">imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Noyes, R. W.; Nisenson, P.; Papaliolios, C.; Stachnik, R. V.</p> <p>1986-04-01</p> <p>This report describes progress under a contract to develop a complete astronomical speckle <span class="hlt">image</span> reconstruction facility and to apply that facility to the solution of astronomical problems. During the course of the contract we have developed the procedures, algorithms, theory and hardware required to perform that function and have made and interpreted astronomical <span class="hlt">observations</span> of substantial significance. A principal result of the program was development of a photon-counting camera of innovative design, the PAPA detector. Development of this device was, in our view, essential to making the speckle process into a useful astronomical tool, since the principal impediment to that circumstance in the past was the necessity for application of photon noise compensation procedures which were difficult if not impossible to calibrate. The photon camera made this procedure unnecessary and permitted precision <span class="hlt">image</span> recovery. The result of this effort and the associated algorithm development was an active program of astronomical <span class="hlt">observation</span> which included investigations into young stellar objects, supergiant structure and measurements of the helium abundance of the early universe. We have also continued research on recovery of high angular resolution <span class="hlt">images</span> of the solar surface working with scientists at the Sacramento Peak Observatory in this area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940026443','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940026443"><span>Solar measurements from the <span class="hlt">Airglow</span>-Solar Spectrometer Instrument (ASSI) on the San Marco 5 satellite</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Woods, Thomas N.</p> <p>1994-01-01</p> <p>The analysis of the solar spectral irradiance from the <span class="hlt">Airglow</span>-Solar Spectrometer Instrument (ASSI) on the San Marco 5 satellite is the focus for this research grant. A pre-print copy of the paper describing the calibrations of and results from the San Marco ASSI is attached to this report. The calibration of the ASSI included (1) transfer of photometric calibration from a rocket experiment and the Solar Mesosphere Explorer (SME), (2) use of the on-board radioactive calibration sources, (3) validation of the ASSI sensitivity over its field of view, and (4) determining the degradation of the spectrometers. We have determined that the absolute values for the solar irradiance needs adjustment in the current proxy models of the solar UV irradiance, and the amount of solar variability from the proxy models are in reasonable agreement with the ASSI measurements. This research grant also has supported the development of a new solar EUV irradiance proxy model. We expected that the magnetic flux is responsible for most of the heating, via Alfen waves, in the chromosphere, transition region, and corona. From examining time series of solar irradiance data and magnetic fields at different levels, we did indeed find that the chromospheric emissions correlate best with the large magnetic field levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999SPIE.3663....8G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999SPIE.3663....8G"><span><span class="hlt">Observer</span> performance assessment of JPEG-compressed high-resolution chest <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Good, Walter F.; Maitz, Glenn S.; King, Jill L.; Gennari, Rose C.; Gur, David</p> <p>1999-05-01</p> <p>The JPEG compression algorithm was tested on a set of 529 chest radiographs that had been digitized at a spatial resolution of 100 micrometer and contrast sensitivity of 12 bits. <span class="hlt">Images</span> were compressed using five fixed 'psychovisual' quantization tables which produced average compression ratios in the range 15:1 to 61:1, and were then printed onto film. Six experienced radiologists read all cases from the laser printed film, in each of the five compressed modes as well as in the non-compressed mode. For comparison purposes, <span class="hlt">observers</span> also read the same cases with reduced pixel resolutions of 200 micrometer and 400 micrometer. The specific task involved detecting masses, pneumothoraces, interstitial disease, alveolar infiltrates and rib fractures. Over the range of compression ratios tested, for <span class="hlt">images</span> digitized at 100 micrometer, we were unable to demonstrate any statistically significant decrease (p greater than 0.05) in <span class="hlt">observer</span> performance as measured by ROC techniques. However, the <span class="hlt">observers</span>' subjective assessments of <span class="hlt">image</span> quality did decrease significantly as <span class="hlt">image</span> resolution was reduced and suggested a decreasing, but nonsignificant, trend as the compression ratio was increased. The seeming discrepancy between our failure to detect a reduction in <span class="hlt">observer</span> performance, and other published studies, is likely due to: (1) the higher resolution at which we digitized our <span class="hlt">images</span>; (2) the higher signal-to-noise ratio of our digitized films versus typical CR <span class="hlt">images</span>; and (3) our particular choice of an optimized quantization scheme.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820028848&hterms=8796&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3D8796','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820028848&hterms=8796&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3D8796"><span>Visible aurora in Jupiter's atmosphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cook, A. F., II; Jones, A. V.; Shemansky, D. E.</p> <p>1981-01-01</p> <p>The darkside limb pictures obtained by the <span class="hlt">imaging</span> experiment on Voyager 1 have been reexamined. It is concluded that the <span class="hlt">observed</span> luminosity is very likely due at least in part to Io torus aurora. If the effective wavelength of the emission lies in the 4000- to 5000-A region, the slant intensity is estimated to be about 20 kR. The <span class="hlt">observed</span> double structure may be due to a number of causes such as horizontal structure in auroral emission, aurora plus twilight or photochemical <span class="hlt">airglow</span> plus aurora.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-STS102-342-024.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-STS102-342-024.html"><span>Earth <span class="hlt">observation</span> taken during STS-102</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2001-06-14</p> <p>STS102-342-024 (8-21 March 2001)--- The largest percentage of astronaut out-the-window photography is scientific in nature. However, occasionally scenes such as this one showing the moon over Earth's <span class="hlt">airglow</span> are irresistable for crew members with cameras. The 35mm scene was recorded by one of the STS-102 astronauts from the aft flight deck of the Space Shuttle Discovery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AJ....149...50O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AJ....149...50O"><span>Difference <span class="hlt">Image</span> Analysis of Defocused <span class="hlt">Observations</span> With CSTAR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oelkers, Ryan J.; Macri, Lucas M.; Wang, Lifan; Ashley, Michael C. B.; Cui, Xiangqun; Feng, Long-Long; Gong, Xuefei; Lawrence, Jon S.; Qiang, Liu; Luong-Van, Daniel; Pennypacker, Carl R.; Yang, Huigen; Yuan, Xiangyan; York, Donald G.; Zhou, Xu; Zhu, Zhenxi</p> <p>2015-02-01</p> <p>The Chinese Small Telescope ARray carried out high-cadence time-series <span class="hlt">observations</span> of 27 square degrees centered on the South Celestial Pole during the Antarctic winter seasons of 2008-2010. Aperture photometry of the 2008 and 2010 i-band <span class="hlt">images</span> resulted in the discovery of over 200 variable stars. Yearly servicing left the array defocused for the 2009 winter season, during which the system also suffered from intermittent frosting and power failures. Despite these technical issues, nearly 800,000 useful <span class="hlt">images</span> were obtained using g, r, and clear filters. We developed a combination of difference <span class="hlt">imaging</span> and aperture photometry to compensate for the highly crowded, blended, and defocused frames. We present details of this approach, which may be useful for the analysis of time-series data from other small-aperture telescopes regardless of their <span class="hlt">image</span> quality. Using this approach, we were able to recover 68 previously known variables and detected variability in 37 additional objects. We also have determined the <span class="hlt">observing</span> statistics for Dome A during the 2009 winter season; we find the extinction due to clouds to be less than 0.1 and 0.4 mag for 40% and 63% of the dark time, respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950015355','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950015355"><span>Application of MCM <span class="hlt">image</span> construction to IRAS comet <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>Schlapfer, Martin F.; Walker, Russell G.</p> <p>1994-01-01</p> <p>There is a wealth of IRAS comet data, obtained in both the survey and pointed <span class="hlt">observations</span> modes. However, these measurements have remained largely untouched due to difficulties in removing instrumental effects from the data. We have developed a version of the Maximum Correlation Method for <span class="hlt">Image</span> Construction algorithm (MCM) which operates in the moving coordinate system of the comet and properly treats both real cometary motion and apparent motion due to spacecraft parallax. This algorithm has been implemented on a 486/33 PC in FORTRAN and IDL codes. Preprocessing of the IRAS CRDD includes baseline removal, deglitching, and removal of long tails due to dielectric time constants of the detectors. The resulting <span class="hlt">images</span> are virtually free from instrumental effects and have the highest possible spatial resolution consistent with the data sampling. We present examples of high resolution IRAS <span class="hlt">images</span> constructed from survey <span class="hlt">observations</span> of Comets P/Tempel 1 and P/Tempel 2, and pointed <span class="hlt">observations</span> of IRAS-Araki-Alcock.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e005352.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e005352.html"><span>Earth <span class="hlt">Observation</span> taken by the Expedition 29 crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-16</p> <p>ISS029-E-005352 (16 Sept. 2011) --? One of the Expedition 29 crew members aboard the International Space Station, flying at an altitude of approximately 220 miles, took this night time picture showing Africa's Congo coast, clouds, <span class="hlt">airglow</span> and Earth's terminator. Nadir coordinates are 5.32 degrees south latitude and 12.86 degrees east longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SoPh..292..106M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SoPh..292..106M"><span>First <span class="hlt">Observations</span> from the Multi-Application Solar Telescope (MAST) Narrow-Band <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mathew, Shibu K.; Bayanna, Ankala Raja; Tiwary, Alok Ranjan; Bireddy, Ramya; Venkatakrishnan, Parameswaran</p> <p>2017-08-01</p> <p>The Multi-Application Solar Telescope is a 50 cm off-axis Gregorian telescope recently installed at the Udaipur Solar Observatory, India. In order to obtain near-simultaneous <span class="hlt">observations</span> at photospheric and chromospheric heights, an <span class="hlt">imager</span> optimized for two or more wavelengths is being integrated with the telescope. Two voltage-tuneable lithium-niobate Fabry-Perot etalons along with a set of interference blocking filters have been used for developing the <span class="hlt">imager</span>. Both of the etalons are used in tandem for photospheric <span class="hlt">observations</span> in Fe i 6173 Å and chromospheric <span class="hlt">observation</span> in Hα 6563 Å spectral lines, whereas only one of the etalons is used for the chromospheric Ca II line at 8542 Å. The <span class="hlt">imager</span> is also being used for spectropolarimetric <span class="hlt">observations</span>. We discuss the characterization of the etalons at the above wavelengths, detail the integration of the <span class="hlt">imager</span> with the telescope, and present a few sets of <span class="hlt">observations</span> taken with the <span class="hlt">imager</span> set-up.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17808181','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17808181"><span>Granular convection <span class="hlt">observed</span> by magnetic resonance <span class="hlt">imaging</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ehrichs, E E; Jaeger, H M; Karczmar, G S; Knight, J B; Kuperman, V Y; Nagel, S R</p> <p>1995-03-17</p> <p>Vibrations in a granular material can spontaneously produce convection rolls reminiscent of those seen in fluids. Magnetic resonance <span class="hlt">imaging</span> provides a sensitive and noninvasive probe for the detection of these convection currents, which have otherwise been difficult to <span class="hlt">observe</span>. A magnetic resonance <span class="hlt">imaging</span> study of convection in a column of poppy seeds yielded data about the detailed shape of the convection rolls and the depth dependence of the convection velocity. The velocity was found to decrease exponentially with depth; a simple model for this behavior is presented here.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003PhDT.........9H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003PhDT.........9H"><span>High-latitude electron density <span class="hlt">observations</span> from the <span class="hlt">IMAGE</span> radio plasma <span class="hlt">imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Henize, Vance Karl</p> <p>2003-11-01</p> <p>Before the <span class="hlt">IMAGE</span> mission, electron densities in the high latitude, high altitude region of the magnetosphere were measured exclusively by in situ means. The Radio Plasma <span class="hlt">Imager</span> instrument onboard <span class="hlt">IMAGE</span> is capable of remotely <span class="hlt">observing</span> electron densities between 0.01 and 100,000 e-/cm-3 from distances of several Earth radii or more. This allows a global view of the high latitude region that has a far greater accuracy than was previously possible. Soundings of the terrestrial magnetic cusp provide the first remote <span class="hlt">observations</span> of the dynamics and poleward density profile of this feature continuously over a 60- minute interval. During steady quiet-time solar wind and interplanetary magnetic field conditions, the cusp is shown to be stable in both position and density structure with only slight variations in both. Peak electron densities within the cusp during this time are found to be somewhat higher than predicted. New procedures for deriving electron densities from radio sounding measurements are developed. The addition of curve fitting algorithms significantly increases the amount of useable data. Incorporating forward modeling techniques greatly reduces the computational time over traditional inversion methods. These methods are described in detail. A large number high latitude <span class="hlt">observations</span> of ducted right-hand extraordinary mode waves made over the course of one year of the <span class="hlt">IMAGE</span> mission are used to create a three dimensional model of the electron density profile of the terrestrial polar cap region. The dependence of electron density in the polar cap on average geocentric distance (d) is found to vary as d-6.6. This is a significantly steeper gradient than cited in earlier works such as Persoon et al., although the introduction of an asymptotic term provides for basic agreement in the limited region of their joint validity. Latitudinal and longitudinal variations are found to be insignificant. Both the mean profile power law index of the electron density profile</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..12.8128K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..12.8128K"><span>Mesospheric Temperatures and Winds measured by a VHF Meteor Radar at King Sejong Station (62.2S, 58.8W), Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Yongha; Kim, Jeong-Han; Jee, Geonwha; Lee, Chang-Sup</p> <p>2010-05-01</p> <p>A VHF radar at King Sejong Station, Antarctica has been measuring meteor echoes since March 2007. Temperatures near the mesopause are derived from meteor decay times with an improved method of selecting meteor echo samples, and compared with <span class="hlt">airglow</span> temperatures simultaneously <span class="hlt">observed</span> by a spectral <span class="hlt">airglow</span> temperature <span class="hlt">imager</span> (SATI). The temperatures derived from meteor decay times are mostly consistent with the rotational temperatures of SATI OH(6-2) and O2(0-1) emissions from March through October. During southern summer when SATI cannot be operated due to brief night time, the meteor radar <span class="hlt">observation</span> shows cold mesospheric temperatures, significantly lower than the CIRA86 model. The meteor radar <span class="hlt">observation</span> also provides wind field information between 80 and 100 km of altitude. The measured meridional winds seem to follow the summer pole to winter pole circulation, and thus are correlated with the measured seasonal temperature change. However, the correlation between meridional winds and temperatures is not found in day by day base, as a previous study reported. Tidal characteristics of both zonal and meridional winds will also be compared with those of other Antarctic stations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1991SPIE.1444..256C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1991SPIE.1444..256C"><span><span class="hlt">Observer</span> detection of <span class="hlt">image</span> degradation caused by irreversible data compression processes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Ji; Flynn, Michael J.; Gross, Barry; Spizarny, David</p> <p>1991-05-01</p> <p>Irreversible data compression methods have been proposed to reduce the data storage and communication requirements of digital <span class="hlt">imaging</span> systems. In general, the error produced by compression increases as an algorithm''s compression ratio is increased. We have studied the relationship between compression ratios and the detection of induced error using radiologic <span class="hlt">observers</span>. The nature of the errors was characterized by calculating the power spectrum of the difference <span class="hlt">image</span>. In contrast with studies designed to test whether detected errors alter diagnostic decisions, this study was designed to test whether <span class="hlt">observers</span> could detect the induced error. A paired-film <span class="hlt">observer</span> study was designed to test whether induced errors were detected. The study was conducted with chest radiographs selected and ranked for subtle evidence of interstitial disease, pulmonary nodules, or pneumothoraces. <span class="hlt">Images</span> were digitized at 86 microns (4K X 5K) and 2K X 2K regions were extracted. A full-frame discrete cosine transform method was used to compress <span class="hlt">images</span> at ratios varying between 6:1 and 60:1. The decompressed <span class="hlt">images</span> were reprinted next to the original <span class="hlt">images</span> in a randomized order with a laser film printer. The use of a film digitizer and a film printer which can reproduce all of the contrast and detail in the original radiograph makes the results of this study insensitive to instrument performance and primarily dependent on radiographic <span class="hlt">image</span> quality. The results of this study define conditions for which errors associated with irreversible compression cannot be detected by radiologic <span class="hlt">observers</span>. The results indicate that an <span class="hlt">observer</span> can detect the errors introduced by this compression algorithm for compression ratios of 10:1 (1.2 bits/pixel) or higher.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910040919&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910040919&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight"><span>A method for the retrieval of atomic oxygen density and temperature profiles from ground-based measurements of the O(+)(2D-2P) 7320 A twilight <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fennelly, J. A.; Torr, D. G.; Richards, P. G.; Torr, M. R.; Sharp, W. E.</p> <p>1991-01-01</p> <p>This paper describes a technique for extracting thermospheric profiles of the atomic-oxygen density and temperature, using ground-based measurements of the O(+)(2D-2P) doublet at 7320 and 7330 A in the twilight <span class="hlt">airglow</span>. In this method, a local photochemical model is used to calculate the 7320-A intensity; the method also utilizes an iterative inversion procedure based on the Levenberg-Marquardt method described by Press et al. (1986). The results demonstrate that, if the measurements are only limited by errors due to Poisson noise, the altitude profiles of neutral temperature and atomic oxygen concentration can be determined accurately using currently available spectrometers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRA..121.3410H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRA..121.3410H"><span>Edge of polar cap patches</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hosokawa, K.; Taguchi, S.; Ogawa, Y.</p> <p>2016-04-01</p> <p>On the night of 4 December 2013, a sequence of polar cap patches was captured by an all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span> (ASI) in Longyearbyen, Norway (78.1°N, 15.5°E). The 630.0 nm <span class="hlt">airglow</span> <span class="hlt">images</span> from the ASI of 4 second exposure time, oversampled the emission of natural lifetime (with quenching) of at least ˜30 sec, introduce no <span class="hlt">observational</span> blurring effects. By using such high-quality ASI <span class="hlt">images</span>, we succeeded in visualizing an asymmetry in the gradients between the leading/trailing edges of the patches in a 2-D fashion. The gradient in the leading edge was found to be 2-3 times steeper than that in the trailing edge. We also identified fingerlike structures, appearing only along the trailing edge of the patches, whose horizontal scale size ranged from 55 to 210 km. These fingers are considered to be manifestations of plasma structuring through the gradient-drift instability (GDI), which is known to occur only along the trailing edge of patches. That is, the current 2-D <span class="hlt">observations</span> visualized, for the first time, how GDI stirs the patch plasma and such a mixing process makes the trailing edge more gradual. This result strongly implies a close connection between the GDI-driven plasma stirring and the asymmetry in the large-scale shape of patches and then suggests that the fingerlike structures can be used as markers to estimate the fine-scale structure in the plasma flow within patches.</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/2060/20000091592','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000091592"><span>High-Definition Television (HDTV) <span class="hlt">Images</span> for Earth <span class="hlt">Observations</span> and Earth Science Applications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Robinson, Julie A.; Holland, S. Douglas; Runco, Susan K.; Pitts, David E.; Whitehead, Victor S.; Andrefouet, Serge M.</p> <p>2000-01-01</p> <p>As part of Detailed Test Objective 700-17A, astronauts acquired Earth <span class="hlt">observation</span> <span class="hlt">images</span> from orbit using a high-definition television (HDTV) camcorder, Here we provide a summary of qualitative findings following completion of tests during missions STS (Space Transport System)-93 and STS-99. We compared HDTV imagery stills to <span class="hlt">images</span> taken using payload bay video cameras, Hasselblad film camera, and electronic still camera. We also evaluated the potential for motion video <span class="hlt">observations</span> of changes in sunlight and the use of multi-aspect viewing to <span class="hlt">image</span> aerosols. Spatial resolution and color quality are far superior in HDTV <span class="hlt">images</span> compared to National Television Systems Committee (NTSC) video <span class="hlt">images</span>. Thus, HDTV provides the first viable option for video-based remote sensing <span class="hlt">observations</span> of Earth from orbit. Although under ideal conditions, HDTV <span class="hlt">images</span> have less spatial resolution than medium-format film cameras, such as the Hasselblad, under some conditions on orbit, the HDTV <span class="hlt">image</span> acquired compared favorably with the Hasselblad. Of particular note was the quality of color reproduction in the HDTV <span class="hlt">images</span> HDTV and electronic still camera (ESC) were not compared with matched fields of view, and so spatial resolution could not be compared for the two <span class="hlt">image</span> types. However, the color reproduction of the HDTV stills was truer than colors in the ESC <span class="hlt">images</span>. As HDTV becomes the operational video standard for Space Shuttle and Space Station, HDTV has great potential as a source of Earth-<span class="hlt">observation</span> data. Planning for the conversion from NTSC to HDTV video standards should include planning for Earth data archiving and distribution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19720027233&hterms=May+13th+1969&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DMay%2B13th%2B1969','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720027233&hterms=May+13th+1969&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DMay%2B13th%2B1969"><span>Simultaneous measurements of the hydrogen <span class="hlt">airglow</span> emissions of Lyman alpha, Lyman beta, and Balmer alpha.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Weller, C. S.; Meier, R. R.; Tinsley, B. A.</p> <p>1971-01-01</p> <p>Comparison of Lyman-alpha, 740- to 1050-A, and Balmer-alpha <span class="hlt">airglow</span> measurements made at 134 deg solar-zenith angle on Oct. 13, 1969, with resonance-scattering models of solar radiation. Model comparison with Lyman-alpha data fixes the hydrogen column abundance over 215 km to 2 x 10 to the 13th per cu cm within a factor of 2. Differences between the Lyman-alpha model and data indicate a polar-equatorial departure from spherical symmetry in the hydrogen distribution. A Lyman-beta model based on the hydrogen distribution found to fit the Lyman-alpha data fits the spatial variation of the 740- to 1050-A data well from 100 to 130 km, but it does not fit the data well at higher altitudes; thus the presence of more rapidly absorbed shorter-wavelength radiation is indicated. This same resonance-scattering model yields Balmer-alpha intensities that result in good spatial agreement with the Balmer-alpha measurements, but a fivefold increase in the measured solar line center Lyman-beta flux is required (as required for the Lyman-beta measurement). The intensity ratio of Lyman-beta and Balmer-alpha at night is found to be a simple measure of the hydrogen optical depth if measurements with good accuracy can be made in the visible and ultraviolet spectrum.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003DPS....35.1107T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003DPS....35.1107T"><span>Cassini <span class="hlt">imaging</span> <span class="hlt">observations</span> of Jupiter's rings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Throop, H. B.; Porco, C. C.; West, R. A.; Burns, J. A.; Showalter, M. R.; Nicholson, P. D.</p> <p>2003-05-01</p> <p>Cassini's <span class="hlt">Imaging</span> Science Subsystem (ISS) instrument took nearly 1200 <span class="hlt">images</span> of the Jupiter ring system during the spacecraft's 6-month encounter with Jupiter. These <span class="hlt">observations</span> constitute the most complete dataset of the ring taken by a single instrument, both in phase angle (0.5 - 120° at seven angles) and wavelength (0.45 - 0.93 μ {m} through eight filters). The main ring was detected in all targeted exposures; the halo and gossamer rings were too faint to be <span class="hlt">observed</span> above the planet's stray light. The optical depth and radial profile of the main ring are unchanged from that of previous studies. No evidence for broad asymmetries within the ring were found; we did identify possible evidence for 1000 km-scale clumps within the ring. Cassini <span class="hlt">observations</span> at a phase angle of 64° place an upper limit on the ring's full thickness of 80 km. We have combined the Cassini ISS and VIMS <span class="hlt">observations</span> with those from Voyager, HST, Keck, Galileo, Palomar, and IRTF. We have fit the entire suite of data using a photometric model that includes microscopic silicate dust grains as well as larger, long-lived `parent bodies' that engender this dust. Our dust grain model considers a range of spheroidal particle shapes computed using the T-matrix method (Mishchenko & Travis 1998). Our best-fit model to all the data indicates an optical depth of small particles of τ s = 4.7x 10-6 and large bodies τ l = 1.3x 10-6. The dust is concentrated about a radius of 15 μ {m}. The data are fit significantly better using non-spherical rather than spherical dust grains. The parent bodies themselves must be very red from 0.4--2.5 μ {m} and may have absorption features near 0.9 μ {m} and 2.2 μ {m}.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880066225&hterms=Abreu&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAbreu%252C%2Bc.','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880066225&hterms=Abreu&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DAbreu%252C%2Bc."><span>The auroral 6300 A emission - <span class="hlt">Observations</span> and modeling</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Solomon, Stanley C.; Hays, Paul B.; Abreu, Vincent J.</p> <p>1988-01-01</p> <p>A tomographic inversion is used to analyze measurements of the auroral atomic oxygen emission line at 6300 A made by the atmosphere explorer visible <span class="hlt">airglow</span> experiment. A comparison is made between emission altitude profiles and the results from an electron transport and chemical reaction model. Measurements of the energetic electron flux, neutral composition, ion composition, and electron density are incorporated in the model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AnGeo..35..567R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AnGeo..35..567R"><span>Seasonal MLT-region nightglow intensities, temperatures, and emission heights at a Southern Hemisphere midlatitude site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Reid, Iain M.; Spargo, Andrew J.; Woithe, Jonathan M.; Klekociuk, Andrew R.; Younger, Joel P.; Sivjee, Gulamabas G.</p> <p>2017-04-01</p> <p>We consider 5 years of spectrometer measurements of OH(6-2) and O2(0-1) <span class="hlt">airglow</span> emission intensities and temperatures made near Adelaide, Australia (35° S, 138° E), between September 2001 and August 2006 and compare them with measurements of the same parameters from at the same site using an <span class="hlt">airglow</span> <span class="hlt">imager</span>, with the intensities of the OH(8-3) and O(1S) emissions made with a filter photometer, and with 2 years of Aura MLS (Microwave Limb Sounder) v3.3 temperatures and 4.5 years of TIMED SABER (Thermosphere Ionosphere Mesosphere Energetics and Dynamics Sounding of the Atmosphere using Broadband Emission Radiometry) v2.0 temperatures for the same site. We also consider whether we can recover the actual emission heights from the intercomparison of the ground-based and satellite <span class="hlt">observations</span>. We find a significant improvement in the correlation between the spectrometer OH and SABER temperatures by interpolating the latter to constant density surfaces determined using a meteor radar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3196628','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3196628"><span>New Developments in <span class="hlt">Observer</span> Performance Methodology in Medical <span class="hlt">Imaging</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>Chakraborty, Dev P.</p> <p>2011-01-01</p> <p>A common task in medical <span class="hlt">imaging</span> is assessing whether a new <span class="hlt">imaging</span> system, or a variant of an existing one, is an improvement over an existing <span class="hlt">imaging</span> technology. <span class="hlt">Imaging</span> systems are generally quite complex, consisting of several components – e.g., <span class="hlt">image</span> acquisition hardware, <span class="hlt">image</span> processing and display hardware and software, and <span class="hlt">image</span> interpretation by radiologists– each of which can affect performance. While it may appear odd to include the radiologist as a “component” of the <span class="hlt">imaging</span> chain, since the radiologist’s decision determines subsequent patient care, the effect of the human interpretation has to be included. Physical measurements like modulation transfer function, signal to noise ratio, etc., are useful for characterizing the non-human parts of the <span class="hlt">imaging</span> chain under idealized and often unrealistic conditions, such as uniform background phantoms, target objects with sharp edges, etc. Measuring the effect on performance of the entire <span class="hlt">imaging</span> chain, including the radiologist, and using real clinical <span class="hlt">images</span>, requires different methods that fall under the rubric of <span class="hlt">observer</span> performance methods or “ROC analysis”. The purpose of this paper is to review recent developments in this field, particularly with respect to the free-response method. PMID:21978444</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006OptRv..13..215M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006OptRv..13..215M"><span><span class="hlt">Imaging</span> Analysis of Near-Field Recording Technique for <span class="hlt">Observation</span> of Biological Specimens</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moriguchi, Chihiro; Ohta, Akihiro; Egami, Chikara; Kawata, Yoshimasa; Terakawa, Susumu; Tsuchimori, Masaaki; Watanabe, Osamu</p> <p>2006-07-01</p> <p>We present an analysis of the properties of an <span class="hlt">imaging</span> based on a near-field recording technique in comparison with simulation results. In the system, the optical field distributions localized near the specimens are recorded as the surface topographic distributions of a photosensitive film. It is possible to <span class="hlt">observe</span> both soft and moving specimens, because the system does not require a scanning probe to obtain the <span class="hlt">observed</span> <span class="hlt">image</span>. The <span class="hlt">imaging</span> properties are evaluated using fine structures of paramecium, and we demonstrate that it is possible to <span class="hlt">observe</span> minute differences of refractive indices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19770028738&hterms=limnology&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dlimnology','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19770028738&hterms=limnology&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dlimnology"><span><span class="hlt">Imaging</span> radar <span class="hlt">observations</span> of frozen Arctic lakes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Elachi, C.; Bryan, M. L.; Weeks, W. F.</p> <p>1976-01-01</p> <p>A synthetic aperture <span class="hlt">imaging</span> L-band radar flown aboard the NASA CV-990 remotely sensed a number of ice-covered lakes about 48 km northwest of Bethel, Alaska. The <span class="hlt">image</span> obtained is a high resolution, two-dimensional representation of the surface backscatter cross section, and large differences in backscatter returns are <span class="hlt">observed</span>: homogeneous low returns, homogeneous high returns and/or low returns near lake borders, and high returns from central areas. It is suggested that a low return indicates that the lake is frozen completely to the bottom, while a high return indicates the presence of fresh water between the ice cover and the lake bed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997AAS...191.1102R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997AAS...191.1102R"><span>First Space VLBI <span class="hlt">Observations</span> and <span class="hlt">Images</span> Using the VLBA and VSOP</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Romney, J. D.; Benson, J. M.; Claussen, M. J.; Desai, K. M.; Flatters, C.; Mioduszewski, A. J.; Ulvestad, J. S.</p> <p>1997-12-01</p> <p>The National Radio Astronomy Observatory (NRAO) is a participant in the VSOP Space VLBI mission, an international collaboration led by Japan's Institute of Space and Astronautical Science. NRAO has committed up to 30% of scheduled <span class="hlt">observing</span> time on the Very Long Baseline Array (VLBA), and corresponding correlation resources, to Space VLBI <span class="hlt">observations</span>. The NRAO Space VLBI Project, funded by NASA, has been working for several years to complete the necessary enhancements to the VLBA correlator and the AIPS <span class="hlt">image</span> processing system. These developments were completed by the time of the successful launch of the VSOP mission's Halca spacecraft on 1997 February 12. As part of the in-orbit checkout phase, the first Space VLBI fringes from a VLBA <span class="hlt">observation</span> were detected on 1997 June 12, and the VSOP mission's first <span class="hlt">images</span>, in both the 1.6- and 5-GHz bands, were obtained shortly thereafter. In-orbit test <span class="hlt">observations</span> continued through early September, with the first General <span class="hlt">Observing</span> Time (GOT) scientific <span class="hlt">observations</span> beginning in July. Through mid-October, a total of 20 Space VLBI <span class="hlt">observations</span>, comprising 190 hours, had been completed at the VLBA correlator. This paper reviews the unique features of correlation and <span class="hlt">imaging</span> of Space VLBI <span class="hlt">observations</span>. These include, for correlation, the ephemeris for an orbiting VLBI ``station'' which is not fixed on the surface of the earth, and the requirement to close the loop on the phase-transfer process from a frequency standard on the ground to the spacecraft. <span class="hlt">Images</span> from a number of early tests and scientific <span class="hlt">observations</span> are presented. NRAO's user-support program, providing expert assistance in data analysis to Space VLBI <span class="hlt">observers</span>, is also described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMGP51A3713A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMGP51A3713A"><span>Small-scale field-aligned currents caused by tropical cyclones as <span class="hlt">observed</span> by the SWARM satellites above the ionosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aoyama, T.; Iyemori, T.; Nakanishi, K.</p> <p>2014-12-01</p> <p>We present case studies of small-scale magnetic fluctuations above typhoons, hurricanes and cyclones as <span class="hlt">observed</span> by the swarm constellation. It is reported lately that AGWs(atmospheric gravity waves) generated by meteorological phenomena in the troposphere such as typhoons and tornadoes, large earthquakes and volcanic eruptions propagate to the mesosphere and thermosphere. We <span class="hlt">observe</span> them in various forms(e.g. <span class="hlt">airglows</span>, ionospheric disturbances and TEC variations). We are proposing the following model. AGWs caused by atmospheric disturbances in the troposphere propagate to the ionospheric E-layer, drive dynamo action and generate field-aligned currents. The satellites <span class="hlt">observe</span> magnetic fluctuations above the ionosphere. In this presentation, we focus on cases of tropical cyclone(hurricanes in North America, typhoons in North-West Pacific).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016EPJWC.11922002S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016EPJWC.11922002S"><span>Development of Fluorescence <span class="hlt">Imaging</span> Lidar for Boat-Based Coral <span class="hlt">Observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sasano, Masahiko; Imasato, Motonobu; Yamano, Hiroya; Oguma, Hiroyuki</p> <p>2016-06-01</p> <p>A fluorescence <span class="hlt">imaging</span> lidar system installed in a boat-towable buoy has been developed for the <span class="hlt">observation</span> of reef-building corals. Long-range fluorescent <span class="hlt">images</span> of the sea bed can be recorded in the daytime with this system. The viability of corals is clear in these fluorescent <span class="hlt">images</span> because of the innate fluorescent proteins. In this study, the specifications and performance of the system are shown.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA34A..07K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA34A..07K"><span>Effect of equatorial electrodynamics on low-latitude thermosphere as inferred from neutral optical dayglow emission <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>Karan, D. K.; Duggirala, P. R.</p> <p>2017-12-01</p> <p>The diurnal variations in daytime <span class="hlt">airglow</span> emission intensity measurements at three wavelengths OI 777.4 nm, OI 630.0 nm, and OI 557.7 nm made from a low-latitude location, Hyderabad (Geographic 17.50 N, 78.40 E; 8.90 N Mag. Lat) in India have been investigated. The intensity patterns showed both symmetric and asymmetric behavior in their respective diurnal emission variability with respect to local noon. The asymmetric diurnal behavior is not expected considering the photochemical nature of the production mechanisms. The reason for this <span class="hlt">observed</span> asymmetric diurnal behavior has been found to be predominantly the temporal variation in the equatorial electrodynamics. The plasma that is transported across latitudes due to the action of varying electric field strength over the magnetic equator in the daytime contributes to the asymmetric diurnal behavior in the neutral daytime <span class="hlt">airglow</span> emissions. Independent magnetic and radio measurements support this finding. It is also noted that this asymmetric diurnal behavior in the neutral emission intensities has a solar cycle dependence with more number of days during high solar activity period showing asymmetric diurnal behavior compared to those during low-solar activity epoch. These intensity variations over long time scale demonstrate that the daytime neutral optical emissions are extremely sensitive to the changes in the eastward electric field over low- and equatorial-latitudes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PASP..129e5002W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PASP..129e5002W"><span>Ground-based <span class="hlt">Observation</span> System Development for the Moon Hyper-spectral <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Yang; Huang, Yu; Wang, Shurong; Li, Zhanfeng; Zhang, Zihui; Hu, Xiuqing; Zhang, Peng</p> <p>2017-05-01</p> <p>The Moon provides a suitable radiance source for on-orbit calibration of space-borne optical instruments. A ground-based <span class="hlt">observation</span> system dedicated to the hyper-spectral radiometry of the Moon has been developed for improving and validating the current lunar model. The <span class="hlt">observation</span> instrument using a dispersive <span class="hlt">imaging</span> spectrometer is particularly designed for high-accuracy <span class="hlt">observations</span> of the lunar radiance. The simulation and analysis of the push-broom mechanism is made in detail for lunar <span class="hlt">observations</span>, and the automated tracking and scanning is well accomplished in different <span class="hlt">observational</span> condition. A three-month series of hyper-spectral <span class="hlt">imaging</span> experiments of the Moon have been performed in the wavelength range from 400 to 1000 nm near Lijiang Observatory (Yunnan, China) at phase angles -83°-87°. Preliminary results and data comparison are presented, and it shows the instrument performance and lunar <span class="hlt">observation</span> capability of this system are well validated. Beyond previous measurements, this <span class="hlt">observation</span> system provides the entire lunar disk <span class="hlt">images</span> of continuous spectral coverage by adopting the push-broom mode with special scanning scheme and leads to the further research of lunar photometric model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRA..121.2569B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRA..121.2569B"><span>C/NOFS <span class="hlt">observations</span> of electromagnetic coupling between magnetically conjugate MSTID structures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Burke, W. J.; Martinis, C. R.; Lai, P. C.; Gentile, L. C.; Sullivan, C.; Pfaff, R. F.</p> <p>2016-03-01</p> <p>This report demonstrates empirically that couplings between magnetically conjugate medium-scale traveling ionospheric disturbances (MSTIDs) are electromagnetic in nature. This is accomplished by comparing plasma density, electric, and magnetic perturbations sampled simultaneously by sensors on the Communication/Navigation Outage Forecasting System (C/NOFS) satellite. During the period of interest on 17 February 2010, C/NOFS made three consecutive orbits while magnetically conjugate to the field of view of an all-sky <span class="hlt">imager</span> located at El Leoncito, Argentina (31.8°S, 69.3°W). <span class="hlt">Imaged</span> 630.0 nm <span class="hlt">airglow</span> was characterized by alternating bands of relatively bright and dark emissions that were aligned from northeast to southwest and propagated toward the northwest, characteristic of MSTIDs in the southern hemisphere. Measurable Poynting fluxes flow along the Earth's magnetic field (S||) from "generator" to "load" hemispheres. While S|| was predominantly away from the ionosphere above El Leoncito, interhemispheric energy flows were not one-way streets. Measured Poynting flux intensities diminished with time over the three C/NOFS passes, suggesting that source mechanisms of MSTIDs were absent or that initial impedance mismatches between the two hemispheres approached an equilibrium status.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20180001887&hterms=electromagnetic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Delectromagnetic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20180001887&hterms=electromagnetic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Delectromagnetic"><span>C/NOFS <span class="hlt">Observations</span> of Electromagnetic Coupling Between Magnetically Conjugate MSTID Structures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Burke, W. J.; Martinis, C. R.; Lai, P. C.; Gentile, L. C.; Sullivan, C.; Pfaff, Robert F.</p> <p>2016-01-01</p> <p>This report demonstrates empirically that couplings between magnetically conjugate medium-scale traveling ionospheric disturbances (MSTIDs) are electromagnetic in nature. This is accomplished by comparing plasma density, electric, and magnetic perturbations sampled simultaneously by sensors on the Communication Navigation Outage Forecasting System (CNOFS) satellite. During the period of interest on 17 February 2010, CNOFS made three consecutive orbits while magnetically conjugate to the field of view of an all-sky <span class="hlt">imager</span> located at El Leoncito, Argentina (31.8degS, 69.3degW). <span class="hlt">Imaged</span> 630.0 nm <span class="hlt">airglow</span> was characterized by alternating bands of relatively bright and dark emissions that were aligned from northeast to southwest and propagated toward the northwest, characteristic of MSTIDs in the southern hemisphere. Measurable Poynting fluxes flow along the Earths magnetic field (S) from generator to load hemispheres. While S was predominantly away from the ionosphere above El Leoncito, interhemispheric energy flows were not one-way streets. Measured Poynting flux intensities diminished with time over the three CNOFS passes, suggesting that source mechanisms of MSTIDs were absent or that initial impedance mismatches between the two hemispheres approached an equilibrium status.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8318E..0TG','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8318E..0TG"><span>Performance characteristics of a visual-search human-model <span class="hlt">observer</span> with sparse PET <span class="hlt">image</span> data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gifford, Howard C.</p> <p>2012-02-01</p> <p>As predictors of human performance in detection-localization tasks, statistical model <span class="hlt">observers</span> can have problems with tasks that are primarily limited by target contrast or structural noise. Model <span class="hlt">observers</span> with a visual-search (VS) framework may provide a more reliable alternative. This framework provides for an initial holistic search that identifies suspicious locations for analysis by a statistical <span class="hlt">observer</span>. A basic VS <span class="hlt">observer</span> for emission tomography focuses on hot "blobs" in an <span class="hlt">image</span> and uses a channelized nonprewhitening (CNPW) <span class="hlt">observer</span> for analysis. In [1], we investigated this model for a contrast-limited task with SPECT <span class="hlt">images</span>; herein, a statisticalnoise limited task involving PET <span class="hlt">images</span> is considered. An LROC study used 2D <span class="hlt">image</span> slices with liver, lung and soft-tissue tumors. Human and model <span class="hlt">observers</span> read the <span class="hlt">images</span> in coronal, sagittal and transverse display formats. The study thus measured the detectability of tumors in a given organ as a function of display format. The model <span class="hlt">observers</span> were applied under several task variants that tested their response to structural noise both at the organ boundaries alone and over the organs as a whole. As measured by correlation with the human data, the VS <span class="hlt">observer</span> outperformed the CNPW scanning <span class="hlt">observer</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.7126E..0MT','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.7126E..0MT"><span>High-speed <span class="hlt">imaging</span> system for <span class="hlt">observation</span> of discharge phenomena</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tanabe, R.; Kusano, H.; Ito, Y.</p> <p>2008-11-01</p> <p>A thin metal electrode tip instantly changes its shape into a sphere or a needlelike shape in a single electrical discharge of high current. These changes occur within several hundred microseconds. To <span class="hlt">observe</span> these high-speed phenomena in a single discharge, an <span class="hlt">imaging</span> system using a high-speed video camera and a high repetition rate pulse laser was constructed. A nanosecond laser, the wavelength of which was 532 nm, was used as the illuminating source of a newly developed high-speed video camera, HPV-1. The time resolution of our system was determined by the laser pulse width and was about 80 nanoseconds. The system can take one hundred pictures at 16- or 64-microsecond intervals in a single discharge event. A band-pass filter at 532 nm was placed in front of the camera to block the emission of the discharge arc at other wavelengths. Therefore, clear <span class="hlt">images</span> of the electrode were recorded even during the discharge. If the laser was not used, only <span class="hlt">images</span> of plasma during discharge and thermal radiation from the electrode after discharge were <span class="hlt">observed</span>. These results demonstrate that the combination of a high repetition rate and a short pulse laser with a high speed video camera provides a unique and powerful method for high speed <span class="hlt">imaging</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.1391G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.1391G"><span>Nightglow on Venus: Venus Express NO(UV), O2(IR), and OH(IR) <span class="hlt">Observations</span> and Implications for Upper Atmosphere Dynamics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gerard, Jean-Claude</p> <p></p> <p>Ground-based and space <span class="hlt">observations</span> have shown the presence of several emissions in the Venus nightglow. The gamma and delta bands of nitric oxide between 190 and 270 nm are ubiquitous on the Venus night side. They are excited by radiative recombination of N and O atoms created by photodissociation of CO2 and N2 molecules on the dayside of the planet. This emission has been extensively <span class="hlt">observed</span> with the SPICAV spectrograph on board Venus Express. It shows a maximum limb brightness near 115 km. Similarly, the O2 (1 ∆) emission at 1.27 µm is excited by three-body recombination of O atoms which produces an <span class="hlt">airglow</span> layer near 96 km, as was demonstrated by several studies based on <span class="hlt">observations</span> with the VIRTIS instrument on Venus Express. The two emissions are variable in space and time and show little spatial correlation. The N and O atoms are transported to the night side by the subsolar to antisolar global circulation in the thermosphere generated by the thermal contrast between the two sides of Venus. A zonal circulation is also <span class="hlt">observed</span> in the mesosphere and a region exists where both transport regimes influence the distribution of O and N atoms and the resulting <span class="hlt">airglow</span> emissions. The statistical location of the NO and O2 bright spots is not identical, which suggests that the dynamical regime is different at the altitudes of the two layers. Finally, the statistical characteristics of the OH Meinel bands in the near infrared will be presented. This emission shows similarities with O2 (1 ∆), presumably because atomic oxygen is a common precursor to both emissions. The growing information on the brightness, vertical and horizontal distribution of these emissions now provides constraints on the dynamics prevailing in the Venus upper atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018uhec.confa1042T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018uhec.confa1042T"><span>NICHE: Non-<span class="hlt">Imaging</span> Cherenkov Light <span class="hlt">Observation</span> at the TA Site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsunesada, Yoshiki; Omura, Yugo; Shin, BokKyun; Bergman, Douglas R.; Krizmanic, John F.; Nonaka, Toshiyuki</p> <p></p> <p>The Non-<span class="hlt">Imaging</span> CHErenkov Array (NICHE) is a low energy extension to the Telescope Array and TALE using an array of closely spaced (70-100 m) light collectors covering an area of up to a square km. The target is cosmic rays with energies above the knee, including the "transition region" above which Galactic cosmic rays are no more confined by the galactic magnetic field. It will be deployed in the field of view of TALE and will overlap it in energy range. TALE can <span class="hlt">observe</span> events in the energy range 3-30 PeV by non-<span class="hlt">imaging</span> air-Cherenkov, so NICHE and TALE will <span class="hlt">observe</span> <span class="hlt">imaging/non-imaging</span> Cherenkov hybrid events. NICHE itself will use both the Cherenkov lateral distribution and the Cherenkov time-width lateral distribution in measuring air showers. These two methods will allow shower energy and Xmax to be determined to infer primary types of cosmic nuclei. A prototype of the array with 15 counters, called j-NICHE, is currently being built. We describe the design of the experiment and the status of the detector development.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19304608','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19304608"><span>The effect of the <span class="hlt">observer</span> vantage point on perceived distortions in linear perspective <span class="hlt">images</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Todorović, Dejan</p> <p>2009-01-01</p> <p>Some features of linear perspective <span class="hlt">images</span> may look distorted. Such distortions appear in two drawings by Jan Vredeman de Vries involving perceived elliptical, instead of circular, pillars and tilted, instead of upright, columns. Distortions may be due to factors intrinsic to the <span class="hlt">images</span>, such as violations of the so-called Perkins's laws, or factors extrinsic to them, such as <span class="hlt">observing</span> the <span class="hlt">images</span> from positions different from their center of projection. When the correct projection centers for the two drawings were reconstructed, it was found that they were very close to the <span class="hlt">images</span> and, therefore, practically unattainable in normal <span class="hlt">observation</span>. In two experiments, enlarged versions of <span class="hlt">images</span> were used as stimuli, making the positions of the projection centers attainable for <span class="hlt">observers</span>. When <span class="hlt">observed</span> from the correct positions, the perceived distortions disappeared or were greatly diminished. Distortions perceived from other positions were smaller than would be predicted by geometrical analyses, possibly due to flatness cues in the <span class="hlt">images</span>. The results are relevant for the practical purposes of creating faithful impressions of 3-D spaces using 2-D <span class="hlt">images</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('https://ntrs.nasa.gov/search.jsp?R=20050215631&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=20050215631&hterms=Cluster+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DCluster%2Banalysis"><span>Analysis of plasmaspheric plumes: CLUSTER and <span class="hlt">IMAGE</span> <span class="hlt">observations</span> and numerical simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Darouzet, Fabien; DeKeyser, Johan; Decreau, Pierrette; Gallagher, Dennis; Pierrard, Viviane; Lemaire, Joseph; Dandouras, Iannis; Matsui, Hiroshi; Dunlop, Malcolm; Andre, Mats</p> <p>2005-01-01</p> <p>Plasmaspheric plumes have been routinely <span class="hlt">observed</span> by CLUSTER and <span class="hlt">IMAGE</span>. The CLUSTER mission provides high time resolution four-point measurements of the plasmasphere near perigee. Total electron density profiles can be derived from the plasma frequency and/or from the spacecraft potential (note that the electron spectrometer is usually not operating inside the plasmasphere); ion velocity is also measured onboard these satellites (but ion density is not reliable because of instrumental limitations). The EUV <span class="hlt">imager</span> onboard the <span class="hlt">IMAGE</span> spacecraft provides global <span class="hlt">images</span> of the plasmasphere with a spatial resolution of 0.1 RE every 10 minutes; such <span class="hlt">images</span> acquired near apogee from high above the pole show the geometry of plasmaspheric plumes, their evolution and motion. We present coordinated <span class="hlt">observations</span> for 3 plume events and compare CLUSTER in-situ data (panel A) with global <span class="hlt">images</span> of the plasmasphere obtained from <span class="hlt">IMAGE</span> (panel B), and with numerical simulations for the formation of plumes based on a model that includes the interchange instability mechanism (panel C). In particular, we study the geometry and the orientation of plasmaspheric plumes by using a four-point analysis method, the spatial gradient. We also compare several aspects of their motion as determined by different methods: (i) inner and outer plume boundary velocity calculated from time delays of this boundary <span class="hlt">observed</span> by the wave experiment WHISPER on the four spacecraft, (ii) ion velocity derived from the ion spectrometer CIS onboard CLUSTER, (iii) drift velocity measured by the electron drift instrument ED1 onboard CLUSTER and (iv) global velocity determined from successive EUV <span class="hlt">images</span>. These different techniques consistently indicate that plasmaspheric plumes rotate around the Earth, with their foot fully co-rotating, but with their tip rotating slower and moving farther out.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10136E..0PL','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10136E..0PL"><span>Foveated model <span class="hlt">observers</span> to predict human performance in 3D <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lago, Miguel A.; Abbey, Craig K.; Eckstein, Miguel P.</p> <p>2017-03-01</p> <p>We evaluate 3D search requires model <span class="hlt">observers</span> that take into account the peripheral human visual processing (foveated models) to predict human <span class="hlt">observer</span> performance. We show that two different 3D tasks, free search and location-known detection, influence the relative human visual detectability of two signals of different sizes in synthetic backgrounds mimicking the noise found in 3D digital breast tomosynthesis. One of the signals resembled a microcalcification (a small and bright sphere), while the other one was designed to look like a mass (a larger Gaussian blob). We evaluated current standard models <span class="hlt">observers</span> (Hotelling; Channelized Hotelling; non-prewhitening matched filter with eye filter, NPWE; and non-prewhitening matched filter model, NPW) and showed that they incorrectly predict the relative detectability of the two signals in 3D search. We propose a new model <span class="hlt">observer</span> (3D Foveated Channelized Hotelling <span class="hlt">Observer</span>) that incorporates the properties of the visual system over a large visual field (fovea and periphery). We show that the foveated model <span class="hlt">observer</span> can accurately predict the rank order of detectability of the signals in 3D <span class="hlt">images</span> for each task. Together, these results motivate the use of a new generation of foveated model <span class="hlt">observers</span> for predicting <span class="hlt">image</span> quality for search tasks in 3D <span class="hlt">imaging</span> modalities such as digital breast tomosynthesis or computed tomography.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820038646&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820038646&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight"><span>Exospheric temperatures deduced from 7320- to 7330-A /O/+//2P/ - O/+//2D// twilight <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>Yee, J. H.; Abreu, V. J.</p> <p>1982-01-01</p> <p>A technique developed to deduce exospheric temperatures from the 7320- to 7330-A emission measured by the visible <span class="hlt">airglow</span> 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 <span class="hlt">observed</span> 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.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008SPIE.7055E..07L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008SPIE.7055E..07L"><span>Dualband infrared <span class="hlt">imaging</span> spectrometer: <span class="hlt">observations</span> 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>LeVan, Paul D.; Beecken, Brian P.; Lindh, Cory</p> <p>2008-08-01</p> <p>We reported previously on full-disk <span class="hlt">observations</span> of the sun through a layer of black polymer, used to protect the entrance aperture of a novel dualband spectrometer while transmitting discrete wavelength regions in the MWIR & LWIR1. More recently, the spectrometer was used to assess the accuracy of recovery of unknown blackbody temperatures2. Here, we briefly describe MWIR <span class="hlt">observations</span> of the full Moon made in Jan 2008. As was the case for the solar <span class="hlt">observations</span>, the Moon was allowed to drift across the spectrometer slit by Earth's rotation. A detailed sensor calibration performed prior to the <span class="hlt">observations</span> accounts for sensor non-uniformities; the spectral <span class="hlt">images</span> of the Moon therefore include atmospheric transmission features. Our plans are to repeat the <span class="hlt">observations</span> at liquid helium temperatures, thereby allowing both MWIR & LWIR spectral coverage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUOSME53A..01P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUOSME53A..01P"><span>Applications of Geostationary Ocean Color <span class="hlt">Imager</span> (GOCI) <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>Park, Y. J.</p> <p>2016-02-01</p> <p>Ocean color remote-sensing technique opened a new era for biological oceanography by providing the global distribution of phytoplankton biomass every a few days. It has been proved useful for a variety of applications in coastal waters as well as oceanic waters. However, most ocean color sensors deliver less than one <span class="hlt">image</span> per day for low and middle latitude areas, and this once a day <span class="hlt">image</span> is insufficient to resolve transient or high frequency processes. Korean Geostationary Ocean Color <span class="hlt">Imager</span> (GOCI), the first ever ocean color instrument operated on geostationary orbit, is collecting ocean color radiometry (OCR) data (multi-band radiances at the visible to NIR spectral wavelengths) since July, 2010. GOCI has an unprecedented capability to provide eight OCR <span class="hlt">images</span> a day with a 500m resolution for the North East Asian seas Monitoring the spatial and temporal variability is important to understand many processes occurring in open ocean and coastal environments. With a series of <span class="hlt">images</span> consecutively acquired by GOCI, we are now able to look into (sub-)diurnal variabilities of coastal ocean color products such as phytoplankton biomass, suspended particles concentrations, and primary production. The eight <span class="hlt">images</span> taken a day provide another way to derive maps of ocean current velocity. Compared to polar orbiters, GOCI delivers more frequent <span class="hlt">images</span> with constant viewing angle, which enables to better monitor and thus respond to coastal water issues such as harmful algal blooms, floating green and brown algae. The frequent <span class="hlt">observation</span> capability for local area allows us to respond timely to natural disasters and hazards. GOCI <span class="hlt">images</span> are often useful to identify sea fog, sea ice, wild fires, volcanic eruptions, transport of dust aerosols, snow covered area, etc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e034092.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e034092.html"><span>Progress 42 re-entry</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-10-29</p> <p>ISS029-E-034092 (29 Oct. 2011) --- This unusual photograph, captured by one of the Expedition 29 crew members aboard the International Space Station, highlights the reentry plasma trail (center) of Progress 42P (M-10M) supply vehicle. Progress 42P docked at the space station on April 29, 2011, and was undocked and de-orbited approximately 183 days later on Oct. 29, 2011. The ISS was located over the southern Pacific Ocean when this <span class="hlt">image</span> was taken. Light from the rising sun illuminates the curvature of the Earth limb (horizon line) at top, but does not completely overwhelm the <span class="hlt">airglow</span> visible at <span class="hlt">image</span> top left. <span class="hlt">Airglow</span> is caused by light emitted at specific wavelengths by atoms and molecules excited by ultraviolet radiation in the upper atmosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JASTP.140....1J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JASTP.140....1J"><span>A case study of A mesoscale gravity wave in the MLT region using simultaneous multi-instruments in Beijing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jia, Mingjiao; Xue, Xianghui; Dou, Xiankang; Tang, Yihuan; Yu, Chao; Wu, Jianfei; Xu, Jiyao; Yang, Guotao; Ning, Baiqi; Hoffmann, Lars</p> <p>2016-03-01</p> <p>In this work, we used <span class="hlt">observational</span> data from an all-sky <span class="hlt">airglow</span> <span class="hlt">imager</span> at Xinglong (40.2 °N, 117.4 °E), a sodium lidar at Yanqing (40.4 °N, 116.0 °E) and a meteor radar at Shisanling (40.3 °N, 116.2 °E) to study the propagation of a mesoscale gravity wave. During the night of March 1, 2011, the <span class="hlt">imager</span> identified a mesoscale gravity wave structure in the OH <span class="hlt">airglow</span> that had a wave period of 2 hours, propagated along an azimuthal direction (clockwise) with an angle of 163°, a phase speed of 73 m/s, and a horizontal wavelength of 566 km. Simultaneous measurements provided by the sodium lidar also showed a perturbation in the sodium layer with a 2-hour period. Based on the SABER/TIMED and radar data, we estimated that the momentum flux and the energy flux of the gravity wave were approximately 0.59 m2/s2 and 0.22 mW/m2, respectively. Ray-tracing analysis showed that the gravity wave was likely generated in the center of Lake Baikal owing to the existence of a jet- front system in the upper troposphere at that time.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4963469','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4963469"><span>Design of a practical model-<span class="hlt">observer</span>-based <span class="hlt">image</span> quality assessment method for x-ray computed tomography <span class="hlt">imaging</span> systems</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Tseng, Hsin-Wu; Fan, Jiahua; Kupinski, Matthew A.</p> <p>2016-01-01</p> <p>Abstract. The use of a channelization mechanism on model <span class="hlt">observers</span> not only makes mimicking human visual behavior possible, but also reduces the amount of <span class="hlt">image</span> data needed to estimate the model <span class="hlt">observer</span> parameters. The channelized Hotelling <span class="hlt">observer</span> (CHO) and channelized scanning linear <span class="hlt">observer</span> (CSLO) have recently been used to assess CT <span class="hlt">image</span> quality for detection tasks and combined detection/estimation tasks, respectively. Although the use of channels substantially reduces the amount of data required to compute <span class="hlt">image</span> quality, the number of scans required for CT <span class="hlt">imaging</span> is still not practical for routine use. It is our desire to further reduce the number of scans required to make CHO or CSLO an <span class="hlt">image</span> quality tool for routine and frequent system validations and evaluations. This work explores different data-reduction schemes and designs an approach that requires only a few CT scans. Three different kinds of approaches are included in this study: a conventional CHO/CSLO technique with a large sample size, a conventional CHO/CSLO technique with fewer samples, and an approach that we will show requires fewer samples to mimic conventional performance with a large sample size. The mean value and standard deviation of areas under ROC/EROC curve were estimated using the well-validated shuffle approach. The results indicate that an 80% data reduction can be achieved without loss of accuracy. This substantial data reduction is a step toward a practical tool for routine-task-based QA/QC CT system assessment. PMID:27493982</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss030e010008.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss030e010008.html"><span>Earth <span class="hlt">Observations</span> taken by Expedition 30 crewmember</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-12-04</p> <p>ISS030-E-010008 (4 Dec. 2011) --- One of the Expedition 30 crew members aboard the Earth-orbiting International Space Station photographed this night time scene of the Iberian Peninsula on Dec. 4, 2011. The city lights of Spain and Portugal define the peninsula. Several large metropolitan areas are visible, marked by their relatively large and brightly lit areas, such as two capital cities -- Madrid, Spain, located near the center of the peninsula?s interior, and Lisbon, Portugal, located along the southwestern coastline. Ancient Seville, visible at <span class="hlt">image</span> right to the north of the approximately 14 kilometer-wide Strait of Gibraltar, is one of the largest cities in Spain. All together, the Principality of Andorra, the Kingdom of Spain and the Portuguese Republic total approximately 590,000 square kilometers of landmass. The peninsula is bounded by the Atlantic Ocean to the northwest, west, and southwest and the Mediterranean Sea to the east. Its northeastern boundary with the rest of continental Europe is marked by the Pyrenees mountain range. The view is looking outwards from the orbital outpost toward the east. The network of smaller cities and towns in the interior and along the coastline attest to the large extent of human presence on the Iberian landscape. Blurring of the city lights is caused by thin cloud cover (<span class="hlt">image</span> left and center), while the cloud tops are dimly illuminated by moonlight. Though obscured, the lights of France are visible near the horizon line at <span class="hlt">image</span> upper left, while the lights of northern Africa are more clearly discernable at <span class="hlt">image</span> right. The gold to green line of <span class="hlt">airglow</span>, caused by excitation of upper atmosphere gas molecules by ultraviolet radiation, parallels the horizon line (or Earth limb).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016MNRAS.458.3248S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016MNRAS.458.3248S"><span>The advantages of using a Lucky <span class="hlt">Imaging</span> camera for <span class="hlt">observations</span> of microlensing events</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sajadian, Sedighe; Rahvar, Sohrab; Dominik, Martin; Hundertmark, Markus</p> <p>2016-05-01</p> <p>In this work, we study the advantages of using a Lucky <span class="hlt">Imaging</span> camera for the <span class="hlt">observations</span> of potential planetary microlensing events. Our aim is to reduce the blending effect and enhance exoplanet signals in binary lensing systems composed of an exoplanet and the corresponding parent star. We simulate planetary microlensing light curves based on present microlensing surveys and follow-up telescopes where one of them is equipped with a Lucky <span class="hlt">Imaging</span> camera. This camera is used at the Danish 1.54-m follow-up telescope. Using a specific <span class="hlt">observational</span> strategy, for an Earth-mass planet in the resonance regime, where the detection probability in crowded fields is smaller, Lucky <span class="hlt">Imaging</span> <span class="hlt">observations</span> improve the detection efficiency which reaches 2 per cent. Given the difficulty of detecting the signal of an Earth-mass planet in crowded-field <span class="hlt">imaging</span> even in the resonance regime with conventional cameras, we show that Lucky <span class="hlt">Imaging</span> can substantially improve the detection efficiency.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930044726&hterms=Low+vision&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DLow%2Bvision','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930044726&hterms=Low+vision&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DLow%2Bvision"><span><span class="hlt">Image</span> enhancement filters significantly improve reading performance for low vision <span class="hlt">observers</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lawton, T. B.</p> <p>1992-01-01</p> <p>As people age, so do their photoreceptors; many photoreceptors in central vision stop functioning when a person reaches their late sixties or early seventies. Low vision <span class="hlt">observers</span> with losses in central vision, those with age-related maculopathies, were studied. Low vision <span class="hlt">observers</span> no longer see high spatial frequencies, being unable to resolve fine edge detail. We developed <span class="hlt">image</span> enhancement filters to compensate for the low vision <span class="hlt">observer</span>'s losses in contrast sensitivity to intermediate and high spatial frequencies. The filters work by boosting the amplitude of the less visible intermediate spatial frequencies. The lower spatial frequencies. These <span class="hlt">image</span> enhancement filters not only reduce the magnification needed for reading by up to 70 percent, but they also increase the <span class="hlt">observer</span>'s reading speed by 2-4 times. A summary of this research is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ESS.....2.2101D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ESS.....2.2101D"><span>Exploring Hitherto Uncharted Planet Territory with Lucky-<span class="hlt">imaging</span> Microlensing <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>Dominik, Martin; Jørgensen, U. G.; Hessman, F. V.; Horne, K.; Harpsøe, K.; Skottfelt, J.; MiNDSTEp Consortium</p> <p>2011-09-01</p> <p>Leading the agenda for pushing the planet sensitivity limit towards the mass of the Moon, we will report first results from our 2011 MiNDSTEp (Microlensing Network for the Detection of Small Terrestrial Exoplanets) lucky-<span class="hlt">imaging</span> microlensing follow-up campaign with the Danish 1.54m at ESO La Silla. It serves as a precursor to <span class="hlt">observations</span> with a global network comprising the LCOGT/SUPAscope, SONG, and MONET 1m-class robotic telescope networks gradually deployed from 2011 to 2014. As for <span class="hlt">observations</span> from space, the lucky-<span class="hlt">imaging</span> technique allows us to get around the atmospheric <span class="hlt">image</span> blurring and to obtain a resolution near the diffraction limit. This enables high-precision photometry on considerably fainter (smaller) stars in the crowded fields towards the Galactic bulge than obtainable from ground-based surveys. Monitoring smaller source stars in turn provides sensitivity to planets with smaller masses orbiting the lens star. M.D. is supported by a Royal Society University Research Fellowship</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998IAUS..185..453N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998IAUS..185..453N"><span>MHD oscillations <span class="hlt">observed</span> in the solar photosphere with the Michelson Doppler <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Norton, A.; Ulrich, R. K.; Bogart, R. S.; Bush, R. I.; Hoeksema, J. T.</p> <p></p> <p>Magnetohydrodynamic oscillations are <span class="hlt">observed</span> in the solar photosphere with the Michelson Doppler <span class="hlt">Imager</span> (MDI). <span class="hlt">Images</span> of solar surface velocity and magnetic field strength with 4'' spatial resolution and a 60 second temporal resolution are analyzed. A two dimensional gaussian aperture with a FWHM of 10'' is applied to the data in regions of sunspot, plage and quiet sun and the resulting averaged signal is returned each minute. Significant power is <span class="hlt">observed</span> in the magnetic field oscillations with periods of five minutes. The effect of misregistration between MDI's left circularly polarized (LCP) and right circularly polarized (RCP) <span class="hlt">images</span> has been investigated and is found not to be the cause of the <span class="hlt">observed</span> magnetic oscillations. It is assumed that the large amplitude acoustic waves with 5 minute periods are the driving mechanism behind the magnetic oscillations. The nature of the magnetohydrodynamic oscillations are characterized by their phase relations with simultaneously <span class="hlt">observed</span> solar surface velocity oscillations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900027756&hterms=nonimaging&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnonimaging','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900027756&hterms=nonimaging&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnonimaging"><span>Simultaneous multi-frequency <span class="hlt">imaging</span> <span class="hlt">observations</span> of solar microwave bursts</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.; White, S. M.; Schmahl, E. J.</p> <p>1989-01-01</p> <p>The results of simultaneous two-frequency <span class="hlt">imaging</span> <span class="hlt">observations</span> of solar microwave bursts with the Very Large Array are reviewed. Simultaneous 2 and 6 cm <span class="hlt">observations</span> have been made of bursts which are optically thin at both frequencies, or optically thick at the lower frequency. In the latter case, the source structure may differ at the two frequencies, but the two sources usually seem to be related. However, this is not always true of simultaneous 6 and 20 cm <span class="hlt">observations</span>. The results have implications for the analysis of nonimaging radio data of solar and stellar flares.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006A%26A...452..119O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006A%26A...452..119O"><span>ESO <span class="hlt">imaging</span> survey: infrared <span class="hlt">observations</span> of CDF-S and HDF-S</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olsen, L. F.; Miralles, J.-M.; da Costa, L.; Benoist, C.; Vandame, B.; Rengelink, R.; Rité, C.; Scodeggio, M.; Slijkhuis, R.; Wicenec, A.; Zaggia, S.</p> <p>2006-06-01</p> <p>This paper presents infrared data obtained from <span class="hlt">observations</span> carried out at the ESO 3.5 m New Technology Telescope (NTT) of the Hubble Deep Field South (HDF-S) and the Chandra Deep Field South (CDF-S). These data were taken as part of the ESO <span class="hlt">Imaging</span> Survey (EIS) program, a public survey conducted by ESO to promote follow-up <span class="hlt">observations</span> with the VLT. In the HDF-S field the infrared <span class="hlt">observations</span> cover an area of ~53 square arcmin, encompassing the HST WFPC2 and STIS fields, in the JHKs passbands. The seeing measured in the final stacked <span class="hlt">images</span> ranges from 0.79 arcsec to 1.22 arcsec and the median limiting magnitudes (AB system, 2'' aperture, 5σ detection limit) are J_AB˜23.0, H_AB˜22.8 and K_AB˜23.0 mag. Less complete data are also available in JKs for the adjacent HST NICMOS field. For CDF-S, the infrared <span class="hlt">observations</span> cover a total area of ~100 square arcmin, reaching median limiting magnitudes (as defined above) of J_AB˜23.6 and K_AB˜22.7 mag. For one CDF-S field H band data are also available. This paper describes the <span class="hlt">observations</span> and presents the results of new reductions carried out entirely through the un-supervised, high-throughput EIS Data Reduction System and its associated EIS/MVM C++-based <span class="hlt">image</span> processing library developed, over the past 5 years, by the EIS project and now publicly available. The paper also presents source catalogs extracted from the final co-added <span class="hlt">images</span> which are used to evaluate the scientific quality of the survey products, and hence the performance of the software. This is done comparing the results obtained in the present work with those obtained by other authors from independent data and/or reductions carried out with different software packages and techniques. The final science-grade catalogs together with the astrometrically and photometrically calibrated co-added <span class="hlt">images</span> are available at CDS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28120394','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28120394"><span>Comparing the Scoring of Human Decomposition from Digital <span class="hlt">Images</span> to Scoring Using On-site <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>Dabbs, Gretchen R; Bytheway, Joan A; Connor, Melissa</p> <p>2017-09-01</p> <p>When in forensic casework or empirical research in-person assessment of human decomposition is not possible, the sensible substitution is color photographic <span class="hlt">images</span>. To date, no research has confirmed the utility of color photographic <span class="hlt">images</span> as a proxy for in situ <span class="hlt">observation</span> of the level of decomposition. Sixteen <span class="hlt">observers</span> scored photographs of 13 human cadavers in varying decomposition stages (PMI 2-186 days) using the Total Body Score system (total n = 929 <span class="hlt">observations</span>). The on-site TBS was compared with recorded <span class="hlt">observations</span> from digital color <span class="hlt">images</span> using a paired samples t-test. The average difference between on-site and photographic <span class="hlt">observations</span> was -0.20 (t = -1.679, df = 928, p = 0.094). Individually, only two <span class="hlt">observers</span>, both students with <1 year of experience, demonstrated TBS statistically significantly different than the on-site value, suggesting that with experience, <span class="hlt">observations</span> of human decomposition based on digital <span class="hlt">images</span> can be substituted for assessments based on <span class="hlt">observation</span> of the corpse in situ, when necessary. © 2017 American Academy of Forensic Sciences.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AAS...198.2103C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AAS...198.2103C"><span>Polarization <span class="hlt">Observations</span> with the Cosmic Background <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cartwright, J. K.; Padin, S.; Pearson, T. J.; Readhead, A. C. S.; Shepherd, M. C.; Taylor, G. B.</p> <p>2001-05-01</p> <p>We describe polarization <span class="hlt">observations</span> of the CMBR with the Cosmic Background <span class="hlt">Imager</span>, a 13 element interferometer which operates in the 26-36 GHz band from a site at 5000m in northern Chile. The array consists of 90-cm Cassegrain antennas mounted on a single, fully steerable platform; this platform can be rotated about the optical axis to facilitate polarization <span class="hlt">observations</span>. The CBI employs single mode circularly polarized receivers, of which 12 are configured for LCP and one is configured for RCP. The 12 cross polarized baselines sample multipoles from l 600 to l 3500. The instrumental polarization of the CBI was calibrated with <span class="hlt">observations</span> of 3C279, a bright polarized source which is unresolved by the CBI. Because the centimeter flux of 3C279 is variable, it was monitored twice per month for 8 months in 2000 with the VLA at 22 and 43 GHz. These <span class="hlt">observations</span> also established the stability of the polarization characteristics of the CBI. This work was made possible by NSF grant AST-9802989</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApJ...852...79Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApJ...852...79Y"><span>Two Solar Tornadoes <span class="hlt">Observed</span> with the Interface Region <span class="hlt">Imaging</span> Spectrograph</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Zihao; Tian, Hui; Peter, Hardi; Su, Yang; Samanta, Tanmoy; Zhang, Jingwen; Chen, Yajie</p> <p>2018-01-01</p> <p>The barbs or legs of some prominences show an apparent motion of rotation, which are often termed solar tornadoes. It is under debate whether the apparent motion is a real rotating motion, or caused by oscillations or counter-streaming flows. We present analysis results from spectroscopic <span class="hlt">observations</span> of two tornadoes by the Interface Region <span class="hlt">Imaging</span> Spectrograph. Each tornado was <span class="hlt">observed</span> for more than 2.5 hr. Doppler velocities are derived through a single Gaussian fit to the Mg II k 2796 Å and Si IV 1393 Å line profiles. We find coherent and stable redshifts and blueshifts adjacent to each other across the tornado axes, which appears to favor the interpretation of these tornadoes as rotating cool plasmas with temperatures of 104 K–105 K. This interpretation is further supported by simultaneous <span class="hlt">observations</span> of the Atmospheric <span class="hlt">Imaging</span> Assembly on board the Solar Dynamics Observatory, which reveal periodic motions of dark structures in the tornadoes. Our results demonstrate that spectroscopic <span class="hlt">observations</span> can provide key information to disentangle different physical processes in solar prominences.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9037E..0LK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9037E..0LK"><span><span class="hlt">Observer</span> assessment of multi-pinhole SPECT geometries for prostate cancer <span class="hlt">imaging</span>: a simulation study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kalantari, Faraz; Sen, Anando; Gifford, Howard C.</p> <p>2014-03-01</p> <p>SPECT <span class="hlt">imaging</span> using In-111 ProstaScint is an FDA-approved method for diagnosing prostate cancer metastases within the pelvis. However, conventional medium-energy parallel-hole (MEPAR) collimators produce poor <span class="hlt">image</span> quality and we are investigating the use of multipinhole (MPH) <span class="hlt">imaging</span> as an alternative. This paper presents a method for evaluating MPH designs that makes use of sampling-sensitive (SS) mathematical model <span class="hlt">observers</span> for tumor detectionlocalization tasks. Key to our approach is the redefinition of a normal (or background) reference <span class="hlt">image</span> that is used with scanning model <span class="hlt">observers</span>. We used this approach to compare different MPH configurations for the task of small-tumor detection in the prostate and surrounding lymph nodes. Four configurations used 10, 20, 30, and 60 pinholes evenly spaced over a complete circular orbit. A fixed-count acquisition protocol was assumed. Spherical tumors were placed within a digital anthropomorphic phantom having a realistic Prostascint biodistribution. <span class="hlt">Imaging</span> data sets were generated with an analytical projector and reconstructed volumes were obtained with the OSEM algorithm. The MPH configurations were compared in a localization ROC (LROC) study with 2D pelvic <span class="hlt">images</span> and both human and model <span class="hlt">observers</span>. Regular and SS versions of the scanning channelized nonprewhitening (CNPW) and visual-search (VS) model <span class="hlt">observers</span> were applied. The SS models demonstrated the highest correlations with the average human-<span class="hlt">observer</span> results</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9787E..0JK','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9787E..0JK"><span>The potential of pigeons as surrogate <span class="hlt">observers</span> in medical <span class="hlt">image</span> perception studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Krupinski, Elizabeth A.; Levenson, Richard M.; Navarro, Victor; Wasserman, Edward A.</p> <p>2016-03-01</p> <p>Assessment of medical <span class="hlt">image</span> quality and how changes in <span class="hlt">image</span> appearance impact performance are critical but assessment can be expensive and time-consuming. Could an animal (pigeon) <span class="hlt">observer</span> with well-known visual skills and documented ability to distinguish complex visual stimuli serve as a surrogate for the human <span class="hlt">observer</span>? Using sets of whole slide pathology (WSI) and mammographic <span class="hlt">images</span> we trained pigeons (cohorts of 4) to detect and/or classify lesions in medical <span class="hlt">images</span>. Standard training methods were used. A chamber equipped with a 15' display with a resistive touchscreen was used to display the <span class="hlt">images</span> and record responses (pecks). Pigeon pellets were dispensed for correct responses. The pigeons readily learned to distinguish benign from malignant breast cancer histopathology in WSI (mean % correct responses rose 50% to 85% over 15 days) and generalized readily from 4X to 10X and 20X magnifications; to detect microcalcifications (mean % correct responses rose 50% to over 85% over 25 days); to distinguish benign from malignant breast masses (3 of 4 birds learned this task to around 80% and 60% over 10 days); and ignore compression artifacts in WSI (performance with uncompressed slides averaged 95% correct; 15:1 and 27:1 compression slides averaged 92% and 90% correct). Pigeons models may help us better understand medical <span class="hlt">image</span> perception and may be useful in quality assessment by serving as surrogate <span class="hlt">observers</span> for certain types of studies.</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/2017EGUGA..19.9014T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.9014T"><span><span class="hlt">Observing</span> hydrological processes: recent advancements in surface flow monitoring through <span class="hlt">image</span> analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tauro, Flavia; Grimaldi, Salvatore</p> <p>2017-04-01</p> <p>Recently, several efforts have been devoted to the design and development of innovative, and often unintended, approaches for the acquisition of hydrological data. Among such pioneering techniques, this presentation reports recent advancements towards the establishment of a novel noninvasive and potentially continuous methodology based on the acquisition and analysis of <span class="hlt">images</span> for spatially distributed <span class="hlt">observations</span> of the kinematics of surface waters. The approach aims at enabling rapid, affordable, and accurate surface flow monitoring of natural streams. Flow monitoring is an integral part of hydrological sciences and is essential for disaster risk reduction and the comprehension of natural phenomena. However, water processes are inherently complex to <span class="hlt">observe</span>: they are characterized by multiscale and highly heterogeneous phenomena which have traditionally demanded sophisticated and costly measurement techniques. Challenges in the implementation of such techniques have also resulted in lack of hydrological data during extreme events, in difficult-to-access environments, and at high temporal resolution. By combining low-cost yet high-resolution <span class="hlt">images</span> and several velocimetry algorithms, noninvasive flow monitoring has been successfully conducted at highly heterogeneous scales, spanning from rills to highly turbulent streams, and medium-scale rivers, with minimal supervision by external users. Noninvasive <span class="hlt">image</span> data acquisition has also afforded <span class="hlt">observations</span> in high flow conditions. Latest novelties towards continuous flow monitoring at the catchment scale have entailed the development of a remote gauge-cam station on the Tiber River and integration of flow monitoring through <span class="hlt">image</span> analysis with unmanned aerial systems (UASs) technology. The gauge-cam station and the UAS platform both afford noninvasive <span class="hlt">image</span> acquisition and calibration through an innovative laser-based setup. Compared to traditional point-based instrumentation, <span class="hlt">images</span> allow for generating surface</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA33A2419V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA33A2419V"><span>Investigating middle-atmospheric gravity waves associated with a sprite-producing mesoscale convective event</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vollmer, D. R.; McHarg, M. G.; Harley, J.; Haaland, R. K.; Stenbaek-Nielsen, H.</p> <p>2016-12-01</p> <p>On 23 July 2014, a mesoscale convective event over western Nebraska produced a large number of sprites. One frame per second <span class="hlt">images</span> obtained from a low-noise Andor Scientific CMOS camera showed regularly-spaced horizontal striations in the <span class="hlt">airglow</span> both before and during several of the sprite events, suggesting the presence of vertically-propagating gravity waves in the middle atmosphere. Previous work hypothesized that the gravity waves were produced by the thunderstorm itself. We compare our <span class="hlt">observations</span> with previous work, and present numerical simulations conducted to determine source, structure, and propagation of atmospheric gravity waves.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015Ap%26SS.358...25G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015Ap%26SS.358...25G"><span>Improved SOT (Hinode mission) high resolution solar <span class="hlt">imaging</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>Goodarzi, H.; Koutchmy, S.; Adjabshirizadeh, A.</p> <p>2015-08-01</p> <p>We consider the best today available <span class="hlt">observations</span> of the Sun free of turbulent Earth atmospheric effects, taken with the Solar Optical Telescope (SOT) onboard the Hinode spacecraft. Both the instrumental smearing and the <span class="hlt">observed</span> stray light are analyzed in order to improve the resolution. The Point Spread Function (PSF) corresponding to the blue continuum Broadband Filter <span class="hlt">Imager</span> (BFI) near 450 nm is deduced by analyzing (i) the limb of the Sun and (ii) <span class="hlt">images</span> taken during the transit of the planet Venus in 2012. A combination of Gaussian and Lorentzian functions is selected to construct a PSF in order to remove both smearing due to the instrumental diffraction effects (PSF core) and the large-angle stray light due to the spiders and central obscuration (wings of the PSF) that are responsible for the parasitic stray light. A Max-likelihood deconvolution procedure based on an optimum number of iterations is discussed. It is applied to several solar field <span class="hlt">images</span>, including the granulation near the limb. The normal non-magnetic granulation is compared to the abnormal granulation which we call magnetic. A new feature appearing for the first time at the extreme- limb of the disk (the last 100 km) is discussed in the context of the definition of the solar edge and of the solar diameter. A single sunspot is considered in order to illustrate how effectively the restoration works on the sunspot core. A set of 125 consecutive deconvolved <span class="hlt">images</span> is assembled in a 45 min long movie illustrating the complexity of the dynamical behavior inside and around the sunspot.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22654012-tu-fg-validation-channelized-hotelling-observer-optimize-chest-radiography-image-processing-nodule-detection-human-observer-study','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22654012-tu-fg-validation-channelized-hotelling-observer-optimize-chest-radiography-image-processing-nodule-detection-human-observer-study"><span>TU-FG-209-11: Validation of a Channelized Hotelling <span class="hlt">Observer</span> to Optimize Chest Radiography <span class="hlt">Image</span> Processing for Nodule Detection: A Human <span class="hlt">Observer</span> Study</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>Sanchez, A; Little, K; Chung, J</p> <p></p> <p>Purpose: To validate the use of a Channelized Hotelling <span class="hlt">Observer</span> (CHO) model for guiding <span class="hlt">image</span> processing parameter selection and enable improved nodule detection in digital chest radiography. Methods: In a previous study, an anthropomorphic chest phantom was <span class="hlt">imaged</span> with and without PMMA simulated nodules using a GE Discovery XR656 digital radiography system. The impact of <span class="hlt">image</span> processing parameters was then explored using a CHO with 10 Laguerre-Gauss channels. In this work, we validate the CHO’s trend in nodule detectability as a function of two processing parameters by conducting a signal-known-exactly, multi-reader-multi-case (MRMC) ROC <span class="hlt">observer</span> study. Five naive readers scored confidencemore » of nodule visualization in 384 <span class="hlt">images</span> with 50% nodule prevalence. The <span class="hlt">image</span> backgrounds were regions-of-interest extracted from 6 normal patient scans, and the digitally inserted simulated nodules were obtained from phantom data in previous work. Each patient <span class="hlt">image</span> was processed with both a near-optimal and a worst-case parameter combination, as determined by the CHO for nodule detection. The same 192 ROIs were used for each <span class="hlt">image</span> processing method, with 32 randomly selected lung ROIs per patient <span class="hlt">image</span>. Finally, the MRMC data was analyzed using the freely available iMRMC software of Gallas et al. Results: The <span class="hlt">image</span> processing parameters which were optimized for the CHO led to a statistically significant improvement (p=0.049) in human <span class="hlt">observer</span> AUC from 0.78 to 0.86, relative to the <span class="hlt">image</span> processing implementation which produced the lowest CHO performance. Conclusion: Differences in user-selectable <span class="hlt">image</span> processing methods on a commercially available digital radiography system were shown to have a marked impact on performance of human <span class="hlt">observers</span> in the task of lung nodule detection. Further, the effect of processing on humans was similar to the effect on CHO performance. Future work will expand this study to include a wider range of detection/classification tasks and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910011702&hterms=tokyo+round+japan&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dtokyo%2Bround%2Bjapan','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910011702&hterms=tokyo+round+japan&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dtokyo%2Bround%2Bjapan"><span>Optical and near-IR <span class="hlt">imaging</span> <span class="hlt">observations</span> of comet Austin 1989c1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Watanabe, J.; Hiromoto, N.; Takami, H.; Aoki, TE.; Nakamura, T.; Takagishi, K.; Hatsukade, I.; Isobe, S.; Sasaki, G.; Sugai, H.</p> <p>1990-01-01</p> <p>Near-nucleus <span class="hlt">imaging</span> <span class="hlt">observations</span> of comet Austin (1989c1) were carried out by the Japanese CCD <span class="hlt">imaging</span> team. Six telescopes were used to monitor the time variation of the near-nucleus <span class="hlt">images</span> in C2, CN, H2O, and Na continuum in the optical region, and in J, H, and K bands in the near-IR region. A featureless, round shape of the comet was revealed in all <span class="hlt">images</span>. Although some of the jet features are recognized by using an <span class="hlt">image</span> enhancement technique, the azimuthal difference of the intensity distribution is about 10 percent. The <span class="hlt">images</span> in the H2O band show complex ion structures near the nucleus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/2134793','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/2134793"><span>[<span class="hlt">Observation</span> of oral actions using digital <span class="hlt">image</span> processing system].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ichikawa, T; Komoda, J; Horiuchi, M; Ichiba, H; Hada, M; Matsumoto, N</p> <p>1990-04-01</p> <p>A new digital <span class="hlt">image</span> processing system to <span class="hlt">observe</span> oral actions is proposed. The system provides analyses of motion pictures along with other physiological signals. The major components are a video tape recorder, a digital <span class="hlt">image</span> processor, a percept scope, a CCD camera, an A/D converter and a personal computer. Five reference points were marked on the lip and eyeglasses of 9 adult subjects. Lip movements were recorded and analyzed using the system when uttering five vowels and [ka, sa, ta, ha, ra, ma, pa, ba[. 1. Positions of the lip when uttering five vowels were clearly classified. 2. Active articulatory movements of the lip were not recognized when uttering consonants [k, s, t, h, r[. It seemed lip movements were dependent on tongue and mandibular movements. Downward and rearward movements of the upper lip, and upward and forward movements of the lower lip were <span class="hlt">observed</span> when uttering consonants [m, p, b[.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017yCat..51530071F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017yCat..51530071F"><span>VizieR Online Data Catalog: Kepler follow-up <span class="hlt">observation</span> program. I. <span class="hlt">Imaging</span> (Furlan+, 2017)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Furlan, E.; Ciardi, D. R.; Everett, M. E.; Saylors, M.; Teske, J. K.; Horch, E. P.; Howell, S. B.; van Belle, G. T.; Hirsch, L. A.; Gautier, T. N.; Adams, E. R.; Barrado, D.; Cartier, K. M. S.; Dressing, C. D.; Dupree, A. K.; Gilliland, R. L.; Lillo-Box, J.; Lucas, P. W.; Wang, J.</p> <p>2017-07-01</p> <p>We present results from six years of follow-up <span class="hlt">imaging</span> <span class="hlt">observations</span> of KOI host stars, including work done by teams from the Kepler Community Follow-up <span class="hlt">Observation</span> Program (CFOP; https://exofop.ipac.caltech.edu/cfop.php) and by other groups. Several <span class="hlt">observing</span> facilities were used to obtain high-resolution <span class="hlt">images</span> of KOI host stars. Table1 lists the various telescopes, instruments used, filter bandpasses, typical Point Spread Function (PSF) widths, number of targets <span class="hlt">observed</span>, and main references for the published results. The four main <span class="hlt">observing</span> techniques employed are adaptive optics (Keck, Palomar, Lick, MMT), speckle interferometry (Gemini North, WIYN, DCT), lucky <span class="hlt">imaging</span> (Calar Alto), and <span class="hlt">imaging</span> from space with HST. A total of 3557 KOI host stars were <span class="hlt">observed</span> at 11 facilities with 9 different instruments, using filters from the optical to the near-infrared. In addition, 10 of these stars were also <span class="hlt">observed</span> at the 8m Gemini North telescope by Ziegler et al. 2016 (AJ accepted, arXiv:1605.03584) using laser guide star adaptive optics. The largest number of KOI host stars (3320) were <span class="hlt">observed</span> using Robo-AO at the Palomar 1.5m telescope (Baranec et al. 2014ApJ...790L...8B; Baranec et al. 2016, Cat. J/AJ/152/18; Law et al. 2014, Cat. J/ApJ/791/35; Ziegler et al. 2016, AJ accepted, arXiv:1605.03584). A total of 8332 <span class="hlt">observations</span> were carried out from 2009 September to 2015 October covering 3557 stars. We carried out <span class="hlt">observations</span> at the Keck, Palomar, and Lick Observatory using the facility adaptive optics systems and near-infrared cameras from 2009 to 2015. At Keck, we <span class="hlt">observed</span> with the 10m Keck II telescope and Near-Infrared Camera, second generation (NIRC2). The pixel scale of NIRC2 was 0.01''/pixel, resulting in a field of view of about 10''*10''. We <span class="hlt">observed</span> our targets in a narrow K-band filter, Brγ, which has a central wavelength of 2.1686μm. In most cases, when a companion was detected, we also <span class="hlt">observed</span> the target in a narrow-band J filter, Jcont, which is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9037E..0OT','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9037E..0OT"><span>Design of a practical model-<span class="hlt">observer</span>-based <span class="hlt">image</span> quality assessment method for CT <span class="hlt">imaging</span> systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tseng, Hsin-Wu; Fan, Jiahua; Cao, Guangzhi; Kupinski, Matthew A.; Sainath, Paavana</p> <p>2014-03-01</p> <p>The channelized Hotelling <span class="hlt">observer</span> (CHO) is a powerful method for quantitative <span class="hlt">image</span> quality evaluations of CT systems and their <span class="hlt">image</span> reconstruction algorithms. It has recently been used to validate the dose reduction capability of iterative <span class="hlt">image</span>-reconstruction algorithms implemented on CT <span class="hlt">imaging</span> systems. The use of the CHO for routine and frequent system evaluations is desirable both for quality assurance evaluations as well as further system optimizations. The use of channels substantially reduces the amount of data required to achieve accurate estimates of <span class="hlt">observer</span> performance. However, the number of scans required is still large even with the use of channels. This work explores different data reduction schemes and designs a new approach that requires only a few CT scans of a phantom. For this work, the leave-one-out likelihood (LOOL) method developed by Hoffbeck and Landgrebe is studied as an efficient method of estimating the covariance matrices needed to compute CHO performance. Three different kinds of approaches are included in the study: a conventional CHO estimation technique with a large sample size, a conventional technique with fewer samples, and the new LOOL-based approach with fewer samples. The mean value and standard deviation of area under ROC curve (AUC) is estimated by shuffle method. Both simulation and real data results indicate that an 80% data reduction can be achieved without loss of accuracy. This data reduction makes the proposed approach a practical tool for routine CT system assessment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016DPS....4812350A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016DPS....4812350A"><span>Analyzing Serendipitous Asteroid <span class="hlt">Observations</span> in <span class="hlt">Imaging</span> Data using PHOTOMETRYPIPELINE</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ard, Christopher; Mommert, Michael; Trilling, David E.</p> <p>2016-10-01</p> <p>Asteroids are nearly ubiquitous in the night sky, making them present in the majority of <span class="hlt">imaging</span> data taken every night. Serendipitous asteroid <span class="hlt">observations</span> represent a treasure trove to Solar System researchers: accurate positional measurements of asteroids provide important constraints on their sometimes highly uncertain orbits, whereas calibrated photometric measurements can be used to establish rotational periods, intrinsic colors, or photometric phase curves.We present an add-on to the PHOTOMETRYPIPELINE (PP, github.com/mommermi/photometrypipeline, see Poster presentation 123.42) that identifies asteroids that have been <span class="hlt">observed</span> serendipitously and extracts astrometry and calibrated photometry for these objects. PP is an open-source Python 2.7 software suite that provides <span class="hlt">image</span> registration, aperture photometry, photometric calibration, and target identification with only minimal human interaction.Asteroids are identified based on approximate positions that are pre-calculated for a range of dates. Using interpolated coordinates, we identify potential asteroids that might be in the <span class="hlt">observed</span> field and query their exact positions and positional uncertainties from the JPL Horizons system. The method results in robust astrometry and calibrated photometry for all asteroids in the field as a function of time. Our measurements will supplement existing photometric databases of asteroids and improve their orbits.We present first results using this procedure based on <span class="hlt">imaging</span> data from the Vatican Advanced Technology Telescope.This work was done in the framework of NAU's REU summer program that is supported by NSF grant AST-1461200. PP was developed in the framework of the "Mission Accessible Near-Earth Object Survey" (MANOS) and is supported by NASA SSO grants NNX15AE90G and NNX14AN82G.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AdSpR..27.1165W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AdSpR..27.1165W"><span><span class="hlt">Observations</span> of OH(3,1) <span class="hlt">airglow</span> emission using a Michelson interferometer at 62° S</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Won, Young-In; Cho, Young-Min; Niciejewski, Rick J.; Kim, Jhoon</p> <p></p> <p>A Michelson interferometer was used to <span class="hlt">observe</span> the hydroxyl (OH) emission in the upper mesosphere at the King Sejong Station (62.22° S, 301.25° E), Antarctica. The instrument was installed in February 1999 and has been in routine operation since then. An intensive operational effort has resulted in a substantial data set between April and June, 1999. A spectral analysis was performed on individual data to examine the information of dominant waves. A harmonic analysis was also carried out on the monthly average data to investigate the characteristics of the major low frequency oscillations. The 12-hr temperature oscillations exhibit a striking agreement with a theoretical tidal model, supporting the tidal (migrating) origin. The 8-hr wave is found to be persistent and dominant, reflecting its major role in the upper mesospheric dynamics at the given latitude. The 6-hr oscillation is <span class="hlt">observed</span> only in May with its value close to the prediction for zonally symmetric tides.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880037707&hterms=sars&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsars','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880037707&hterms=sars&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dsars"><span><span class="hlt">Observation</span> of sea-ice dynamics using synthetic aperture radar <span class="hlt">images</span>: Automated analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vesecky, John F.; Samadani, Ramin; Smith, Martha P.; Daida, Jason M.; Bracewell, Ronald N.</p> <p>1988-01-01</p> <p>The European Space Agency's ERS-1 satellite, as well as others planned to follow, is expected to carry synthetic-aperture radars (SARs) over the polar regions beginning in 1989. A key component in utilization of these SAR data is an automated scheme for extracting the sea-ice velocity field from a time sequence of SAR <span class="hlt">images</span> of the same geographical region. Two techniques for automated sea-ice tracking, <span class="hlt">image</span> pyramid area correlation (hierarchical correlation) and feature tracking, are described. Each technique is applied to a pair of Seasat SAR sea-ice <span class="hlt">images</span>. The results compare well with each other and with manually tracked estimates of the ice velocity. The advantages and disadvantages of these automated methods are pointed out. Using these ice velocity field estimates it is possible to construct one sea-ice <span class="hlt">image</span> from the other member of the pair. Comparing the reconstructed <span class="hlt">image</span> with the <span class="hlt">observed</span> <span class="hlt">image</span>, errors in the estimated velocity field can be recognized and a useful probable error display created automatically to accompany ice velocity estimates. It is suggested that this error display may be useful in segmenting the sea ice <span class="hlt">observed</span> into regions that move as rigid plates of significant ice velocity shear and distortion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010019249','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010019249"><span>Earth <span class="hlt">Observing</span>-1 Advanced Land <span class="hlt">Imager</span>: Radiometric Response Calibration</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mendenhall, J. A.; Lencioni, D. E.; Evans, J. B.</p> <p>2000-01-01</p> <p>The Advanced Land <span class="hlt">Imager</span> (ALI) is one of three instruments to be flown on the first Earth <span class="hlt">Observing</span> mission (EO-1) under NASA's New Millennium Program (NMP). ALI contains a number of innovative features, including a wide field of view optical design, compact multispectral focal plane arrays, non-cryogenic HgCdTe detectors for the short wave infrared bands, and silicon carbide optics. This document outlines the techniques adopted during ground calibration of the radiometric response of the Advanced Land <span class="hlt">Imager</span>. Results from system level measurements of the instrument response, signal-to-noise ratio, saturation radiance, and dynamic range for all detectors of every spectral band are also presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss041e067595.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss041e067595.html"><span>Earth <span class="hlt">Observations</span> taken by Expedition 41 crewmember.</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2014-10-06</p> <p>ISS041-E-067595 (6 Oct. 2014) --- This moonlit panorama was shot recently with a wide-angle lens by an Expedition 41 crew member aboard the International Space Station, as they looked southwest from a point over Nebraska. The wide-angle lens shows a huge swath of country that stretches from Portland, Oregon (right) to Phoenix, Arizona (left). The largest string of lights is the Ogden-Salt Lake City-Provo area (lower center) in Utah. The Los Angeles and San Francisco metropolitan regions, and the cities of the central valley of California (Bakersfield to Redding) stretch across the horizon. The green <span class="hlt">airglow</span> layer always appears in night <span class="hlt">images</span>. Moonlight shows the red tinge of the space station?s solar arrays top left. Moonlight emphasizes the broader-scale geological zones. Nevada?s short, dark, parallel mountain ranges of the basin and range geological province (center) contrast with the expanses of flat terrain of the Colorado Plateau (left) in Colorado, Arizona, Utah and New Mexico. The near-full moon even reveals the vast dry lake bed known as the Bonneville Salt Flats. The black line of the Sierra Nevada marks the edge of California?s well-lit central valley (directly below the San Francisco Bay area).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910069026&hterms=L37&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DL37','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910069026&hterms=L37&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DL37"><span><span class="hlt">Observations</span> of Comet Levy (1990c) with the Hopkins Ultraviolet Telescope</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Feldman, P. D.; Davidsen, A. F.; Blair, W. P.; Bowers, C. W.; Dixon, W. V.; Durrance, S. T.; Henry, R. C.; Ferguson, H. C.; Kimble, R. A.; Gull, Theodore R.</p> <p>1991-01-01</p> <p><span class="hlt">Observations</span> of Comet Levy (1990c) were made with the Hopkins Ultraviolet Telescope during the Astro-1 Space Shuttle mission on December 10, 1990. The spectrum, covering the wavelength range 415-1850 A at a spectral resolution of 3 A, shows the presence of carbon monoxide and atomic hydrogen, carbon, and sulfur in the coma. Aside from H I Lyman-beta, no cometary features are detected below 1200 A, although cometary O I and O II would be masked by the same emissions present in the day <span class="hlt">airglow</span> spectrum. The 9.4 x 116 arcsecond aperture corresponds to 12,000 x 148,000 km at the comet. The derived production rate of CO relative to water is 0.11 + or - 0.02, compared with 0.04 + or - 0.01 derived from IUE <span class="hlt">observations</span> (made in September 1990) which sample a much smaller region of the coma. This suggests the presence of an extended source of CO, as was found in comet Halley. Upper limits on Ne and Ar abundance are within one order of magnitude of solar abundances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMSA12A..08S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMSA12A..08S"><span>Remote sensing of the low-latitude daytime ionosphere: ICON simulations and retrievals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stephan, A. W.; Korpela, E.; England, S.; Immel, T. J.</p> <p>2016-12-01</p> <p>The Ionospheric Connection Explorer (ICON) sensor suite includes a spectrograph that will provide altitude profiles of the OII 61.7 and 83.4 nm <span class="hlt">airglow</span> features, from which the daytime F-region ionosphere can be inferred. To make the connection between these extreme-ultraviolet (EUV) <span class="hlt">airglow</span> emissions and ionospheric densities, ICON will use a method that has matured significantly in the last decade with the analysis of data from the Remote Atmospheric and Ionospheric Detection System (RAIDS) on the International Space Station, and the Special Sensor Ultraviolet Limb <span class="hlt">Imager</span> (SSULI) sensors on the Defense Meteorological Satellite Program (DMSP) series of satellites. We will present end-to-end simulations of ICON EUV <span class="hlt">airglow</span> measurements and data inversion for the expected viewing geometry and sensor capabilities, including noise. While we will focus on the performance of the algorithm for ICON within the context of the current state of knowledge, we will also identify areas where fundamental information can be gained from the high-sensitivity ICON measurements that could be used as feedback to directly improve the overall performance of the algorithm itself.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24514183','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24514183"><span>Nonlinear regression method for estimating neutral wind and temperature from Fabry-Perot interferometer data.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Harding, Brian J; Gehrels, Thomas W; Makela, Jonathan J</p> <p>2014-02-01</p> <p>The Earth's thermosphere plays a critical role in driving electrodynamic processes in the ionosphere and in transferring solar energy to the atmosphere, yet measurements of thermospheric state parameters, such as wind and temperature, are sparse. One of the most popular techniques for measuring these parameters is to use a Fabry-Perot interferometer to monitor the Doppler width and breadth of naturally occurring <span class="hlt">airglow</span> emissions in the thermosphere. In this work, we present a technique for estimating upper-atmospheric winds and temperatures from <span class="hlt">images</span> of Fabry-Perot fringes captured by a CCD detector. We estimate instrument parameters from fringe patterns of a frequency-stabilized laser, and we use these parameters to estimate winds and temperatures from <span class="hlt">airglow</span> fringe patterns. A unique feature of this technique is the model used for the laser and <span class="hlt">airglow</span> fringe patterns, which fits all fringes simultaneously and attempts to model the effects of optical defects. This technique yields accurate estimates for winds, temperatures, and the associated uncertainties in these parameters, as we show with a Monte Carlo simulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40833&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40833&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dphoto%2Bimage"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Earth <span class="hlt">observation</span> photo taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). This <span class="hlt">image</span> shows a 50 by 120 kilometer (30 by 75 mile) area of the Mediterranean Sea and the eastern coast of Central Sardinia (left). The city of Arbatose is seen as a bright area along the coast in the lower part of the <span class="hlt">image</span>, and the star-like spot off the coast is a ship's reflection. The Gulf of Orsei is near the top of the <span class="hlt">image</span>. Bright, mottled features in the sea (right) represent surface choppiness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10136E..10H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10136E..10H"><span>The disagreement between the ideal <span class="hlt">observer</span> and human <span class="hlt">observers</span> in hardware and software <span class="hlt">imaging</span> system optimization: theoretical explanations and evidence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>He, Xin</p> <p>2017-03-01</p> <p>The ideal <span class="hlt">observer</span> is widely used in <span class="hlt">imaging</span> system optimization. One practical question remains open: do the ideal and human <span class="hlt">observers</span> have the same preference in system optimization and evaluation? Based on the ideal <span class="hlt">observer</span>'s mathematical properties proposed by Barrett et. al. and the empirical properties of human <span class="hlt">observers</span> investigated by Myers et. al., I attempt to pursue the general rules regarding the applicability of the ideal <span class="hlt">observer</span> in system optimization. Particularly, in software optimization, the ideal <span class="hlt">observer</span> pursues data conservation while humans pursue data presentation or perception. In hardware optimization, the ideal <span class="hlt">observer</span> pursues a system with the maximum total information, while humans pursue a system with the maximum selected (e.g., certain frequency bands) information. These different objectives may result in different system optimizations between human and the ideal <span class="hlt">observers</span>. Thus, an ideal <span class="hlt">observer</span> optimized system is not necessarily optimal for humans. I cite empirical evidence in search and detection tasks, in hardware and software evaluation, in X-ray CT, pinhole <span class="hlt">imaging</span>, as well as emission computed tomography to corroborate the claims. (Disclaimer: the views expressed in this work do not necessarily represent those of the FDA)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40832&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40832&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Earth <span class="hlt">observation</span> photo taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). This <span class="hlt">image</span> shows the Los Angeles basin. The area's freeways are visible as dark lines. The Los Angles harbor breakwater off Long Beach is seen as a bright line. Vessels in the harbor show as bright points.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014cosp...40E3685Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014cosp...40E3685Y"><span>Study of medium-scale traveling ionospheric disturbances (MSTID) with sounding rockets and ground <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>Yamamoto, Mamoru; Abe, Takumi; Kumamoto, Atsushi; Yokoyama, Tatsuhiro; Bernhardt, Paul; Watanabe, Shigeto; Yamamoto, Masa-yuki; Larsen, Miguel; Saito, Susumu; Tsugawa, Takuya; Ishisaka, Keigo; Iwagami, Naomoto; Nishioka, Michi; Kato, Tomohiro; Takahashi, Takao; Tanaka, Makoto; Mr</p> <p></p> <p>Medium-scale traveling ionospheric disturbance (MSTID) is an interesting phenomenon in the F-region. The MSTID is frequent in summer nighttime over Japan, showing wave structures with wavelengths of 100-200 km, periodicity of about 1 hour, and propagation toward the southwest. The phenomena are <span class="hlt">observed</span> by the total electron content (TEC) from GEONET, Japanese dense network of GPS receivers, and 630 nm <span class="hlt">airglow</span> <span class="hlt">imagers</span> as horizontal pattern. It was also measured as Spread-F events of ionograms or as field-aligned echoes of the MU radar. MSTID was, in the past, explained by Perkins instability (Perkins, 1973) while its low growth rate was a problem. Recently 3D simulation study by Yokoyama et al (2009) hypothesized a generation mechanism of the MSTID, which stands on electromagnetic E/F-region coupling of the ionosphere. The hypothesis is that the MSTID first grows with polarization electric fields from sporadic-E, then show spatial structures resembling to the Perkins instability. We recently conducted a <span class="hlt">observation</span> campaign to check this hypothesis. We launched JASA ISAS sounding rockets S-310-42 and S-520-27 at 23:00 JST and 23:57JST on July 20, 2013 while an MSTID event was monitored in real-time by the GPS-TEC from GEONET. We found 1-5mV/m northeastward/eastward electric fields during the flight. Variation of electric fileds were associated with horizontal distribution of plasma density. Wind velocity was measured by the TME and Lithium releases from S-310-42 and S-520-27 rockets, respectively, showing southward wind near the sporadic-E layer heights. These results are consistent to the expected generation mechanism shown above. In the presentation we will discuss electric-field results and its relationship with plasma density variability together with preliminary results from the neutral-wind <span class="hlt">observations</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_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://hdl.handle.net/2060/19900013510','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19900013510"><span>Spaceborne radar <span class="hlt">observations</span>: A guide for Magellan radar-<span class="hlt">image</span> analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ford, J. P.; Blom, R. G.; Crisp, J. A.; Elachi, Charles; Farr, T. G.; Saunders, R. Stephen; Theilig, E. E.; Wall, S. D.; Yewell, S. B.</p> <p>1989-01-01</p> <p>Geologic analyses of spaceborne radar <span class="hlt">images</span> of Earth are reviewed and summarized with respect to detecting, mapping, and interpreting impact craters, volcanic landforms, eolian and subsurface features, and tectonic landforms. Interpretations are illustrated mostly with Seasat synthetic aperture radar and shuttle-<span class="hlt">imaging</span>-radar <span class="hlt">images</span>. Analogies are drawn for the potential interpretation of radar <span class="hlt">images</span> of Venus, with emphasis on the effects of variation in Magellan look angle with Venusian latitude. In each landform category, differences in feature perception and interpretive capability are related to variations in <span class="hlt">imaging</span> geometry, spatial resolution, and wavelength of the <span class="hlt">imaging</span> radar systems. Impact craters and other radially symmetrical features may show apparent bilateral symmetry parallel to the illumination vector at low look angles. The styles of eruption and the emplacement of major and minor volcanic constructs can be interpreted from morphological features <span class="hlt">observed</span> in <span class="hlt">images</span>. Radar responses that are governed by small-scale surface roughness may serve to distinguish flow types, but do not provide unambiguous information. <span class="hlt">Imaging</span> of sand dunes is rigorously constrained by specific angular relations between the illumination vector and the orientation and angle of repose of the dune faces, but is independent of radar wavelength. With a single look angle, conditions that enable shallow subsurface <span class="hlt">imaging</span> to occur do not provide the information necessary to determine whether the radar has recorded surface or subsurface features. The topographic linearity of many tectonic landforms is enhanced on <span class="hlt">images</span> at regional and local scales, but the detection of structural detail is a strong function of illumination direction. Nontopographic tectonic lineaments may appear in response to contrasts in small-surface roughness or dielectric constant. The breakpoint for rough surfaces will vary by about 25 percent through the Magellan viewing geometries from low to high</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007SPIE.6826E..2PO','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007SPIE.6826E..2PO"><span>Methods on <span class="hlt">observation</span> of fluorescence micro-<span class="hlt">imaging</span> for microalgae</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ou, Lin; Zhuang, Hui-ru; Chen, Rong; Lei, Jin-pin; Liao, Xiao-hua; Lin, Wen-suo</p> <p>2007-11-01</p> <p>Objective: Auto-fluorescence micro-<span class="hlt">imaging</span> of microalgae are <span class="hlt">observed</span> by using of laser scanning confocal microscopy (LSCM) and fluorescence microscopy, so as to investigate the effect of auto fluorescence alteration on growth of irradiated microalgae irradiated, meanwhile, the method of microalgae cells stained also to be studied. Methods: Platymonas subcordiformis, Phaeodactylum tricormutum and Isochyrsis zhanjiangensis cells are stained with acridine orange, and <span class="hlt">observed</span> by fluorescence microscopy; the three types microalgae mentioned above are irradiated by Nd:YAP laser with 10w at 1341nm, irradiating time:12s, 30s, 35s and 55s, than to be cultured 6 days, and the auto fluorescence <span class="hlt">images</span> and fluorescence spectra of algae cells are obtained by LSCM on lambda scan mode, at excitation 488nm (Ar + laser). Results: It is showed that the shapes and the structural features of microalgae cells stained can be seen clearly, and the cytoplasm and nucleus also can be <span class="hlt">observed</span>. The chloroplasts in cell is bigger on promoting effects, conversely, it is to be mutilated, deformation and shrink. Contrast to the CK, the peak positions of fluorescence of algae cells irradiated is similar to the whole while the peak light intensity alters. On irradiation of promoting dose, however, the auto fluorescence intensity is enhanced more than control. Conclusions: The method of cell stained can be used to <span class="hlt">observed</span> genetic material in microalgae. There are obvious effects for laser irradiating to chloroplasts in cells, the bigger chloroplasts the greater fluorescence intensity. Physiological incentive effects of microalgae irradiated can be given expression on fluorescence characteristics and fluorescence intensity alteration of cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000AAS...197.5502C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000AAS...197.5502C"><span>Polarization <span class="hlt">Observations</span> with the Cosmic Background <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cartwright, J. K.; Padin, S.; Pearson, T. J.; Readhead, A. C. S.; Shepherd, M. C.; Taylor, G. B.</p> <p>2000-12-01</p> <p>The linear polarization of the Cosmic Microwave Background Radiation is a fundamental prediction of the standard model. We report a limit on the polarization of the CMBR for l ~660. This limit was obtained with the Cosmic Background <span class="hlt">Imager</span>, a 13 element interferometer which operates in the 26-36 GHz band from a site at 5000m in northern Chile. The array consists of 90-cm Cassegrain antennas mounted on a single, fully steerable platform; this platform can be rotated about the optical axis to facilitate polarization <span class="hlt">observations</span>. The CBI employs single mode circularly polarized receivers, of which 12 are configured for LCP and one is configured for RCP. The 12 cross polarized baselines sample multipoles from l ~600 to l ~3500. The instrumental polarization of the CBI was calibrated with <span class="hlt">observations</span> of 3C279, a bright polarized source which is unresolved by the CBI. Because the centimeter flux of 3C279 is variable, it was monitored twice per month for 8 months in '00 with the VLA at 22 and 43 GHz. These <span class="hlt">observations</span> also established the stability of the polarization characteristics of the CBI. This work was made possible by NSF grant AST-9802989</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40829&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40829&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Photos of earth <span class="hlt">observations</span> taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). This <span class="hlt">image</span> shows Lake Okeechobee (right) and Lake Istokopoga (left) in Central Florida. Lake Okeechobee is bounded on the east by rectangular agricultural fields and to the south and west by swamps and wetlands which appear as bright features.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9100E..04B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9100E..04B"><span>Performance of PHOTONIS' low light level CMOS <span class="hlt">imaging</span> sensor for long range <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bourree, Loig E.</p> <p>2014-05-01</p> <p>Identification of potential threats in low-light conditions through <span class="hlt">imaging</span> is commonly achieved through closed-circuit television (CCTV) and surveillance cameras by combining the extended near infrared (NIR) response (800-10000nm wavelengths) of the <span class="hlt">imaging</span> sensor with NIR LED or laser illuminators. Consequently, camera systems typically used for purposes of long-range <span class="hlt">observation</span> often require high-power lasers in order to generate sufficient photons on targets to acquire detailed <span class="hlt">images</span> at night. While these systems may adequately identify targets at long-range, the NIR illumination needed to achieve such functionality can easily be detected and therefore may not be suitable for covert applications. In order to reduce dependency on supplemental illumination in low-light conditions, the frame rate of the <span class="hlt">imaging</span> sensors may be reduced to increase the photon integration time and thus improve the signal to noise ratio of the <span class="hlt">image</span>. However, this may hinder the camera's ability to <span class="hlt">image</span> moving objects with high fidelity. In order to address these particular drawbacks, PHOTONIS has developed a CMOS <span class="hlt">imaging</span> sensor (CIS) with a pixel architecture and geometry designed specifically to overcome these issues in low-light level <span class="hlt">imaging</span>. By combining this CIS with field programmable gate array (FPGA)-based <span class="hlt">image</span> processing electronics, PHOTONIS has achieved low-read noise <span class="hlt">imaging</span> with enhanced signal-to-noise ratio at quarter moon illumination, all at standard video frame rates. The performance of this CIS is discussed herein and compared to other commercially available CMOS and CCD for long-range <span class="hlt">observation</span> applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22522212-dynamics-disk-plumes-observed-interface-region-imaging-spectrograph-atmospheric-imaging-assembly-helioseismic-magnetic-imager','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22522212-dynamics-disk-plumes-observed-interface-region-imaging-spectrograph-atmospheric-imaging-assembly-helioseismic-magnetic-imager"><span>DYNAMICS OF ON-DISK PLUMES AS <span class="hlt">OBSERVED</span> WITH THE INTERFACE REGION <span class="hlt">IMAGING</span> SPECTROGRAPH, THE ATMOSPHERIC <span class="hlt">IMAGING</span> ASSEMBLY, AND THE HELIOSEISMIC AND MAGNETIC <span class="hlt">IMAGER</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>Pant, Vaibhav; Mazumder, Rakesh; Banerjee, Dipankar</p> <p>2015-07-01</p> <p>We examine the role of small-scale transients in the formation and evolution of solar coronal plumes. We study the dynamics of plume footpoints seen in the vicinity of a coronal hole using the Atmospheric <span class="hlt">Imaging</span> Assembly (AIA) <span class="hlt">images</span>, the Helioseismic and Magnetic <span class="hlt">Imager</span> magnetogram on board the Solar Dynamics Observatory and spectroscopic data from the Interface Region <span class="hlt">Imaging</span> Spectrograph (IRIS). Quasi-periodic brightenings are <span class="hlt">observed</span> in the base of the plumes and are associated with magnetic flux changes. With the high spectral and spatial resolution of IRIS, we identify the sources of these oscillations and try to understand what role themore » transients at the footpoints can play in sustaining the coronal plumes. IRIS “sit-and-stare” <span class="hlt">observations</span> provide a unique opportunity to study the evolution of footpoints of the plumes. We notice enhanced line width and intensity, and large deviation from the average Doppler shift in the line profiles at specific instances, which indicate the presence of flows at the footpoints of plumes. We propose that outflows (jet-like features) as a result of small-scale reconnections affect the line profiles. These jet-like features may also be responsible for the generation of propagating disturbances (PDs) within the plumes, which are <span class="hlt">observed</span> to be propagating to larger distances as recorded from multiple AIA channels. These PDs can be explained in terms of slow magnetoacoustic waves.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040074235&hterms=energy+solar&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Denergy%2Bsolar','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040074235&hterms=energy+solar&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Denergy%2Bsolar"><span>Solar Flares <span class="hlt">Observed</span> with the Ramaty High Energy Solar Spectroscopic <span class="hlt">Imager</span> (RHESSI)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Holman, Gordon D.</p> <p>2004-01-01</p> <p>Solar flares are impressive examples of explosive energy release in unconfined, magnetized plasma. It is generally believed that the flare energy is derived from the coronal magnetic field. However, we have not been able to establish the specific energy release mechanism(s) or the relative partitioning of the released energy between heating, particle acceleration (electrons and ions), and mass motions. NASA's RHESSI Mission was designed to study the acceleration and evolution of electrons and ions in flares by <span class="hlt">observing</span> the X-ray and gamma-ray emissions these energetic particles produce. This is accomplished through the combination of high-resolution spectroscopy and spectroscopic <span class="hlt">imaging</span>, including the first <span class="hlt">images</span> of flares in gamma rays. RHESSI has <span class="hlt">observed</span> over 12,000 solar flares since its launch on February 5, 2002. I will demonstrate how we use the RHESSI spectra to deduce physical properties of accelerated electrons and hot plasma in flares. Using <span class="hlt">images</span> to estimate volumes, w e typically find that the total energy in accelerated electrons is comparable to that in the thermal plasma. I will also present flare <span class="hlt">observations</span> that provide strong support for the presence of magnetic reconnection in a large-scale, vertical current sheet in the solar corona. RHESSI <span class="hlt">observations</span> such as these are allowing us to probe more deeply into the physics of solar flares.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24925100','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24925100"><span>Fine structure in diabetic retinopathy lesions as <span class="hlt">observed</span> by adaptive optics <span class="hlt">imaging</span>. A qualitative study.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bek, Toke</p> <p>2014-12-01</p> <p>Diabetic retinopathy is diagnosed by fundus photography and optical coherence tomography (OCT) scanning. However, adaptive optics (AO) <span class="hlt">imaging</span> can be expected to add new aspects to the knowledge of diabetic retinopathy because photographic resolution is improved by reducing the influence of optical aberrations on retinal <span class="hlt">imaging</span>. Nineteen patients with diabetes mellitus were subjected to fundus photography, OCT scanning and AO <span class="hlt">imaging</span>. The fundus photographs were scaled to the same magnification as that of the AO <span class="hlt">image</span>, and qualitative aspects of AO <span class="hlt">images</span> of each retinopathy lesion <span class="hlt">observed</span> on fundus photographs and OCT scans were assessed. All red lesions on fundus photographs appeared on AO <span class="hlt">images</span> as dark hyporeflective elements, but it could not be verified whether lesions represented haemorrhages or microaneurysms. The smallest of these lesions were circular with a size corresponding to that of blood cells. Hard exudates had irregular surfaces with buddings of various sizes protruding from the lesions. Areas of retinal oedema <span class="hlt">observed</span> by fundus <span class="hlt">imaging</span> and OCT scanning resulted in blurring of AO <span class="hlt">images</span>, but cystoid spaces <span class="hlt">observed</span> by OCT could be seen on AO <span class="hlt">images</span> to have a sharp delimitation with a darker hyporeflective rim at the internal lining of the cyst wall. AO <span class="hlt">imaging</span> may potentially assist in detecting diabetic retinopathy at an earlier stage, may help elucidating the pathophysiology of the diseases and may be used for evaluating the effects of clinical interventions on diabetic retinopathy and other retinal vascular diseases. © 2014 Acta Ophthalmologica Scandinavica Foundation. Published by John Wiley & Sons Ltd.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMAE43A0245S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMAE43A0245S"><span>Monochromatic <span class="hlt">imaging</span> <span class="hlt">observation</span> of sprites with the Reimei satellite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sakanoi, T.; Adachi, T.; Sato, M.; Yamazaki, A.; Asamura, K.; Hirahara, M.</p> <p>2012-12-01</p> <p>The sprite emission is characterized by vertically extending fine structures (called as streamer, halo, etc.) in the approximate altitude range from 40 to 90 km. Satellite <span class="hlt">observation</span> is useful to investigate the global distributions of sprite since an optical instrument on a satellite can measure the sprite in the wide range without atmospheric absorption. However, the sprite has been measured mainly by ground-based instruments, and there is no monochromatic <span class="hlt">imaging</span> data from space. The multi-spectral camera (MAC) on Reimei was originally designed to measure auroral emissions in the polar region taking the monochromatic <span class="hlt">images</span> at wavelengths of 428 nm, 558 nm, and 670 nm. Since March 2008, MAC has been operated in the mid- and low-latitudes viewing the limb direction with an exposure time of 957 ms to measure the monochromatic <span class="hlt">image</span> of sprite emission. The spatial resolution at a tangential point is approximately 4 km. According to the noon-midnight sun-synchronous orbit of Reimei at an altitude of 640 km, the <span class="hlt">observation</span> is made around the midnight sector. So far, we found seven sprits events in N2 1P (670 nm) <span class="hlt">images</span>, and on six events the simultaneous <span class="hlt">observations</span> between N2+ 1N (428nm) and N2 1P were performed. The electron temperature and electric field associated with a sprite can be estimated from the intensity ratio between emission of N2+ 1N and that of N2 1P. However, we did not obtain the N2+ 1N emission intensity due to the low sensitivity of 428 nm channel of MAC. Therefore, the N2+ 1N intensities of sprites are estimated to be less than the noise level (26 - 54 R), while the measured N2 1P intensities of sprites are 2.9 - 3.6 kR. Using these data, we estimated the upper limit of electron temperature and electric field associated with sprites. The altitude of sprite emission was accurately determined with the satellite attitude data and the field-of-view direction of MAC. On the 2008 Sep.2 case, we obtained sprite events at 26.6 GLAT and 107.6 GELON</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss030e060478.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss030e060478.html"><span>Earth <span class="hlt">Observations</span> taken by Expedition 30 crewmember</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2012-01-30</p> <p>ISS030-E-060478 (30 Jan. 2012) --- The city lights of Madrid (just right of center) stand out in this photograph from the International Space Station. Recorded by one of the Expedition 30 crew members, the view shows almost the entire Iberian Peninsula (both Spain and Portugal) with the Strait of Gibraltar and Morocco appearing at lower left. What is thought to be a blur of the moon appears in upper left corner. The faint gold or brownish line of <span class="hlt">airglow</span>?caused by ultraviolet radiation exciting the gas molecules in the upper atmosphere?parallels the horizon or Earth limb.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040110411&hterms=image+processing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dimage%2Bprocessing','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040110411&hterms=image+processing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dimage%2Bprocessing"><span>Retinex <span class="hlt">Image</span> Processing: Improved Fidelity To Direct Visual <span class="hlt">Observation</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jobson, Daniel J.; Rahman, Zia-Ur; Woodell, Glenn A.</p> <p>1996-01-01</p> <p>Recorded color <span class="hlt">images</span> differ from direct human viewing by the lack of dynamic range compression and color constancy. Research is summarized which develops the center/surround retinex concept originated by Edwin Land through a single scale design to a multi-scale design with color restoration (MSRCR). The MSRCR synthesizes dynamic range compression, color constancy, and color rendition and, thereby, approaches fidelity to direct <span class="hlt">observation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss025e009858.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss025e009858.html"><span>Night Earth <span class="hlt">Observation</span> taken by the Expedition 25 crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2010-10-28</p> <p>ISS025-E-009858 (28 Oct. 2010) --- From 220 miles above Earth, one of the Expedition 25 crew members on the International Space Station took this night time photo featuring the bright lights of Cairo and Alexandria, Egypt on the Mediterranean coast. The Nile River and its delta stand out clearly as well. On the horizon, the <span class="hlt">airglow</span> of the atmosphere is seen across the Mediterranean. The Sinai Peninsula, at right, is outlined with lights highlighting the Gulf of Suez and Gulf of Aqaba.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20060000018','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20060000018"><span>Solid Hydrogen Experiments for Atomic Propellants: Particle Formation, <span class="hlt">Imaging</span>, <span class="hlt">Observations</span>, and Analyses</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Palaszewski, Bryan</p> <p>2005-01-01</p> <p>This report presents particle formation <span class="hlt">observations</span> and detailed analyses of the <span class="hlt">images</span> from experiments that were conducted on the formation of solid hydrogen particles in liquid helium. Hydrogen was frozen into particles in liquid helium, and <span class="hlt">observed</span> with a video camera. The solid hydrogen particle sizes and the total mass of hydrogen particles were estimated. These newly analyzed data are from the test series held on February 28, 2001. Particle sizes from previous testing in 1999 and the testing in 2001 were similar. Though the 2001 testing created similar particles sizes, many new particle formation phenomena were <span class="hlt">observed</span>: microparticles and delayed particle formation. These experiment <span class="hlt">image</span> analyses are some of the first steps toward visually characterizing these particles, and they allow designers to understand what issues must be addressed in atomic propellant feed system designs for future aerospace vehicles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA21A2503F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA21A2503F"><span>Estimation of 557.7 nm Emission Altitude using Co-located Lidars and Photometers over Arecibo</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Franco, E.; Raizada, S.; Lautenbach, J.; Brum, C. G. M.</p> <p>2017-12-01</p> <p><span class="hlt">Airglow</span> at 557.7 nm (green line emission) is generated through the Barth mechanism in the E-region altitude and is sometimes associated with red line (630.0 nm) originating at F-region altitudes. Photons at 557.7 nm are produced through the quenching of excited atomic oxygen atoms, O(1S), while 630.0 nm results through the de-excitation of O(1D) atoms. Even though, the contribution of the green line from F-region is negligible and the significant component comes from the mesosphere, this uncertainty gives rise to a question related to its precise altitude. Previous studies have shown that perturbations generated by atmospheric gravity Waves (GWs) alter the <span class="hlt">airglow</span> intensity and can be used for studying dynamics of the region where it originates. The uncertainty in the emission altitude of green line can be resolved by using co-located lidars, which provide altitude resolved metal densities. At Arecibo, the resonance lidars tuned to Na and K resonance wavelengths at 589 nm and 770 nm can be used in conjunction with simultaneous measurements from green line photometer to resolve this issue. Both photometer and lidars have narrow field of view as compared to <span class="hlt">airglow</span> <span class="hlt">imagers</span>, and hence provide an added advantage that these instruments sample same GW spectrum. Hence, correlation between density perturbations inferred from lidars and <span class="hlt">airglow</span> intensity perturbations can shed light on the exact altitude of green line emission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4239938','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4239938"><span>High-Precision <span class="hlt">Image</span> Aided Inertial Navigation with Known Features: <span class="hlt">Observability</span> Analysis and Performance Evaluation</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jiang, Weiping; Wang, Li; Niu, Xiaoji; Zhang, Quan; Zhang, Hui; Tang, Min; Hu, Xiangyun</p> <p>2014-01-01</p> <p>A high-precision <span class="hlt">image</span>-aided inertial navigation system (INS) is proposed as an alternative to the carrier-phase-based differential Global Navigation Satellite Systems (CDGNSSs) when satellite-based navigation systems are unavailable. In this paper, the <span class="hlt">image</span>/INS integrated algorithm is modeled by a tightly-coupled iterative extended Kalman filter (IEKF). Tightly-coupled integration ensures that the integrated system is reliable, even if few known feature points (i.e., less than three) are <span class="hlt">observed</span> in the <span class="hlt">images</span>. A new global <span class="hlt">observability</span> analysis of this tightly-coupled integration is presented to guarantee that the system is <span class="hlt">observable</span> under the necessary conditions. The analysis conclusions were verified by simulations and field tests. The field tests also indicate that high-precision position (centimeter-level) and attitude (half-degree-level)-integrated solutions can be achieved in a global reference. PMID:25330046</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22398947-experimental-observation-sub-rayleigh-quantum-imaging-two-photon-entangled-source','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22398947-experimental-observation-sub-rayleigh-quantum-imaging-two-photon-entangled-source"><span>Experimental <span class="hlt">observation</span> of sub-Rayleigh quantum <span class="hlt">imaging</span> with a two-photon entangled source</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>Xu, De-Qin; School of Science, Tianjin University of Technology and Education, Tianjin 300222; Song, Xin-Bing</p> <p></p> <p>It has been theoretically predicted that N-photon quantum <span class="hlt">imaging</span> can realize either an N-fold resolution improvement (Heisenberg-like scaling) or a √(N)-fold resolution improvement (standard quantum limit) beyond the Rayleigh diffraction bound, over classical <span class="hlt">imaging</span>. Here, we report the experimental study on spatial sub-Rayleigh quantum <span class="hlt">imaging</span> using a two-photon entangled source. Two experimental schemes are proposed and performed. In a Fraunhofer diffraction scheme with a lens, two-photon Airy disk pattern is <span class="hlt">observed</span> with subwavelength diffraction property. In a lens <span class="hlt">imaging</span> apparatus, however, two-photon sub-Rayleigh <span class="hlt">imaging</span> for an object is realized with super-resolution property. The experimental results agree with the theoretical predictionmore » in the two-photon quantum <span class="hlt">imaging</span> regime.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090006704&hterms=michael+porter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmichael%2Bporter','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090006704&hterms=michael+porter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dmichael%2Bporter"><span>The Sheath Transport <span class="hlt">Observer</span> for the Redistribution of Mass (STORM) <span class="hlt">Image</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kuntz, Kip; Collier, Michael; Sibeck, David G.; Porter, F. Scott; Carter, J. A.; Cravens, Thomas; Omidi, N.; Robertson, Ina; Sembay, S.; Snowden, Steven L.</p> <p>2008-01-01</p> <p>All of the solar wind energy that powers magnetospheric processes passes through the magnetosheath and magnetopause. Global <span class="hlt">images</span> of the magnetosheath and magnetopause boundary layers will resolve longstanding controversy surrounding fundamental phenomena that occur at the magnetopause and provide information needed to improve operational space weather models. Recent developments showing that soft X-rays (0.15-1 keV) result from high charge state solar wind ions undergoing charge exchange recombination through collisions with exospheric neutral atoms has led to the realization that soft X-ray <span class="hlt">imaging</span> can provide global maps of the high-density shocked solar wind within the magnetosheath and cusps, regions lying between the lower density solar wind and magnetosphere. We discuss an instrument concept called the Sheath Transport <span class="hlt">Observer</span> for the Redistribution of Mass (STORM), an X-ray <span class="hlt">imager</span> suitable for simultaneously <span class="hlt">imaging</span> the dayside magnetosheath, the magnetopause boundary layers, and the cusps.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110005606','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110005606"><span>The Sheath Transport <span class="hlt">Observer</span> for the Redistribution of Mass (STORM) <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Collier, Michael R.; Sibeck, David G.; Porter, F. Scott; Burch, J.; Carter, J. A.; Cravens, Thomas; Kuntz, Kip; Omidi, N.; Read, A.; Robertson, Ina; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20110005606'); toggleEditAbsImage('author_20110005606_show'); toggleEditAbsImage('author_20110005606_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20110005606_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20110005606_hide"></p> <p>2010-01-01</p> <p>All of the solar wind energy that powers magnetospheric processes passes through the magnetosheath and magnetopause. Global <span class="hlt">images</span> of the magnetosheath and magnetopause boundary layers will resolve longstanding controversies surrounding fundamental phenomena that occur at the magnetopause and provide information needed to improve operational space weather models. Recent developments showing that soft X-rays (0.15-1 keV) result from high charge state solar wind ions undergoing charge exchange recombination through collisions with exospheric neutral atoms has led to the realization that soft X-ray <span class="hlt">imaging</span> can provide global maps of the high-density shocked solar wind within the magnetosheath and cusps, regions lying between the lower density solar wind and magnetosphere. We discuss an instrument concept called the Sheath Transport <span class="hlt">Observer</span> for the Redistribution of Mass (STORM), an X-ray <span class="hlt">imager</span> suitable for simultaneously <span class="hlt">imaging</span> the dayside magnetosheath, the magnetopause boundary layers, and the cusps.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9301E..3DD','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9301E..3DD"><span>Center determination for trailed sources in astronomical <span class="hlt">observation</span> <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Du, Jun Ju; Hu, Shao Ming; Chen, Xu; Guo, Di Fu</p> <p>2014-11-01</p> <p><span class="hlt">Images</span> with trailed sources can be obtained when <span class="hlt">observing</span> near-Earth objects, such as small astroids, space debris, major planets and their satellites, no matter the telescopes track on sidereal speed or the speed of target. The low centering accuracy of these trailed sources is one of the most important sources of the astrometric uncertainty, but how to determine the central positions of the trailed sources accurately remains a significant challenge to <span class="hlt">image</span> processing techniques, especially in the study of faint or fast moving objects. According to the conditions of one-meter telescope at Weihai Observatory of Shandong University, moment and point-spread-function (PSF) fitting were chosen to develop the <span class="hlt">image</span> processing pipeline for space debris. The principles and the implementations of both two methods are introduced in this paper. And some simulated <span class="hlt">images</span> containing trailed sources are analyzed with each technique. The results show that two methods are comparable to obtain the accurate central positions of trailed sources when the signal to noise (SNR) is high. But moment tends to fail for the objects with low SNR. Compared with moment, PSF fitting seems to be more robust and versatile. However, PSF fitting is quite time-consuming. Therefore, if there are enough bright stars in the field, or the high astronometric accuracy is not necessary, moment is competent. Otherwise, the combination of moment and PSF fitting is recommended.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSH51B2492B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSH51B2492B"><span>Simulating multi-spacecraft Heliospheric <span class="hlt">Imager</span> <span class="hlt">observations</span> for tomographic reconstruction of interplanetary CMEs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barnes, D.</p> <p>2017-12-01</p> <p>The multiple, spatially separated vantage points afforded by the STEREO and SOHO missions provide physicists with a means to infer the three-dimensional structure of the solar corona via tomographic <span class="hlt">imaging</span>. The reconstruction process combines these multiple projections of the optically thin plasma to constrain its three-dimensional density structure and has been successfully applied to the low corona using the STEREO and SOHO coronagraphs. However, the technique is also possible at larger, inter-planetary distances using wide-angle <span class="hlt">imagers</span>, such as the STEREO Heliospheric <span class="hlt">Imagers</span> (HIs), to <span class="hlt">observe</span> faint solar wind plasma and Coronal Mass Ejections (CMEs). Limited small-scale structure may be inferred from only three, or fewer, viewpoints and the work presented here is done so with the aim of establishing techniques for <span class="hlt">observing</span> CMEs with upcoming and future HI-like technology. We use simulated solar wind densities to compute realistic white-light HI <span class="hlt">observations</span>, with which we explore the requirements of such instruments for determining solar wind plasma density structure via tomography. We exploit this information to investigate the optimal orbital characteristics, such as spacecraft number, separation, inclination and eccentricity, necessary to perform the technique with HIs. Further to this we argue that tomography may be greatly enhanced by means of improved instrumentation; specifically, the use of wide-angle <span class="hlt">imagers</span> capable of measuring polarised light. This work has obvious space weather applications, serving as a demonstration for potential future missions (such as at L1 and L5) and will prove timely in fully exploiting the science return from the upcoming Solar Orbiter and Parker Solar Probe missions.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_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://hdl.handle.net/2060/19910016676','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19910016676"><span>HUT <span class="hlt">observations</span> of carbon monoxide in the coma of Comet Levy (1990c)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Feldman, P. D.; Davidsen, A. F.; Blair, W. P.; Bowers, C. W.; Dixon, W. V.; Durrance, S. T.; Henry, R. C.; Kriss, G. A.; Kruk, J.; Moos, H. W.</p> <p>1991-01-01</p> <p><span class="hlt">Observations</span> of comet Levy (1990c) were made with the Hopkins Ultraviolet Telescope during the Astro-1 Space Shuttle mission on 10 Dec. 1990. The spectrum, covering the wavelength range 415 to 1850 A at a spectral emission of 3 A (in first order), shows the presence of carbon monoxide and atomic hydrogen, carbon, and sulfur in the coma. Aside from H I Lyman-beta, no cometary features are detected below 1200 A, although cometary O I and O II would be masked by the same emissions present in the day <span class="hlt">airglow</span> spectrum. The 9.4 x 116 arcsec aperture corresponds to 12,000 x 148,000 km at the comet. The derived production rate of CO relative to water, 0.13 + or - 0.02, compared with the same ratio derived from IUE <span class="hlt">observations</span> (made in Sep. 1990) which sample a much smaller region of the coma, 0.04 + or - 0.01, suggests the presence of an extended source of CO, as was found in comet Halley. Upper limits on Ne and Ar abundance are within an order of magnitude or solar abundances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss028e029679.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss028e029679.html"><span>Earth <span class="hlt">observation</span> taken by the Expedition 28 crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-08-21</p> <p>ISS028-E-029679 (21 Aug. 2011) --- A night time view of India-Pakistan borderlands is featured in this <span class="hlt">image</span> photographed by an Expedition 28 crew member on the International Space Station. Clusters of yellow lights on the Indo-Gangetic Plain of northern India and northern Pakistan reveal numerous cities both large and small in this photograph. Of the hundreds of clusters, the largest are the metropolitan areas associated with the capital cities of Islamabad, Pakistan in the foreground and New Delhi, India at the top?for scale these metropolitan areas are approximately 700 kilometers apart. The lines of major highways connecting the larger cities also stand out. More subtle but still visible at night are the general outlines of the towering and partly cloud-covered Himalayan ranges immediately to the north (left). A striking feature of this photograph is the line of lights, with a distinctly more orange hue, snaking across the central part of the <span class="hlt">image</span>. It appears to be more continuous and brighter than most highways in the view. This is the fenced and floodlit border zone between the countries of India and Pakistan. The fence is designed to discourage smuggling and arms trafficking between the two countries. A similar fenced zone separates India?s eastern border from Bangladesh (not visible). This <span class="hlt">image</span> was taken with a 16-mm lens, which provides the wide field of view, as the space station was tracking towards the southeast across the subcontinent of India. The station crew took the <span class="hlt">image</span> as part of a continuous series of frames, each frame taken with a one-second exposure time to maximize light collection ? unfortunately, this also causes blurring of some ground features. The distinct, bright zone above the horizon (visible at top) is produced by <span class="hlt">airglow</span>, a phenomena caused by excitation of atoms and molecules high in the atmosphere (above 80 kilometers, or 50 miles altitude) by ultraviolet radiation from the sun. Part of the ISS Permanent Multipurpose Module</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140001025','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140001025"><span>Interactions between Coronal Mass Ejections Viewed in Coordinated <span class="hlt">Imaging</span> 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>Liu, Ying D.; Luhmann, Janet G.; Moestl, Christian; Martinez-Oliveros, Juan C.; Bale, Stewart D.; Lin, Robert P.; Harrison, Richard A.; Temmer, Manuela; Webb, David F.; Odstrcil, Dusan</p> <p>2013-01-01</p> <p>The successive coronal mass ejections (CMEs) from 2010 July 30 - August 1 present us the first opportunity to study CME-CME interactions with unprecedented heliospheric <span class="hlt">imaging</span> and in situ <span class="hlt">observations</span> from multiple vantage points. We describe two cases of CME interactions: merging of two CMEs launched close in time and overtaking of a preceding CME by a shock wave. The first two CMEs on August 1 interact close to the Sun and form a merged front, which then overtakes the July 30 CME near 1 AU, as revealed by wide-angle <span class="hlt">imaging</span> <span class="hlt">observations</span>. Connections between <span class="hlt">imaging</span> <span class="hlt">observations</span> and in situ signatures at 1 AU suggest that the merged front is a shock wave, followed by two ejecta <span class="hlt">observed</span> at Wind which seem to have already merged. In situ measurements show that the CME from July 30 is being overtaken by the shock at 1 AU and is significantly compressed, accelerated and heated. The interaction between the preceding ejecta and shock also results in variations in the shock strength and structure on a global scale, as shown by widely separated in situ measurements from Wind and STEREO B. These results indicate important implications of CME-CME interactions for shock propagation, particle acceleration and space weather forecasting.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860012558&hterms=barium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dbarium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860012558&hterms=barium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dbarium"><span><span class="hlt">Observations</span> of barium ion jets in the magnetosphere using Doppler <span class="hlt">imaging</span> systems and very sensitive <span class="hlt">imaging</span> systems using <span class="hlt">imaging</span> photon detectors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rees, D.; Conboy, J.; Heinz, W.; Heppner, J. P.</p> <p>1985-01-01</p> <p><span class="hlt">Observations</span> of four shaped charge releases from rockets launched from Alaska are described. Results demonstrate that <span class="hlt">imaging</span> and Doppler <span class="hlt">imaging</span> instruments, based on exploiting the <span class="hlt">imaging</span> photon detector, provide additional insight into the motion and development of low intensity targets such as the fast ion jets produced by shaped charge releases. It is possible to trace the motion of fast ion jets to very great distances, of the order of 50,000 km, outward along the Earth's magnetic field, when the conditions are suitable for the outward (upward) motion and/or acceleration of such ion jets. It is shown that ion jets, which fade below the lower sensitivity threshold of previous instruments, do not always disappear. There is no evidence of an abrupt field-aligned shear-type acceleration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23878358','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23878358"><span>Diffusion-weighted magnetic resonance <span class="hlt">imaging</span> combined with T2-weighted <span class="hlt">images</span> in the detection of small breast cancer: a single-center multi-<span class="hlt">observer</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>Wu, Lian-Ming; Chen, Jie; Hu, Jiani; Gu, Hai-Yan; Xu, Jian-Rong; Hua, Jia</p> <p>2014-02-01</p> <p>Breast cancer is the most common cancer in women worldwide. However, it remains a difficult diagnosis problem to differentiate between benign and malignant breast lesions, especially in small early breast lesions. To assess the diagnostic value of diffusion-weighted <span class="hlt">imaging</span> (DWI) combined with T2-weighted <span class="hlt">imaging</span> (T2WI) for small breast cancer characterization. Fifty-eight patients (65 lesions) with a lesion <2 cm in diameter underwent 3.0 Tesla breast magnetic resonance <span class="hlt">imaging</span> (MRI) including DWI and histological analysis. Three <span class="hlt">observers</span> with varying experience levels reviewed MRI. The probability of breast cancer in each lesion on MR <span class="hlt">images</span> was recorded with a 5-point scale. Areas under the receiver-operating characteristic curve (AUCs) were compared by using the Z test; sensitivity and specificity were determined with the Z test after adjusting for data clustering. AUC of T2WI and DWI (<span class="hlt">Observer</span> 1, 0.95; <span class="hlt">Observer</span> 2, 0.91; <span class="hlt">Observer</span> 3, 0.83) was greater than that of T2WI (<span class="hlt">Observer</span> 1, 0.80; <span class="hlt">Observer</span> 2, 0.74; <span class="hlt">Observer</span> 3, 0.70) for all <span class="hlt">observers</span> (P < 0.0001 in all comparisons). Sensitivity of T2WI and DWI (<span class="hlt">Observer</span> 1, 90%; <span class="hlt">Observer</span> 2, 93%; and <span class="hlt">Observer</span> 3, 86%) was greater than that of T2WI alone (<span class="hlt">Observer</span> 1, 76%; <span class="hlt">Observer</span> 2, 83%; <span class="hlt">Observer</span> 3, 79%) for all <span class="hlt">observers</span> (P < 0.0001 in all comparisons). Specificity of T2WI and DWI was greater than that of T2WI alone for <span class="hlt">observer</span> 1 (89% vs. 72%, P < 0.01) and <span class="hlt">observer</span> 2 (94% vs. 78%, P < 0.001). DWI combined with T2WI can improve the diagnostic performance of MRI in small breast cancer characterization. It should be considered selectively in the preoperative evaluation of patients with small lesions of the breast.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950029592&hterms=twilight&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950029592&hterms=twilight&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtwilight"><span>Simultaneous retrieval of the solar EUV flux and neutral thermospheric O, O2, N2, and temperature from twilight <span class="hlt">airglow</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fennelly, J. A.; Torr, D. G.; Richards, P. G.; Torr, M. R.</p> <p>1994-01-01</p> <p>We present a method to retrieve neutral thermospheric composition and the solar EUV flux from ground-based twilight optical measurements of the O(+) ((exp 2)P) 7320 A and O((exp 1)D) 6300 A <span class="hlt">airglow</span> emissions. The parameters retrieved are the neutral temperature, the O, O2, N2 density profiles, and a scaling factor for the solar EUV flux spectrum. The temperature, solar EUV flux scaling factor, and atomic oxygen density are first retrieved from the 7320-A emission, which are then used with the 6300-A emission to retrieve the O2 and N2 densities. The retrieval techniques have been verified by computer simulations. We have shown that the retrieval technique is able to statistically retrieve values, between 200 and 400 km, within an average error of 3.1 + or - 0.6% for thermospheric temperature, 3.3 + or - 2.0% for atomic oxygen, 2.3 + or - 1.3% for molecular oxygen, and 2.4 + or - 1.3% for molecular nitrogen. The solar EUV flux scaling factor was found to have a retrieval error of 5.1 + or - 2.3%. All the above errors have a confidence level of 95%. The purpose of this paper is to prove the viability and usefulness of the retrieval technique by demonstrating the ability to retrieve known quantities under a realistic simulation of the measurement process, excluding systematic effects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007SPIE.6515E..19L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007SPIE.6515E..19L"><span>Combining a wavelet transform with a channelized Hotelling <span class="hlt">observer</span> for tumor detection in 3D PET oncology <span class="hlt">imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lartizien, Carole; Tomei, Sandrine; Maxim, Voichita; Odet, Christophe</p> <p>2007-03-01</p> <p>This study evaluates new <span class="hlt">observer</span> models for 3D whole-body Positron Emission Tomography (PET) <span class="hlt">imaging</span> based on a wavelet sub-band decomposition and compares them with the classical constant-Q CHO model. Our final goal is to develop an original method that performs guided detection of abnormal activity foci in PET oncology <span class="hlt">imaging</span> based on these new <span class="hlt">observer</span> models. This computer-aided diagnostic method would highly benefit to clinicians for diagnostic purpose and to biologists for massive screening of rodents populations in molecular <span class="hlt">imaging</span>. Method: We have previously shown good correlation of the channelized Hotelling <span class="hlt">observer</span> (CHO) using a constant-Q model with human <span class="hlt">observer</span> performance for 3D PET oncology <span class="hlt">imaging</span>. We propose an alternate method based on combining a CHO <span class="hlt">observer</span> with a wavelet sub-band decomposition of the <span class="hlt">image</span> and we compare it to the standard CHO implementation. This method performs an undecimated transform using a biorthogonal B-spline 4/4 wavelet basis to extract the features set for input to the Hotelling <span class="hlt">observer</span>. This work is based on simulated 3D PET <span class="hlt">images</span> of an extended MCAT phantom with randomly located lesions. We compare three evaluation criteria: classification performance using the signal-to-noise ratio (SNR), computation efficiency and visual quality of the derived 3D maps of the decision variable λ. The SNR is estimated on a series of test <span class="hlt">images</span> for a variable number of training <span class="hlt">images</span> for both <span class="hlt">observers</span>. Results: Results show that the maximum SNR is higher with the constant-Q CHO <span class="hlt">observer</span>, especially for targets located in the liver, and that it is reached with a smaller number of training <span class="hlt">images</span>. However, preliminary analysis indicates that the visual quality of the 3D maps of the decision variable λ is higher with the wavelet-based CHO and the computation time to derive a 3D λ-map is about 350 times shorter than for the standard CHO. This suggests that the wavelet-CHO <span class="hlt">observer</span> is a good candidate for use in our guided</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.H11E1222F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.H11E1222F"><span>Riverine Bathymetry <span class="hlt">Imaging</span> with Indirect <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>Farthing, M.; Lee, J. H.; Ghorbanidehno, H.; Hesser, T.; Darve, E. F.; Kitanidis, P. K.</p> <p>2017-12-01</p> <p>Bathymetry, i.e, depth, <span class="hlt">imaging</span> in a river is of crucial importance for shipping operations and flood management. With advancements in sensor technology and computational resources, various types of indirect measurements can be used to estimate high-resolution riverbed topography. Especially, the use of surface velocity measurements has been actively investigated recently since they are easy to acquire at a low cost in all river conditions and surface velocities are sensitive to the river depth. In this work, we <span class="hlt">image</span> riverbed topography using depth-averaged quasi-steady velocity <span class="hlt">observations</span> related to the topography through the 2D shallow water equations (SWE). The principle component geostatistical approach (PCGA), a fast and scalable variational inverse modeling method powered by low-rank representation of covariance matrix structure, is presented and applied to two "twin" riverine bathymetry identification problems. To compare the efficiency and effectiveness of the proposed method, an ensemble-based approach is also applied to the test problems. Results demonstrate that PCGA is superior to the ensemble-based approach in terms of computational effort and accuracy. Especially, the results obtained from PCGA capture small-scale bathymetry features irrespective of the initial guess through the successive linearization of the forward model. Analysis on the direct survey data of the riverine bathymetry used in one of the test problems shows an efficient, parsimonious choice of the solution basis in PCGA so that the number of the numerical model runs used to achieve the inversion results is close to the minimum number that reconstructs the underlying bathymetry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...849...10F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...849...10F"><span>A Compressed Sensing-based <span class="hlt">Image</span> Reconstruction Algorithm for Solar Flare X-Ray <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>Felix, Simon; Bolzern, Roman; Battaglia, Marina</p> <p>2017-11-01</p> <p>One way of <span class="hlt">imaging</span> X-ray emission from solar flares is to measure Fourier components of the spatial X-ray source distribution. We present a new compressed sensing-based algorithm named VIS_CS, which reconstructs the spatial distribution from such Fourier components. We demonstrate the application of the algorithm on synthetic and <span class="hlt">observed</span> solar flare X-ray data from the Reuven Ramaty High Energy Solar Spectroscopic <span class="hlt">Imager</span> satellite and compare its performance with existing algorithms. VIS_CS produces competitive results with accurate photometry and morphology, without requiring any algorithm- and X-ray-source-specific parameter tuning. Its robustness and performance make this algorithm ideally suited for the generation of quicklook <span class="hlt">images</span> or large <span class="hlt">image</span> cubes without user intervention, such as for <span class="hlt">imaging</span> spectroscopy analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940031102','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940031102"><span>Obtaining coincident <span class="hlt">image</span> <span class="hlt">observations</span> for Mission to Planet Earth science data return</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Newman, Lauri Kraft; Folta, David C.; Farrell, James P.</p> <p>1994-01-01</p> <p>One objective of the Mission to Planet Earth (MTPE) program involves comparing data from various instruments on multiple spacecraft to obtain a total picture of the Earth's systems. To correlate <span class="hlt">image</span> data from instruments on different spacecraft, these spacecraft must be able to <span class="hlt">image</span> the same location on the Earth at approximately the same time. Depending on the orbits of the spacecraft involved, complicated operational details must be considered to obtain such <span class="hlt">observations</span>. If the spacecraft are in similar orbits, close formation flying or synchronization techniques may be used to assure coincident <span class="hlt">observations</span>. If the orbits are dissimilar, the launch time of the second satellite may need to be restricted in order to align its orbit with that of the first satellite launched. This paper examines strategies for obtaining coincident <span class="hlt">observations</span> for spacecraft in both similar and dissimilar orbits. Although these calculations may be performed easily for coplanar spacecraft, the non-coplanar case involves additional considerations which are incorporated into the algorithms presented herein.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22482368-computational-human-observer-image-quality-evaluation-low-dose-knowledge-based-ct-iterative-reconstruction','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22482368-computational-human-observer-image-quality-evaluation-low-dose-knowledge-based-ct-iterative-reconstruction"><span>Computational and human <span class="hlt">observer</span> <span class="hlt">image</span> quality evaluation of low dose, knowledge-based CT iterative reconstruction</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>Eck, Brendan L.; Fahmi, Rachid; Miao, Jun</p> <p>2015-10-15</p> <p>Purpose: Aims in this study are to (1) develop a computational model <span class="hlt">observer</span> which reliably tracks the detectability of human <span class="hlt">observers</span> in low dose computed tomography (CT) <span class="hlt">images</span> reconstructed with knowledge-based iterative reconstruction (IMR™, Philips Healthcare) and filtered back projection (FBP) across a range of independent variables, (2) use the model to evaluate detectability trends across reconstructions and make predictions of human <span class="hlt">observer</span> detectability, and (3) perform human <span class="hlt">observer</span> studies based on model predictions to demonstrate applications of the model in CT <span class="hlt">imaging</span>. Methods: Detectability (d′) was evaluated in phantom studies across a range of conditions. <span class="hlt">Images</span> were generated usingmore » a numerical CT simulator. Trained <span class="hlt">observers</span> performed 4-alternative forced choice (4-AFC) experiments across dose (1.3, 2.7, 4.0 mGy), pin size (4, 6, 8 mm), contrast (0.3%, 0.5%, 1.0%), and reconstruction (FBP, IMR), at fixed display window. A five-channel Laguerre–Gauss channelized Hotelling <span class="hlt">observer</span> (CHO) was developed with internal noise added to the decision variable and/or to channel outputs, creating six different internal noise models. Semianalytic internal noise computation was tested against Monte Carlo and used to accelerate internal noise parameter optimization. Model parameters were estimated from all experiments at once using maximum likelihood on the probability correct, P{sub C}. Akaike information criterion (AIC) was used to compare models of different orders. The best model was selected according to AIC and used to predict detectability in blended FBP-IMR <span class="hlt">images</span>, analyze trends in IMR detectability improvements, and predict dose savings with IMR. Predicted dose savings were compared against 4-AFC study results using physical CT phantom <span class="hlt">images</span>. Results: Detection in IMR was greater than FBP in all tested conditions. The CHO with internal noise proportional to channel output standard deviations, Model-k4, showed the best trade</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010020056&hterms=Mason&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D30%26Ntt%3DS%2BJ%2BMason','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010020056&hterms=Mason&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D30%26Ntt%3DS%2BJ%2BMason"><span>First Intrinsic Anisotropy <span class="hlt">Observations</span> With the Cosmic Background <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Padin, S.; Cartwright, J. K.; Mason, B. S.; Pearson, T. J.; Readhead, A. C. S.; Shepherd, M. C.; Sievers, J.; Udomprasert, P. S.; Holzapfel, W. L.; Myers, S. T.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20010020056'); toggleEditAbsImage('author_20010020056_show'); toggleEditAbsImage('author_20010020056_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20010020056_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20010020056_hide"></p> <p>2001-01-01</p> <p>We present the first results of <span class="hlt">observations</span> of the intrinsic anisotropy of the cosmic microwave background radiation with the Cosmic Background <span class="hlt">Imager</span> from a site at 5080 in altitude in northern Chile. Our <span class="hlt">observations</span> show a sharp decrease in C_l in the range l = 400 - 1500. Such a decrease in power at high l is one of the fundamental predictions of the standard cosmological model, and these are the first <span class="hlt">observations</span> which cover a broad enough 1-range to show this decrease in a single experiment. The power, C_l, at l approximately 600 is higher than measured by Boomerang and Maxima, with the differences being significant at the 2.7sigma and 1.9sigma levels, respectively. The C_l we have measured enable us to place limits on the density parameter, Omega(tot) <= 0.4 or Omega(tot) >= 0.7 (90% confidence).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19810025174','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810025174"><span>Large storms: <span class="hlt">Airglow</span> and related measurements. VLF <span class="hlt">observations</span>, volume 4</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>The data presented show the typical values and range of ionospheric and magnetospheric characteristics, as viewed from 1400 km with the ISIS 2 instruments. The definition of each data set depends partly on geophysical parameters and partly on satellite operating mode. Preceding the data set is a description of the organizational parameters and a review of the objectives and general characteristics of the data set. The data are shown as a selection from 12 different data formats. Each data set has a different selection of formats, but uniformity of a given format selection is preserved throughout each data set. Each data set consists of a selected number of passes, each comprising a format combination that is most appropriae for the particular data set. Description of ISIS 2 instruments are provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016amos.confE..26S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016amos.confE..26S"><span>Automated Astrometric Analysis of Satellite <span class="hlt">Observations</span> using Wide-field <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Skuljan, J.; Kay, J.</p> <p>2016-09-01</p> <p>An <span class="hlt">observational</span> trial was conducted in the South Island of New Zealand from 24 to 28 February 2015, as a collaborative effort between the United Kingdom and New Zealand in the area of space situational awareness. The aim of the trial was to <span class="hlt">observe</span> a number of satellites in low Earth orbit using wide-field <span class="hlt">imaging</span> from two separate locations, in order to determine the space trajectory and compare the measurements with the predictions based on the standard two-line elements. This activity was an initial step in building a space situational awareness capability at the Defence Technology Agency of the New Zealand Defence Force. New Zealand has an important strategic position as the last land mass that many satellites selected for deorbiting pass before entering the Earth's atmosphere over the dedicated disposal area in the South Pacific. A preliminary analysis of the trial data has demonstrated that relatively inexpensive equipment can be used to successfully detect satellites at moderate altitudes. A total of 60 satellite passes were <span class="hlt">observed</span> over the five nights of <span class="hlt">observation</span> and about 2600 <span class="hlt">images</span> were collected. A combination of cooled CCD and standard DSLR cameras were used, with a selection of lenses between 17 mm and 50 mm in focal length, covering a relatively wide field of view of 25 to 60 degrees. The CCD cameras were equipped with custom-made GPS modules to record the time of exposure with a high accuracy of one millisecond, or better. Specialised software has been developed for automated astrometric analysis of the trial data. The astrometric solution is obtained as a two-dimensional least-squares polynomial fit to the measured pixel positions of a large number of stars (typically 1000) detected across the <span class="hlt">image</span>. The star identification is fully automated and works well for all camera-lens combinations used in the trial. A moderate polynomial degree of 3 to 5 is selected to take into account any <span class="hlt">image</span> distortions introduced by the lens. A typical RMS</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MNRAS.473.1038P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MNRAS.473.1038P"><span>Robust sparse <span class="hlt">image</span> reconstruction of radio interferometric <span class="hlt">observations</span> with PURIFY</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pratley, Luke; McEwen, Jason D.; d'Avezac, Mayeul; Carrillo, Rafael E.; Onose, Alexandru; Wiaux, Yves</p> <p>2018-01-01</p> <p>Next-generation radio interferometers, such as the Square Kilometre Array, will revolutionize our understanding of the Universe through their unprecedented sensitivity and resolution. However, to realize these goals significant challenges in <span class="hlt">image</span> and data processing need to be overcome. The standard methods in radio interferometry for reconstructing <span class="hlt">images</span>, such as CLEAN, have served the community well over the last few decades and have survived largely because they are pragmatic. However, they produce reconstructed interferometric <span class="hlt">images</span> that are limited in quality and scalability for big data. In this work, we apply and evaluate alternative interferometric reconstruction methods that make use of state-of-the-art sparse <span class="hlt">image</span> reconstruction algorithms motivated by compressive sensing, which have been implemented in the PURIFY software package. In particular, we implement and apply the proximal alternating direction method of multipliers algorithm presented in a recent article. First, we assess the impact of the interpolation kernel used to perform gridding and degridding on sparse <span class="hlt">image</span> reconstruction. We find that the Kaiser-Bessel interpolation kernel performs as well as prolate spheroidal wave functions while providing a computational saving and an analytic form. Secondly, we apply PURIFY to real interferometric <span class="hlt">observations</span> from the Very Large Array and the Australia Telescope Compact Array and find that <span class="hlt">images</span> recovered by PURIFY are of higher quality than those recovered by CLEAN. Thirdly, we discuss how PURIFY reconstructions exhibit additional advantages over those recovered by CLEAN. The latest version of PURIFY, with developments presented in this work, is made publicly available.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013DPS....4521103S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013DPS....4521103S"><span>Deriving Atmospheric Properties and Escape Rates from MAVEN's <span class="hlt">Imaging</span> UV Spectrograph (IUVS)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schneider, Nicholas M.; IUVS Science Team</p> <p>2013-10-01</p> <p>MAVEN (Mars Volatile and Atmosphere EvolutioN) is a Mars Scout mission being readied for launch in November 2013. The key hardware and management partners are University of Colorado, Goddard Space Flight Center, University of California at Berkeley, Lockheed Martin, and the Jet Propulsion Laboratory. MAVEN carries a powerful suite of fields and particles instruments and a sophisticated remote sensing instrument, the <span class="hlt">Imaging</span> UltraViolet Spectrograph (IUVS). This presentation begins by describing IUVS' science goals, instrument design, operational approach and data analysis strategy. IUVS supports the top-level MAVEN science goals: measure the present state of the atmosphere, <span class="hlt">observe</span> its response to varying solar stimuli, and use the information to estimate loss from Mars' atmosphere over time. The instrument operates at low spectral resolution spanning the FUV and MUV ranges in separate channels, and at high resolution around the hydrogen Lyman alpha line to measure the D/H ratio in the upper atmosphere. MAVEN carries the instrument on an Articulated Payload Platform which orients the instrument for optimal <span class="hlt">observations</span> during four segments of its 4.5 hr elliptical orbit. During periapse passage, IUVS uses a scan mirror to obtain vertical profiles of emissions from the atmosphere and ionosphere. Around apoapse, the instrument builds up low-resolution <span class="hlt">images</span> of the atmosphere at multiple wavelengths. In between, the instrument measures emissions from oxygen, hydrogen and deuterium in the corona. IUVS also undertakes day-long stellar occultation campaigns at 2 month intervals, to measure the state of the atmosphere at altitudes below the <span class="hlt">airglow</span> layer and in situ sampling. All data will be pipeline-processed from line brightnesses to column abundances, local densities and global 3-D maps. The focus of the presentation is development of these automatic processing algorithms and the data products they will provide to the Mars community through the PDS Atmospheres Node</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss016e015496.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss016e015496.html"><span>View of Solar Array Panels taken during Expedition 16</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2007-12-09</p> <p>ISS016-E-015496 (9 Dec. 2007) --- Solar array panels of the International Space Station are featured in this <span class="hlt">image</span> photographed by an Expedition 16 crewmember (out of frame) from a window on the station. The blackness of space and <span class="hlt">airglow</span> of Earth's horizon provide the backdrop for the scene.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20020049425&hterms=Galileo&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DGalileo','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20020049425&hterms=Galileo&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DGalileo"><span>Monte Carlo Radiative Transfer Modeling of Lightning <span class="hlt">Observed</span> in Galileo <span class="hlt">Images</span> of Jupiter</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dyudine, U. A.; Ingersoll, Andrew P.</p> <p>2002-01-01</p> <p>We study lightning on Jupiter and the clouds illuminated by the lightning using <span class="hlt">images</span> taken by the Galileo orbiter. The Galileo <span class="hlt">images</span> have a resolution of 25 km/pixel and axe able to resolve the shape of the single lightning spots in the <span class="hlt">images</span>, which have full widths at half the maximum intensity in the range of 90-160 km. We compare the measured lightning flash <span class="hlt">images</span> with simulated <span class="hlt">images</span> produced by our ED Monte Carlo light-scattering model. The model calculates Monte Carlo scattering of photons in a ED opacity distribution. During each scattering event, light is partially absorbed. The new direction of the photon after scattering is chosen according to a Henyey-Greenstein phase function. An <span class="hlt">image</span> from each direction is produced by accumulating photons emerging from the cloud in a small range (bins) of emission angles. Lightning bolts are modeled either as points or vertical lines. Our results suggest that some of the <span class="hlt">observed</span> scattering patterns axe produced in a 3-D cloud rather than in a plane-parallel cloud layer. Lightning is estimated to occur at least as deep as the bottom of the expected water cloud. For the six cases studied, we find that the clouds above the lightning are optically thick (tau > 5). Jovian flashes are more regular and circular than the largest terrestrial flashes <span class="hlt">observed</span> from space. On Jupiter there is nothing equivalent to the 30-40-km horizontal flashes which axe seen on Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015DPS....4710508R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015DPS....4710508R"><span>Pluto's Extended Atmosphere: New Horizons Alice Lyman-α <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Retherford, Kurt D.; Gladstone, G. Randall; Stern, S. Alan; Weaver, Harold A.; Young, Leslie A.; Ennico, Kimberly A.; Olkin, Cathy B.; Cheng, Andy F.; Greathouse, Thomas K.; Hinson, David P.; Kammer, Joshua A.; Linscott, Ivan R.; Parker, Alex H.; Parker, Joel Wm.; Pryor, Wayne R.; Schindhelm, Eric; Singer, Kelsi N.; Steffl, Andrew J.; Strobel, Darrell F.; Summers, Michael E.; Tsang, Constantine C. C.; Tyler, G. Len; Versteeg, Maarten H.; Woods, William W.; Cunningham, Nathaniel J.; Curdt, Werner</p> <p>2015-11-01</p> <p>Pluto's upper atmosphere is expected to extend several planetary radii, proportionally more so than for any planet in our solar system. Atomic hydrogen is readily produced at lower altitudes due to photolysis of methane and transported upward to become an important constituent. The Interplanetary Medium (IPM) provides a natural light source with which to study Pluto's atomic hydrogen atmosphere. While direct solar Lyman-α emissions dominate the signal at 121.6 nm at classical solar system distances, the contribution of diffuse illumination by IPM Lyman-α sky-glow is roughly on par at Pluto (Gladstone et al., Icarus, 2015). Hydrogen atoms in Pluto's upper atmosphere scatter these bright Lyα emission lines, and detailed simulations of the radiative transfer for these photons indicate that Pluto would appear dark against the IPM Lyα background. The Pluto-Alice UV <span class="hlt">imaging</span> spectrograph on New Horizons conducted several <span class="hlt">observations</span> of Pluto during the encounter to search for <span class="hlt">airglow</span> emissions, characterize its UV reflectance spectra, and to measure the radial distribution of IPM Lyα near the disk. Our early results suggest that these model predictions for the darkening of IPM Lyα with decreasing altitude being measureable by Pluto-Alice were correct. We'll report our progress toward extracting H and CH4 density profiles in Pluto's upper atmosphere through comparisons of these data with detailed radiative transfer modeling. These New Horizons findings will have important implications for determining the extent of Pluto's atmosphere and related constraints to high-altitude vertical temperature structure and atmospheric escape.This work was supported by NASA's New Horizons project.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20076181','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20076181"><span>NASA Program of Airborne Optical <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>Bader, M; Wagoner, C B</p> <p>1970-02-01</p> <p>NASA's Ames Research Center currently operates a Convair 990 four-engine jet transport as a National Facility for airborne scientific research (astronomy, aurora, <span class="hlt">airglow</span>, meteorology, earth resources). This aircraft can carry about twelve experiments to 12 km for several hours. A second aircraft, a twin-engine Lear Jet, has been used on a limited basis for airborne science and can carry one experiment to 15 km for 1 h. Mobility and altitude are the principal advantages over ground sites, while large payload and personnel carrying capabilities, combined with ease of operations and relatively low cost, are the main advantages compared to balloons, rockets, or satellites. Typical airborne instrumentation and scientific results are presented.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPhCS.969a2153M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPhCS.969a2153M"><span><span class="hlt">Observation</span> of ferroelastic domains in layered magnetic compounds using birefringence <span class="hlt">imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Miura, Yoko; Okumura, Kazuya; Manaka, Hirotaka</p> <p>2018-03-01</p> <p>The two-dimensional Heisenberg antiferromagnet (C2H5NH3)2CuCl4 is a candidate compound for the coexistence of ferroelectricity and ferroelasticity; however, the microscopic <span class="hlt">observations</span> of multiferroic domains may still be unclear. In-plane birefringence <span class="hlt">imaging</span> measurements were performed to <span class="hlt">observe</span> the manner in which the ferroelectric and the ferroelastic domains change during phase transitions between 15 K and 300 K. It was found that 90° ferroelastic domains appeared in the ab-plane at 300 K. As the temperature decreased toward 15 K, each domain inverted at a certain temperature (T a) without structural or magnetic phase transitions. The value of T a was found to be significantly influenced by external stresses; therefore, birefringence <span class="hlt">imaging</span> techniques are useful for investigating variations in ferroelastic domains with temperature. Furthermore, a structural phase transition from orthorhombic to monoclinic or triclinic occurred at 230 ~ 240 K; however, no spontaneous polarization appeared in the ab-plane over the entire investigated range.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22679727-compressed-sensing-based-image-reconstruction-algorithm-solar-flare-ray-observations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22679727-compressed-sensing-based-image-reconstruction-algorithm-solar-flare-ray-observations"><span>A Compressed Sensing-based <span class="hlt">Image</span> Reconstruction Algorithm for Solar Flare X-Ray <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>Felix, Simon; Bolzern, Roman; Battaglia, Marina, E-mail: simon.felix@fhnw.ch, E-mail: roman.bolzern@fhnw.ch, E-mail: marina.battaglia@fhnw.ch</p> <p></p> <p>One way of <span class="hlt">imaging</span> X-ray emission from solar flares is to measure Fourier components of the spatial X-ray source distribution. We present a new compressed sensing-based algorithm named VIS-CS, which reconstructs the spatial distribution from such Fourier components. We demonstrate the application of the algorithm on synthetic and <span class="hlt">observed</span> solar flare X-ray data from the Reuven Ramaty High Energy Solar Spectroscopic <span class="hlt">Imager</span> satellite and compare its performance with existing algorithms. VIS-CS produces competitive results with accurate photometry and morphology, without requiring any algorithm- and X-ray-source-specific parameter tuning. Its robustness and performance make this algorithm ideally suited for the generation ofmore » quicklook <span class="hlt">images</span> or large <span class="hlt">image</span> cubes without user intervention, such as for <span class="hlt">imaging</span> spectroscopy analysis.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140002252','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140002252"><span><span class="hlt">Images</span> of Bottomside Irregularities <span class="hlt">Observed</span> at Topside Altitudes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Burke, William J.; Gentile, Louise C.; Shomo, Shannon R.; Roddy, Patrick A.; Pfaff, Robert F.</p> <p>2012-01-01</p> <p>We analyzed plasma and field measurements acquired by the Communication/ Navigation Outage Forecasting System (C/NOFS) satellite during an eight-hour period on 13-14 January 2010 when strong to moderate 250 MHz scintillation activity was <span class="hlt">observed</span> at nearby Scintillation Network Decision Aid (SCINDA) ground stations. C/NOFS consistently detected relatively small-scale density and electric field irregularities embedded within large-scale (approx 100 km) structures at topside altitudes. Significant spectral power measured at the Fresnel (approx 1 km) scale size suggests that C/NOFS was magnetically conjugate to bottomside irregularities similar to those directly responsible for the <span class="hlt">observed</span> scintillations. Simultaneous ion drift and plasma density measurements indicate three distinct types of large-scale irregularities: (1) upward moving depletions, (2) downward moving depletions, and (3) upward moving density enhancements. The first type has the characteristics of equatorial plasma bubbles; the second and third do not. The data suggest that both downward moving depletions and upward moving density enhancements and the embedded small-scale irregularities may be regarded as Alfvenic <span class="hlt">images</span> of bottomside irregularities. This interpretation is consistent with predictions of previously reported theoretical modeling and with satellite <span class="hlt">observations</span> of upward-directed Poynting flux in the low-latitude ionosphere.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-STS099-349-002.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-STS099-349-002.html"><span>Views of a sunrise and an aurora taken from OV-105 during STS-99</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2000-04-06</p> <p>STS099-349-002 (11-22 February 2000) ---The Space Shuttle Endeavour's vertical stabilizer is visible in the foreground of this 35mm frame featuring <span class="hlt">airglow</span>, the thin greenish band above the horizon. <span class="hlt">Airglow</span> is radiation emitted by the atmosphere from a layer about 30 kilometers thick and about 100 kilometers altitude. The predominant emission in <span class="hlt">airglow</span> is the green 5577-Angstrom wavelength emission from atomic oxygen atoms. <span class="hlt">Airglow</span> is always and everywhere present in the atmosphere; it results from the recombination of molecules that have been broken apart by solar radiation during the day. But <span class="hlt">airglow</span> is so faint that it can only be seen at night by looking "edge on" at the emission layer, such as the view astronauts have in orbit.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JaJAP..53dEL02T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JaJAP..53dEL02T"><span>A CMOS <span class="hlt">image</span> sensor with stacked photodiodes for lensless <span class="hlt">observation</span> system of digital enzyme-linked immunosorbent assay</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takehara, Hironari; Miyazawa, Kazuya; Noda, Toshihiko; Sasagawa, Kiyotaka; Tokuda, Takashi; Kim, Soo Hyeon; Iino, Ryota; Noji, Hiroyuki; Ohta, Jun</p> <p>2014-01-01</p> <p>A CMOS <span class="hlt">image</span> sensor with stacked photodiodes was fabricated using 0.18 µm mixed signal CMOS process technology. Two photodiodes were stacked at the same position of each pixel of the CMOS <span class="hlt">image</span> sensor. The stacked photodiodes consist of shallow high-concentration N-type layer (N+), P-type well (PW), deep N-type well (DNW), and P-type substrate (P-sub). PW and P-sub were shorted to ground. By monitoring the voltage of N+ and DNW individually, we can <span class="hlt">observe</span> two monochromatic colors simultaneously without using any color filters. The CMOS <span class="hlt">image</span> sensor is suitable for fluorescence <span class="hlt">imaging</span>, especially contact <span class="hlt">imaging</span> such as a lensless <span class="hlt">observation</span> system of digital enzyme-linked immunosorbent assay (ELISA). Since the fluorescence increases with time in digital ELISA, it is possible to <span class="hlt">observe</span> fluorescence accurately by calculating the difference from the initial relation between the pixel values for both photodiodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22634775-su-computer-simulation-model-observer-task-based-image-quality-assessment-radiation-therapy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22634775-su-computer-simulation-model-observer-task-based-image-quality-assessment-radiation-therapy"><span>SU-F-J-178: A Computer Simulation Model <span class="hlt">Observer</span> for Task-Based <span class="hlt">Image</span> Quality Assessment in Radiation Therapy</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>Dolly, S; Mutic, S; Anastasio, M</p> <p></p> <p>Purpose: Traditionally, <span class="hlt">image</span> quality in radiation therapy is assessed subjectively or by utilizing physically-based metrics. Some model <span class="hlt">observers</span> exist for task-based medical <span class="hlt">image</span> quality assessment, but almost exclusively for diagnostic <span class="hlt">imaging</span> tasks. As opposed to disease diagnosis, the task for <span class="hlt">image</span> <span class="hlt">observers</span> in radiation therapy is to utilize the available <span class="hlt">images</span> to design and deliver a radiation dose which maximizes patient disease control while minimizing normal tissue damage. The purpose of this study was to design and implement a new computer simulation model <span class="hlt">observer</span> to enable task-based <span class="hlt">image</span> quality assessment in radiation therapy. Methods: A modular computer simulation framework wasmore » developed to resemble the radiotherapy <span class="hlt">observer</span> by simulating an end-to-end radiation therapy treatment. Given <span class="hlt">images</span> and the ground-truth organ boundaries from a numerical phantom as inputs, the framework simulates an external beam radiation therapy treatment and quantifies patient treatment outcomes using the previously defined therapeutic operating characteristic (TOC) curve. As a preliminary demonstration, TOC curves were calculated for various CT acquisition and reconstruction parameters, with the goal of assessing and optimizing simulation CT <span class="hlt">image</span> quality for radiation therapy. Sources of randomness and bias within the system were analyzed. Results: The relationship between CT <span class="hlt">imaging</span> dose and patient treatment outcome was objectively quantified in terms of a singular value, the area under the TOC (AUTOC) curve. The AUTOC decreases more rapidly for low-dose <span class="hlt">imaging</span> protocols. AUTOC variation introduced by the dose optimization algorithm was approximately 0.02%, at the 95% confidence interval. Conclusion: A model <span class="hlt">observer</span> has been developed and implemented to assess <span class="hlt">image</span> quality based on radiation therapy treatment efficacy. It enables objective determination of appropriate <span class="hlt">imaging</span> parameter values (e.g. <span class="hlt">imaging</span> dose). Framework flexibility allows for</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020083315','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020083315"><span>Earth <span class="hlt">Observation</span> Services (<span class="hlt">Image</span> Processing Software)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1992-01-01</p> <p>San Diego State University and Environmental Systems Research Institute, with other agencies, have applied satellite <span class="hlt">imaging</span> and <span class="hlt">image</span> processing techniques to geographic information systems (GIS) updating. The resulting <span class="hlt">images</span> display land use and are used by a regional planning agency for applications like mapping vegetation distribution and preserving wildlife habitats. The EOCAP program provides government co-funding to encourage private investment in, and to broaden the use of NASA-developed technology for analyzing information about Earth and ocean resources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA198889','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA198889"><span>Dynamics of the Polar Mesopause and Lower Thermosphere Region as <span class="hlt">Observed</span> in the Night <span class="hlt">Airglow</span> Emissions</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1988-02-01</p> <p>feasible. However, only the winter 80-110 km region could be studied by this approach, due to the limitation in the <span class="hlt">observing</span> method , i.e. nightglow...epoch methods have been employed. The diurnal tide component at the mesopause, that according to the latest tidal models should dominate in polar region...temperature was calculated by Kvifte’s method . tweem temperature and intensity was found by Shaqae [1974] using the intensity ratio of the P,(2) and P,(3</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16134866','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16134866"><span>Dependence of reconstructed <span class="hlt">image</span> characteristics on the <span class="hlt">observation</span> condition in light-in-flight recording by holography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Komatsu, Aya; Awatsuji, Yasuhiro; Kubota, Toshihiro</p> <p>2005-08-01</p> <p>We analyze the dependence of the reconstructed <span class="hlt">image</span> characteristic on the <span class="hlt">observation</span> condition in the light-in-flight recording by holography both theoretically and experimentally. This holography makes it possible to record a propagating light pulse. We have found that the shape of the reconstructed <span class="hlt">image</span> is changed when the <span class="hlt">observation</span> position is vertically moved along the hologram plane. The reconstructed <span class="hlt">image</span> is numerically simulated on the basis of the theory and is experimentally obtained by using a 373 fs pulsed laser. The numerical results agree with the experimental result, and the validity of the theory is verified. Also, experimental results are analyzed and the restoration of the reconstructed <span class="hlt">image</span> is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015A%26A...574A.120J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015A%26A...574A.120J"><span>High-contrast <span class="hlt">imaging</span> with Spitzer: deep <span class="hlt">observations</span> of Vega, Fomalhaut, and ɛ Eridani</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Janson, Markus; Quanz, Sascha P.; Carson, Joseph C.; Thalmann, Christian; Lafrenière, David; Amara, Adam</p> <p>2015-02-01</p> <p>Stars with debris disks are intriguing targets for direct-<span class="hlt">imaging</span> exoplanet searches, owing both to previous detections of wide planets in debris disk systems, and to commonly existing morphological features in the disks themselves that may be indicative of a planetary influence. Here we present <span class="hlt">observations</span> of three of the most nearby young stars, which are also known to host massive debris disks: Vega, Fomalhaut, and ɛ Eri. The Spitzer Space Telescope is used at a range of orientation angles for each star to supply a deep contrast through angular differential <span class="hlt">imaging</span> combined with high-contrast algorithms. The <span class="hlt">observations</span> provide the opportunity to probe substantially colder bound planets (120-330 K) than is possible with any other technique or instrument. For Vega, some apparently very red candidate point sources detected in the 4.5 μm <span class="hlt">image</span> remain to be tested for common proper motion. The <span class="hlt">images</span> are sensitive to ~2 Mjup companions at 150 AU in this system. The <span class="hlt">observations</span> presented here represent the first search for planets around Vega using Spitzer. The upper 4.5 μm flux limit on Fomalhaut b could be further constrained relative to previous data. In the case of ɛ Eri, planets below both the effective temperature and the mass of Jupiter could be probed from 80 AU and outward, although no such planets were found. The data sensitively probe the regions around the edges of the debris rings in the systems where planets can be expected to reside. These <span class="hlt">observations</span> validate previous results showing that more than an order of magnitude improvement in performance in the contrast-limited regime can be acquired with respect to conventional methods by applying sophisticated high-contrast techniques to space-based telescopes, thanks to the high degree of PSF stability provided in this environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28134567','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28134567"><span>Blurred digital mammography <span class="hlt">images</span>: an analysis of technical recall and <span class="hlt">observer</span> detection performance.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ma, Wang Kei; Borgen, Rita; Kelly, Judith; Millington, Sara; Hilton, Beverley; Aspin, Rob; Lança, Carla; Hogg, Peter</p> <p>2017-03-01</p> <p>Blurred <span class="hlt">images</span> in full-field digital mammography are a problem in the UK Breast Screening Programme. Technical recalls may be due to blurring not being seen on lower resolution monitors used for review. This study assesses the visual detection of blurring on a 2.3-MP monitor and a 5-MP report grade monitor and proposes an <span class="hlt">observer</span> standard for the visual detection of blurring on a 5-MP reporting grade monitor. 28 <span class="hlt">observers</span> assessed 120 <span class="hlt">images</span> for blurring; 20 <span class="hlt">images</span> had no blurring present, whereas 100 <span class="hlt">images</span> had blurring imposed through mathematical simulation at 0.2, 0.4, 0.6, 0.8 and 1.0 mm levels of motion. Technical recall rate for both monitors and angular size at each level of motion were calculated. χ 2 tests were used to test whether significant differences in blurring detection existed between 2.3- and 5-MP monitors. The technical recall rate for 2.3- and 5-MP monitors are 20.3% and 9.1%, respectively. The angular size for 0.2- to 1-mm motion varied from 55 to 275 arc s. The minimum amount of motion for visual detection of blurring in this study is 0.4 mm. For 0.2-mm simulated motion, there was no significant difference [χ 2 (1, N = 1095) = 1.61, p = 0.20] in blurring detection between the 2.3- and 5-MP monitors. According to this study, monitors ≤2.3 MP are not suitable for technical review of full-field digital mammography <span class="hlt">images</span> for the detection of blur. Advances in knowledge: This research proposes the first <span class="hlt">observer</span> standard for the visual detection of blurring.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5601529','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5601529"><span>Blurred digital mammography <span class="hlt">images</span>: an analysis of technical recall and <span class="hlt">observer</span> detection performance</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Borgen, Rita; Kelly, Judith; Millington, Sara; Hilton, Beverley; Aspin, Rob; Lança, Carla; Hogg, Peter</p> <p>2017-01-01</p> <p>Objective: Blurred <span class="hlt">images</span> in full-field digital mammography are a problem in the UK Breast Screening Programme. Technical recalls may be due to blurring not being seen on lower resolution monitors used for review. This study assesses the visual detection of blurring on a 2.3-MP monitor and a 5-MP report grade monitor and proposes an <span class="hlt">observer</span> standard for the visual detection of blurring on a 5-MP reporting grade monitor. Methods: 28 <span class="hlt">observers</span> assessed 120 <span class="hlt">images</span> for blurring; 20 <span class="hlt">images</span> had no blurring present, whereas 100 <span class="hlt">images</span> had blurring imposed through mathematical simulation at 0.2, 0.4, 0.6, 0.8 and 1.0 mm levels of motion. Technical recall rate for both monitors and angular size at each level of motion were calculated. χ2 tests were used to test whether significant differences in blurring detection existed between 2.3- and 5-MP monitors. Results: The technical recall rate for 2.3- and 5-MP monitors are 20.3% and 9.1%, respectively. The angular size for 0.2- to 1-mm motion varied from 55 to 275 arc s. The minimum amount of motion for visual detection of blurring in this study is 0.4 mm. For 0.2-mm simulated motion, there was no significant difference [χ2 (1, N = 1095) = 1.61, p = 0.20] in blurring detection between the 2.3- and 5-MP monitors. Conclusion: According to this study, monitors ≤2.3 MP are not suitable for technical review of full-field digital mammography <span class="hlt">images</span> for the detection of blur. Advances in knowledge: This research proposes the first <span class="hlt">observer</span> standard for the visual detection of blurring. PMID:28134567</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980038128','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980038128"><span>Maynooth Optical Aeronomical Facility</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mulligan, Francis J.; Niciejewski, Rick J.</p> <p>1994-01-01</p> <p>Ground-based measurements of upper atmospheric parameters, such as temperature and wind velocity, can be made by <span class="hlt">observing</span> <span class="hlt">airglow</span> emissions that have a well-defined altitude profile and that are known to be representative of the emitting region. We describe the optical observatory at Maynooth (53.23 deg N, 6.4 deg W) at which two instruments, a Fabry-Perot interferometer and a Fourier transform spectrometer, are used to record atmospheric <span class="hlt">airglow</span> emissions in Ireland at visible and near-infrared wavelengths, respectively. Descriptions of the instruments, data acquisition, and analysis procedures are provided, together with some sample results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12016864','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12016864"><span>[Scanning electron microscope <span class="hlt">observation</span> and <span class="hlt">image</span> quantitative analysis of Hippocampi].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Z; Pu, Z; Xu, L; Xu, G; Wang, Q; Xu, G; Wu, L; Chen, J</p> <p>1998-12-01</p> <p>The "scale-like projects" on the derma of 3 species of Hippocampi, H. kuda Bleerer, H. trimaculatus Leach and H. japonicus Kaup were <span class="hlt">observed</span> by scanning electron microscope (SEM). Results showed that some characteristics such us size, shape and type of arrangement of the "scale-like projects" can be considered as the evidence for microanalysis. <span class="hlt">Image</span> quantitative analysis of the "scale-like project" was carried out on 45 pieces of photograph using area, long diameter, short diameter and shape factor as parameters. No difference among the different parts of the same species was <span class="hlt">observed</span>, but significant differences were found among the above 3 species.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020083270','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020083270"><span>Earth <span class="hlt">Observation</span> Services Weather <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1992-01-01</p> <p>Microprocessor-based systems for processing satellite data offer mariners real-time <span class="hlt">images</span> of weather systems, day and night, of large areas or allow them to zoom in on a few square miles. Systems West markets these commercial <span class="hlt">image</span> processing systems, which have significantly decreased the cost of satellite weather stations. The company was assisted by the EOCAP program, which provides government co-funding to encourage private investment in, and to broaden the use of, NASA-developed technology for analyzing information about Earth and ocean resources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19770032925&hterms=Tidal+waves&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DTidal%2Bwaves','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19770032925&hterms=Tidal+waves&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DTidal%2Bwaves"><span>Internal wave <span class="hlt">observations</span> made with an airborne synthetic aperture <span class="hlt">imaging</span> radar</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Elachi, C.; Apel, J. R.</p> <p>1976-01-01</p> <p>Synthetic aperture L-band radar flown aboard the NASA CV-990 has <span class="hlt">observed</span> periodic striations on the ocean surface off the coast of Alaska which have been interpreted as tidally excited oceanic internal waves of less than 500 m length. These radar <span class="hlt">images</span> are compared to photographic imagery of similar waves taken from Landsat 1. Both the radar and Landsat <span class="hlt">images</span> reveal variations in reflectivity across each wave in a packet that range from low to high to normal. The variations point to the simultaneous existence of two mechanisms for the surface signatures of internal waves: roughening due to wave-current interactions, and smoothing due to slick formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8318E..0RO','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8318E..0RO"><span>Creation of an ensemble of simulated cardiac cases and a human <span class="hlt">observer</span> study: tools for the development of numerical <span class="hlt">observers</span> for SPECT myocardial perfusion <span class="hlt">imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>O'Connor, J. Michael; Pretorius, P. Hendrik; Gifford, Howard C.; Licho, Robert; Joffe, Samuel; McGuiness, Matthew; Mehurg, Shannon; Zacharias, Michael; Brankov, Jovan G.</p> <p>2012-02-01</p> <p>Our previous Single Photon Emission Computed Tomography (SPECT) myocardial perfusion <span class="hlt">imaging</span> (MPI) research explored the utility of numerical <span class="hlt">observers</span>. We recently created two hundred and eighty simulated SPECT cardiac cases using Dynamic MCAT (DMCAT) and SIMIND Monte Carlo tools. All simulated cases were then processed with two reconstruction methods: iterative ordered subset expectation maximization (OSEM) and filtered back-projection (FBP). <span class="hlt">Observer</span> study sets were assembled for both OSEM and FBP methods. Five physicians performed an <span class="hlt">observer</span> study on one hundred and seventy-nine <span class="hlt">images</span> from the simulated cases. The <span class="hlt">observer</span> task was to indicate detection of any myocardial perfusion defect using the American Society of Nuclear Cardiology (ASNC) 17-segment cardiac model and the ASNC five-scale rating guidelines. Human <span class="hlt">observer</span> Receiver Operating Characteristic (ROC) studies established the guidelines for the subsequent evaluation of numerical model <span class="hlt">observer</span> (NO) performance. Several NOs were formulated and their performance was compared with the human <span class="hlt">observer</span> performance. One type of NO was based on evaluation of a cardiac polar map that had been pre-processed using a gradient-magnitude watershed segmentation algorithm. The second type of NO was also based on analysis of a cardiac polar map but with use of a priori calculated average <span class="hlt">image</span> derived from an ensemble of normal cases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27828274','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27828274"><span>Interference data correction methods for lunar <span class="hlt">observation</span> with a large-aperture static <span class="hlt">imaging</span> spectrometer.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Geng; Wang, Shuang; Li, Libo; Hu, Xiuqing; Hu, Bingliang</p> <p>2016-11-01</p> <p>The lunar spectrum has been used in radiometric calibration and sensor stability monitoring for spaceborne optical sensors. A ground-based large-aperture static <span class="hlt">image</span> spectrometer (LASIS) can be used to acquire the lunar spectral <span class="hlt">image</span> for lunar radiance model improvement when the moon orbits over its viewing field. The lunar orbiting behavior is not consistent with the desired scanning speed and direction of LASIS. To correctly extract interferograms from the obtained data, a translation correction method based on <span class="hlt">image</span> correlation is proposed. This method registers the frames to a reference frame to reduce accumulative errors. Furthermore, we propose a circle-matching-based approach to achieve even higher accuracy during <span class="hlt">observation</span> of the full moon. To demonstrate the effectiveness of our approaches, experiments are run on true lunar <span class="hlt">observation</span> data. The results show that the proposed approaches outperform the state-of-the-art methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9033E..5HL','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9033E..5HL"><span>Validation of an <span class="hlt">image</span>-based technique to assess the perceptual quality of clinical chest radiographs with an <span class="hlt">observer</span> study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lin, Yuan; Choudhury, Kingshuk R.; McAdams, H. Page; Foos, David H.; Samei, Ehsan</p> <p>2014-03-01</p> <p>We previously proposed a novel <span class="hlt">image</span>-based quality assessment technique1 to assess the perceptual quality of clinical chest radiographs. In this paper, an <span class="hlt">observer</span> study was designed and conducted to systematically validate this technique. Ten metrics were involved in the <span class="hlt">observer</span> study, i.e., lung grey level, lung detail, lung noise, riblung contrast, rib sharpness, mediastinum detail, mediastinum noise, mediastinum alignment, subdiaphragm-lung contrast, and subdiaphragm area. For each metric, three tasks were successively presented to the <span class="hlt">observers</span>. In each task, six ROI <span class="hlt">images</span> were randomly presented in a row and <span class="hlt">observers</span> were asked to rank the <span class="hlt">images</span> only based on a designated quality and disregard the other qualities. A range slider on the top of the <span class="hlt">images</span> was used for <span class="hlt">observers</span> to indicate the acceptable range based on the corresponding perceptual attribute. Five boardcertificated radiologists from Duke participated in this <span class="hlt">observer</span> study on a DICOM calibrated diagnostic display workstation and under low ambient lighting conditions. The <span class="hlt">observer</span> data were analyzed in terms of the correlations between the <span class="hlt">observer</span> ranking orders and the algorithmic ranking orders. Based on the collected acceptable ranges, quality consistency ranges were statistically derived. The <span class="hlt">observer</span> study showed that, for each metric, the averaged ranking orders of the participated <span class="hlt">observers</span> were strongly correlated with the algorithmic orders. For the lung grey level, the <span class="hlt">observer</span> ranking orders completely accorded with the algorithmic ranking orders. The quality consistency ranges derived from this <span class="hlt">observer</span> study were close to these derived from our previous study. The <span class="hlt">observer</span> study indicates that the proposed <span class="hlt">image</span>-based quality assessment technique provides a robust reflection of the perceptual <span class="hlt">image</span> quality of the clinical chest radiographs. The derived quality consistency ranges can be used to automatically predict the acceptability of a clinical chest radiograph.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PMB....56.5771D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PMB....56.5771D"><span>Multi-<span class="hlt">observation</span> PET <span class="hlt">image</span> analysis for patient follow-up quantitation and therapy assessment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>David, S.; Visvikis, D.; Roux, C.; Hatt, M.</p> <p>2011-09-01</p> <p>In positron emission tomography (PET) <span class="hlt">imaging</span>, an early therapeutic response is usually characterized by variations of semi-quantitative parameters restricted to maximum SUV measured in PET scans during the treatment. Such measurements do not reflect overall tumor volume and radiotracer uptake variations. The proposed approach is based on multi-<span class="hlt">observation</span> <span class="hlt">image</span> analysis for merging several PET acquisitions to assess tumor metabolic volume and uptake variations. The fusion algorithm is based on iterative estimation using a stochastic expectation maximization (SEM) algorithm. The proposed method was applied to simulated and clinical follow-up PET <span class="hlt">images</span>. We compared the multi-<span class="hlt">observation</span> fusion performance to threshold-based methods, proposed for the assessment of the therapeutic response based on functional volumes. On simulated datasets the adaptive threshold applied independently on both <span class="hlt">images</span> led to higher errors than the ASEM fusion and on clinical datasets it failed to provide coherent measurements for four patients out of seven due to aberrant delineations. The ASEM method demonstrated improved and more robust estimation of the evaluation leading to more pertinent measurements. Future work will consist in extending the methodology and applying it to clinical multi-tracer datasets in order to evaluate its potential impact on the biological tumor volume definition for radiotherapy applications.</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://www.osti.gov/biblio/21455148-testing-hair-theorem-observations-electromagnetic-spectrum-ii-black-hole-images','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21455148-testing-hair-theorem-observations-electromagnetic-spectrum-ii-black-hole-images"><span>TESTING THE NO-HAIR THEOREM WITH <span class="hlt">OBSERVATIONS</span> IN THE ELECTROMAGNETIC SPECTRUM. II. BLACK HOLE <span class="hlt">IMAGES</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>Johannsen, Tim; Psaltis, Dimitrios, E-mail: timj@physics.arizona.ed, E-mail: dpsaltis@email.arizona.ed</p> <p>2010-07-20</p> <p>According to the no-hair theorem, all astrophysical black holes are fully described by their masses and spins. This theorem can be tested <span class="hlt">observationally</span> by measuring (at least) three different multipole moments of the spacetimes of black holes. In this paper, we analyze <span class="hlt">images</span> of black holes within a framework that allows us to calculate <span class="hlt">observables</span> in the electromagnetic spectrum as a function of the mass, spin, and, independently, the quadrupole moment of a black hole. We show that a deviation of the quadrupole moment from the expected Kerr value leads to <span class="hlt">images</span> of black holes that are either prolate ormore » oblate depending on the sign and magnitude of the deviation. In addition, there is a ring-like structure around the black hole shadow with a diameter of {approx}10 black hole masses that is substantially brighter than the <span class="hlt">image</span> of the underlying accretion flow and that is independent of the astrophysical details of accretion flow models. We show that the shape of this ring depends directly on the mass, spin, and quadrupole moment of the black hole and can be used for an independent measurement of all three parameters. In particular, we demonstrate that this ring is highly circular for a Kerr black hole with a spin a {approx}< 0.9 M, independent of the <span class="hlt">observer</span>'s inclination, but becomes elliptical and asymmetric if the no-hair theorem is violated. Near-future very long baseline interferometric <span class="hlt">observations</span> of Sgr A* will <span class="hlt">image</span> this ring and may allow for an <span class="hlt">observational</span> test of the no-hair theorem.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19720003234&hterms=dim+light&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Ddim%2Blight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720003234&hterms=dim+light&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Ddim%2Blight"><span>Experiment S030: Dim sky photography/orthicon</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dunkelman, L.; Mercer, R. D.; Ney, E. P.; Hemenway, C. L.</p> <p>1971-01-01</p> <p>During Gemini missions, the <span class="hlt">image</span> orthicon system was used to obtain photographic data on faint and diffuse astronomical phenomena. Results show that the photographs may be used to determine the <span class="hlt">airglow</span> geometry. Although it was sensitive, the original photographic system was unsuitable for use in the study of dim and diffuse astronomical light sources.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930017533','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930017533"><span>Wide-field direct CCD <span class="hlt">observations</span> supporting the Astro-1 Space Shuttle mission's Ultraviolet <span class="hlt">Imaging</span> Telescope</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hintzen, Paul; Angione, Ron; Talbert, Freddie; Cheng, K.-P.; Smith, Eric; Stecher, Theodore P.</p> <p>1993-01-01</p> <p>Wide field direct CCD <span class="hlt">observations</span> are being obtained to support and complement the vacuum-ultraviolet (VUV) <span class="hlt">images</span> provided by Astro's Ultraviolet <span class="hlt">Imaging</span> Telescope (UIT) during a Space Shuttle flight in December 1990. Because of the wide variety of projects addressed by UIT, the fields <span class="hlt">observed</span> include (1) galactic supernova remnants such as the Cygnus Loop and globular clusters such as Omega Cen and M79; (2) the Magellanic Clouds, M33, M81, and other galaxies in the Local Group; and (3) rich clusters of galaxies, principally the Perseus cluster and Abell 1367. Ground-based <span class="hlt">observations</span> have been obtained for virtually all of the Astro-1 UIT fields. The optical <span class="hlt">images</span> allow identification of individual UV sources in each field and provide the long baseline in wavelength necessary for accurate analysis of UV-bright sources. To facilitate use of our optical <span class="hlt">images</span> for analysis of UIT data and other projects, we plan to archive them, with the UIT <span class="hlt">images</span>, at the National Space Science Data Center (NSSDC), where they will be universally accessible via anonymous FTP. The UIT, one of three telescopes comprising the Astro spacecraft, is a 38-cm f/9 Ritchey-Chretien telescope on which high quantum efficiency, solar-blind <span class="hlt">image</span> tubes are used to record VUV <span class="hlt">images</span> on photographic film. Five filters with passbands centered between 1250A and 2500A provide both VUV colors and a measurement of extinction via the 2200A dust feature. The resulting calibrated VUV pictures are 40 arcminutes in diameter at 2.5 arcseconds resolution. The capabilities of UIT, therefore, complement HST's WFPC: the latter has 40 times greater collecting area, while UIT's usable field has 170 times WFPC's field area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22679605-decameter-stationary-type-iv-burst-imaging-observations-september','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22679605-decameter-stationary-type-iv-burst-imaging-observations-september"><span>A DECAMETER STATIONARY TYPE IV BURST IN <span class="hlt">IMAGING</span> <span class="hlt">OBSERVATIONS</span> ON 2014 SEPTEMBER 6</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>Koval, Artem; Chen, Yao; Feng, Shiwei</p> <p>2016-08-01</p> <p>First-of-its-kind radio <span class="hlt">imaging</span> of a decameter solar stationary type IV radio burst has been presented in this paper. On 2014 September 6 the <span class="hlt">observations</span> of type IV burst radio emission were carried out with the two-dimensional heliograph based on the Ukrainian T-shaped radio telescope (UTR-2), together with other telescope arrays. Starting at ∼09:55 UT and for ∼3 hr, the radio emission was kept within the <span class="hlt">observational</span> session of UTR-2. The interesting <span class="hlt">observation</span> covered the full evolution of this burst, “from birth to death.” During the event lifetime, two C-class solar X-ray flares with peak times 11:29 UT and 12:24 UTmore » took place. The time profile of this burst in radio has a double-humped shape that can be explained by injection of energetic electrons, accelerated by the two flares, into the burst source. According to the heliographic <span class="hlt">observations</span>, we suggest that the burst source was confined within a high coronal loop, which was part of a relatively slow coronal mass ejection. The latter has been developed for several hours before the onset of the event. Through analysis of about 1.5 × 10{sup 6} heliograms (3700 temporal frames with 4096 <span class="hlt">images</span> in each frame that correspond to the number of frequency channels), the radio burst source <span class="hlt">imaging</span> shows a fascinating dynamical evolution. Both space-based ( GOES , SDO , SOHO , STEREO ) data and various ground-based instrumentation (ORFEES, NDA, RSTO, NRH) records have been used for this study.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22325154-su-sample-size-dependence-model-observers-estimating-low-contrast-detection-performance-from-ct-images','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22325154-su-sample-size-dependence-model-observers-estimating-low-contrast-detection-performance-from-ct-images"><span>SU-E-I-46: Sample-Size Dependence of Model <span class="hlt">Observers</span> for Estimating Low-Contrast Detection Performance From CT <span class="hlt">Images</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>Reiser, I; Lu, Z</p> <p>2014-06-01</p> <p>Purpose: Recently, task-based assessment of diagnostic CT systems has attracted much attention. Detection task performance can be estimated using human <span class="hlt">observers</span>, or mathematical <span class="hlt">observer</span> models. While most models are well established, considerable bias can be introduced when performance is estimated from a limited number of <span class="hlt">image</span> samples. Thus, the purpose of this work was to assess the effect of sample size on bias and uncertainty of two channelized Hotelling <span class="hlt">observers</span> and a template-matching <span class="hlt">observer</span>. Methods: The <span class="hlt">image</span> data used for this study consisted of 100 signal-present and 100 signal-absent regions-of-interest, which were extracted from CT slices. The experimental conditions includedmore » two signal sizes and five different x-ray beam current settings (mAs). Human <span class="hlt">observer</span> performance for these <span class="hlt">images</span> was determined in 2-alternative forced choice experiments. These data were provided by the Mayo clinic in Rochester, MN. Detection performance was estimated from three <span class="hlt">observer</span> models, including channelized Hotelling <span class="hlt">observers</span> (CHO) with Gabor or Laguerre-Gauss (LG) channels, and a template-matching <span class="hlt">observer</span> (TM). Different sample sizes were generated by randomly selecting a subset of <span class="hlt">image</span> pairs, (N=20,40,60,80). <span class="hlt">Observer</span> performance was quantified as proportion of correct responses (PC). Bias was quantified as the relative difference of PC for 20 and 80 <span class="hlt">image</span> pairs. Results: For n=100, all <span class="hlt">observer</span> models predicted human performance across mAs and signal sizes. Bias was 23% for CHO (Gabor), 7% for CHO (LG), and 3% for TM. The relative standard deviation, σ(PC)/PC at N=20 was highest for the TM <span class="hlt">observer</span> (11%) and lowest for the CHO (Gabor) <span class="hlt">observer</span> (5%). Conclusion: In order to make <span class="hlt">image</span> quality assessment feasible in the clinical practice, a statistically efficient <span class="hlt">observer</span> model, that can predict performance from few samples, is needed. Our results identified two <span class="hlt">observer</span> models that may be suited for this task.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009SPIE.7263E..0PP','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009SPIE.7263E..0PP"><span>Optimization of medical <span class="hlt">imaging</span> display systems: using the channelized Hotelling <span class="hlt">observer</span> for detecting lung nodules: experimental study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Platisa, Ljiljana; Vansteenkiste, Ewout; Goossens, Bart; Marchessoux, Cédric; Kimpe, Tom; Philips, Wilfried</p> <p>2009-02-01</p> <p>Medical-<span class="hlt">imaging</span> systems are designed to aid medical specialists in a specific task. Therefore, the physical parameters of a system need to optimize the task performance of a human <span class="hlt">observer</span>. This requires measurements of human performance in a given task during the system optimization. Typically, psychophysical studies are conducted for this purpose. Numerical <span class="hlt">observer</span> models have been successfully used to predict human performance in several detection tasks. Especially, the task of signal detection using a channelized Hotelling <span class="hlt">observer</span> (CHO) in simulated <span class="hlt">images</span> has been widely explored. However, there are few studies done for clinically acquired <span class="hlt">images</span> that also contain anatomic noise. In this paper, we investigate the performance of a CHO in the task of detecting lung nodules in real radiographic <span class="hlt">images</span> of the chest. To evaluate variability introduced by the limited available data, we employ a commonly used study of a multi-reader multi-case (MRMC) scenario. It accounts for both case and reader variability. Finally, we use the "oneshot" methods to estimate the MRMC variance of the area under the ROC curve (AUC). The obtained AUC compares well to those reported for human <span class="hlt">observer</span> study on a similar data set. Furthermore, the "one-shot" analysis implies a fairly consistent performance of the CHO with the variance of AUC below 0.002. This indicates promising potential for numerical <span class="hlt">observers</span> in optimization of medical <span class="hlt">imaging</span> displays and encourages further investigation on the subject.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ISPAr42.3.2287Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ISPAr42.3.2287Z"><span>Possibility of Cloudless Optical Remote Sensing <span class="hlt">Images</span> Acquisition Study by Using Meteorological 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>Zhang, T.; Lei, B.; Hu, Y.; Liu, K.; Gan, Y.</p> <p>2018-04-01</p> <p>Optical remote sensing <span class="hlt">images</span> have been widely used in feature interpretation and geo-information extraction. All the fundamental applications of optical remote sensing, are greatly influenced by cloud coverage. Generally, the availability of cloudless <span class="hlt">images</span> depends on the meteorological conditions for a given area. In this study, the cloud total amount (CTA) products of the Fengyun (FY) satellite were introduced to explore the meteorological changes in a year over China. The cloud information of CTA products were tested by using ZY-3 satellite <span class="hlt">images</span> firstly. CTA products from 2006 to 2017 were used to get relatively reliable results. The window period of cloudless <span class="hlt">images</span> acquisition for different areas in China was then determined. This research provides a feasible way to get the cloudless <span class="hlt">images</span> acquisition window by using meteorological <span class="hlt">observations</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-iss029e006855.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-iss029e006855.html"><span>Air glow and Terminator view taken by the Expedition 29 crew</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2011-09-18</p> <p>ISS029-E-006855 (18 Sept. 2011) --- This is one of a series of night time <span class="hlt">images</span> photographed by one of the Expedition 29 crew members from the International Space Station. It features <span class="hlt">airglow</span>, Earth?s terminator and parts of the Central Pacific Ocean. Nadir coordinates are 10.11 degrees north latitude and 169.92 degrees west longitude.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.P13E..02R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.P13E..02R"><span>The Ultraviolet Spectrograph on the Europa Mission (Europa-UVS)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Retherford, K. D.; Gladstone, R.; Greathouse, T. K.; Steffl, A.; Davis, M. W.; Feldman, P. D.; McGrath, M. A.; Roth, L.; Saur, J.; Spencer, J. R.; Stern, S. A.; Pope, S.; Freeman, M. A.; Persyn, S. C.; Araujo, M. F.; Cortinas, S. C.; Monreal, R. M.; Persson, K. B.; Trantham, B. J.; Versteeg, M. H.; Walther, B. C.</p> <p>2015-12-01</p> <p>NASA's Europa multi-flyby mission is designed to provide a diversity of measurements suited to enrich our understanding of the potential habitability of this intriguing ocean world. The Europa mission's Ultraviolet Spectrograph, Europa-UVS, is the sixth in a series of successful ultraviolet <span class="hlt">imaging</span> spectrographs (Rosetta-Alice, New Horizons Pluto-Alice, LRO-LAMP) and, like JUICE-UVS (now under Phase B development), is largely based on the most recent of these to fly, Juno-UVS. Europa-UVS <span class="hlt">observes</span> photons in the 55-210 nm wavelength range, at moderate spectral and spatial resolution along a 7.5° slit. Three distinct apertures send light to the off-axis telescope mirror feeding the long-slit spectrograph: i) a main entrance <span class="hlt">airglow</span> port is used for most <span class="hlt">observations</span> (e.g., <span class="hlt">airglow</span>, aurora, surface mapping, and stellar occultations); ii) a high-spatial-resolution port consists of a small hole in an additional aperture door, and is used for detailed <span class="hlt">observations</span> of bright targets; and iii) a separate solar port allows for solar occultations, viewing at a 60° offset from the nominal payload boresight. Photon event time-tagging (pixel list mode) and programmable spectral <span class="hlt">imaging</span> (histogram mode) allow for <span class="hlt">observational</span> flexibility and optimal science data management. As on Juno-UVS, the effects of penetrating electron radiation on electronic parts and data quality are mitigated through contiguous shielding, filtering of pulse height amplitudes, management of high-voltage settings, and careful use of radiation-hard parts. The science goals of Europa-UVS are to: 1) Determine the composition & chemistry, source & sinks, and structure & variability of Europa's atmosphere, from equator to pole; 2) Search for and characterize active plumes in terms of global distribution, structure, composition, and variability; 3) Explore the surface composition & microphysics and their relation to endogenic & exogenic processes; and 4) Investigate how energy and mass flow in the Europa</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40830&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40830&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Earth <span class="hlt">observation</span> photo taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). This <span class="hlt">image</span> shows a 50 by 100 kilometer (30 by 60 mile) area of the Imperial Valley in Southern California and neighboring Mexico. The checkered patterns represent agricultural fields where different types of crops in different stages of growth are cultivated. The very bright areas are (top left to lower right) the U.S. towns of Brawley, Imperial, El Centro, Calexico and the Mexican city of Mexicali. The bright L-shaped line (upper right) is the All-American water canal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFMSA11A1579K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFMSA11A1579K"><span>The Sondrestrom Research Facility All-sky <span class="hlt">Imagers</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kendall, E. A.; Grill, M.; Gudmundsson, E.; Stromme, A.</p> <p>2010-12-01</p> <p> facility is a color all-sky <span class="hlt">imager</span> (CASI). The CASI instrument is a low-cost Keo Scientific Ltd. system similar to cameras designed for the THEMIS satellite ground-based <span class="hlt">imaging</span> network. This camera captures all visible wavelengths simultaneously at a higher data rate than the ASI. While it is not possible to resolve fine spectral features as with narrowband filters on the ASI, this camera provides context on wavelengths not covered by other <span class="hlt">imagers</span>, and makes it much simpler to distinguish clouds from <span class="hlt">airglow</span> and aurora. As with the ASI, this <span class="hlt">imager</span> collects data during periods of dark skies and the <span class="hlt">images</span> are posted to the web for community viewing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22645710-observer-evaluation-metal-artifact-reduction-algorithm-applied-head-neck-cone-beam-computed-tomographic-images','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22645710-observer-evaluation-metal-artifact-reduction-algorithm-applied-head-neck-cone-beam-computed-tomographic-images"><span><span class="hlt">Observer</span> Evaluation of a Metal Artifact Reduction Algorithm Applied to Head and Neck Cone Beam Computed Tomographic <span class="hlt">Images</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>Korpics, Mark; Surucu, Murat; Mescioglu, Ibrahim</p> <p></p> <p>Purpose and Objectives: To quantify, through an <span class="hlt">observer</span> study, the reduction in metal artifacts on cone beam computed tomographic (CBCT) <span class="hlt">images</span> using a projection-interpolation algorithm, on <span class="hlt">images</span> containing metal artifacts from dental fillings and implants in patients treated for head and neck (H&N) cancer. Methods and Materials: An interpolation-substitution algorithm was applied to H&N CBCT <span class="hlt">images</span> containing metal artifacts from dental fillings and implants. <span class="hlt">Image</span> quality with respect to metal artifacts was evaluated subjectively and objectively. First, 6 independent radiation oncologists were asked to rank randomly sorted blinded <span class="hlt">images</span> (before and after metal artifact reduction) using a 5-point rating scalemore » (1 = severe artifacts; 5 = no artifacts). Second, the standard deviation of different regions of interest (ROI) within each <span class="hlt">image</span> was calculated and compared with the mean rating scores. Results: The interpolation-substitution technique successfully reduced metal artifacts in 70% of the cases. From a total of 60 <span class="hlt">images</span> from 15 H&N cancer patients undergoing <span class="hlt">image</span> guided radiation therapy, the mean rating score on the uncorrected <span class="hlt">images</span> was 2.3 ± 1.1, versus 3.3 ± 1.0 for the corrected <span class="hlt">images</span>. The mean difference in ranking score between uncorrected and corrected <span class="hlt">images</span> was 1.0 (95% confidence interval: 0.9-1.2, P<.05). The standard deviation of each ROI significantly decreased after artifact reduction (P<.01). Moreover, a negative correlation between the mean rating score for each <span class="hlt">image</span> and the standard deviation of the oral cavity and bilateral cheeks was <span class="hlt">observed</span>. Conclusion: The interpolation-substitution algorithm is efficient and effective for reducing metal artifacts caused by dental fillings and implants on CBCT <span class="hlt">images</span>, as demonstrated by the statistically significant increase in <span class="hlt">observer</span> <span class="hlt">image</span> quality ranking and by the decrease in ROI standard deviation between uncorrected and corrected <span class="hlt">images</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4667371','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4667371"><span>Method for optimizing channelized quadratic <span class="hlt">observers</span> for binary classification of large-dimensional <span class="hlt">image</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>Kupinski, M. K.; Clarkson, E.</p> <p>2015-01-01</p> <p>We present a new method for computing optimized channels for channelized quadratic <span class="hlt">observers</span> (CQO) that is feasible for high-dimensional <span class="hlt">image</span> data. The method for calculating channels is applicable in general and optimal for Gaussian distributed <span class="hlt">image</span> data. Gradient-based algorithms for determining the channels are presented for five different information-based figures of merit (FOMs). Analytic solutions for the optimum channels for each of the five FOMs are derived for the case of equal mean data for both classes. The optimum channels for three of the FOMs under the equal mean condition are shown to be the same. This result is critical since some of the FOMs are much easier to compute. Implementing the CQO requires a set of channels and the first- and second-order statistics of channelized <span class="hlt">image</span> data from both classes. The dimensionality reduction from M measurements to L channels is a critical advantage of CQO since estimating <span class="hlt">image</span> statistics from channelized data requires smaller sample sizes and inverting a smaller covariance matrix is easier. In a simulation study we compare the performance of ideal and Hotelling <span class="hlt">observers</span> to CQO. The optimal CQO channels are calculated using both eigenanalysis and a new gradient-based algorithm for maximizing Jeffrey's divergence (J). Optimal channel selection without eigenanalysis makes the J-CQO on large-dimensional <span class="hlt">image</span> data feasible. PMID:26366764</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015ApJ...810...46H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015ApJ...810...46H"><span>Cool Transition Region Loops <span class="hlt">Observed</span> by the Interface Region <span class="hlt">Imaging</span> Spectrograph</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Zhenghua; Xia, Lidong; Li, Bo; Madjarska, Maria S.</p> <p>2015-09-01</p> <p>We report on the first Interface Region <span class="hlt">Imaging</span> Spectrograph (IRIS) study of cool transition region loops, a class of loops that has received little attention in the literature. A cluster of such loops was <span class="hlt">observed</span> on the solar disk in active region NOAA11934, in the Si iv 1402.8 Å spectral raster and 1400 Å slit-jaw <span class="hlt">images</span>. We divide the loops into three groups and study their dynamics. The first group comprises relatively stable loops, with 382-626 km cross-sections. <span class="hlt">Observed</span> Doppler velocities are suggestive of siphon flows, gradually changing from -10 km s-1 at one end to 20 km s-1 at the other end of the loops. Nonthermal velocities of 15 ˜ 25 km s-1 were determined. Magnetic cancellation with a rate of 1015 Mx s-1 is found at the blueshifted footpoints. These physical properties suggest that these loops are impulsively heated by magnetic reconnection, and the siphon flows play an important role in the energy redistribution. The second group corresponds to two footpoints rooted in mixed-magnetic-polarity regions, where magnetic cancellation with a rate of 1015 Mx s-1 and explosive-event line profiles with enhanced wings of up to 200 km s-1 were <span class="hlt">observed</span>. In the third group, interaction between two cool loop systems is <span class="hlt">observed</span>. Evidence for magnetic reconnection between the two loop systems is reflected in the explosive-event line profiles and magnetic cancellation with a rate of 3× {10}15 Mx s-1 <span class="hlt">observed</span> in the corresponding area. The IRIS has provided opportunity for in-depth investigations of cool transition region loops. Further numerical experiments are crucial for understanding their physics and their roles in the coronal heating processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2715171','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2715171"><span><span class="hlt">Observer</span> efficiency in discrimination tasks simulating malignant and benign breast lesions <span class="hlt">imaged</span> with ultrasound</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Abbey, Craig K.; Zemp, Roger J.; Liu, Jie; Lindfors, Karen K.; Insana, Michael F.</p> <p>2009-01-01</p> <p>We investigate and extend the ideal <span class="hlt">observer</span> methodology developed by Smith and Wagner to detection and discrimination tasks related to breast sonography. We provide a numerical approach for evaluating the ideal <span class="hlt">observer</span> acting on radio-frequency (RF) frame data, which involves inversion of large nonstationary covariance matrices, and we describe a power-series approach to computing this inverse. Considering a truncated power series suggests that the RF data be Wiener-filtered before forming the final envelope <span class="hlt">image</span>. We have compared human performance for Wiener-filtered and conventional B-mode envelope <span class="hlt">images</span> using psychophysical studies for 5 tasks related to breast cancer classification. We find significant improvements in visual detection and discrimination efficiency in four of these five tasks. We also use the Smith-Wagner approach to distinguish between human and processing inefficiencies, and find that generally the principle limitation comes from the information lost in computing the final envelope <span class="hlt">image</span>. PMID:16468454</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014amos.confE..86O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014amos.confE..86O"><span>Optical <span class="hlt">Observation</span>, <span class="hlt">Image</span>-processing, and Detection of Space Debris in Geosynchronous Earth Orbit</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oda, H.; Yanagisawa, T.; Kurosaki, H.; Tagawa, M.</p> <p>2014-09-01</p> <p>We report on optical <span class="hlt">observations</span> and an efficient detection method of space debris in the geosynchronous Earth orbit (GEO). We operate our new Australia Remote Observatory (ARO) where an 18 cm optical telescope with a charged-coupled device (CCD) camera covering a 3.14-degree field of view is used for GEO debris survey, and analyse datasets of successive CCD <span class="hlt">images</span> using the line detection method (Yanagisawa and Nakajima 2005). In our operation, the exposure time of each CCD <span class="hlt">image</span> is set to be 3 seconds (or 5 seconds), and the time interval of CCD shutter open is about 4.7 seconds (or 6.7 seconds). In the line detection method, a sufficient number of sample objects are taken from each <span class="hlt">image</span> based on their shape and intensity, which includes not only faint signals but also background noise (we take 500 sample objects from each <span class="hlt">image</span> in this paper). Then we search a sequence of sample objects aligning in a straight line in the successive <span class="hlt">images</span> to exclude the noise sample. We succeed in detecting faint signals (down to about 1.8 sigma of background noise) by applying the line detection method to 18 CCD <span class="hlt">images</span>. As a result, we detected about 300 GEO objects up to magnitude of 15.5 among 5 nights data. We also calculate orbits of objects detected using the Simplified General Perturbations Satellite Orbit Model 4(SGP4), and identify the objects listed in the two-line-element (TLE) data catalogue publicly provided by the U.S. Strategic Command (USSTRATCOM). We found that a certain amount of our detections are new objects that are not contained in the catalogue. We conclude that our ARO and detection method posse a high efficiency detection of GEO objects despite the use of comparatively-inexpensive <span class="hlt">observation</span> and analysis system. We also describe the <span class="hlt">image</span>-processing specialized for the detection of GEO objects (not for usual astronomical objects like stars) in this paper.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870060961&hterms=Evidence+atomic+theory&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DEvidence%2Batomic%2Btheory','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870060961&hterms=Evidence+atomic+theory&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DEvidence%2Batomic%2Btheory"><span>Hydrogen Balmer alpha intensity distributions and line profiles from multiple scattering theory using realistic geocoronal models</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Anderson, D. E., Jr.; Meier, R. R.; Hodges, R. R., Jr.; Tinsley, B. A.</p> <p>1987-01-01</p> <p>The H Balmer alpha nightglow is investigated by using Monte Carlo models of asymmetric geocoronal atomic hydrogen distributions as input to a radiative transfer model of solar Lyman-beta radiation in the thermosphere and atmosphere. It is shown that it is essential to include multiple scattering of Lyman-beta radiation in the interpretation of Balmer alpha <span class="hlt">airglow</span> data. <span class="hlt">Observations</span> of diurnal variation in the Balmer alpha <span class="hlt">airglow</span> showing slightly greater intensities in the morning relative to evening are consistent with theory. No evidence is found for anything other than a single sinusoidal diurnal variation of exobase density. Dramatic changes in effective temperature derived from the <span class="hlt">observed</span> Balmer alpha line profiles are expected on the basis of changing illumination conditions in the thermosphere and exosphere as different regions of the sky are scanned.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992lest.rept....3K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992lest.rept....3K"><span>GAP: yet another <span class="hlt">image</span> processing system for solar <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>Keller, C. U.</p> <p></p> <p>GAP is a versatile, interactive <span class="hlt">image</span> processing system for analyzing solar <span class="hlt">observations</span>, in particular extended time sequences, and for preparing publication quality figures. It consists of an interpreter that is based on a language with a control flow similar to PASCAL and C. The interpreter may be accessed from a command line editor and from user-supplied functions, procedures, and command scripts. GAP is easily expandable via external FORTRAN programs that are linked to the GAP interface routines. The current version of GAP runs on VAX, DECstation, Sun, and Apollo computers. Versions for MS-DOS and OS/2 are in preparation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890002224','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890002224"><span>Structural anomalies in undoped Gallium Arsenide <span class="hlt">observed</span> in high resolution diffraction <span class="hlt">imaging</span> with monochromatic synchrotron radiation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Steiner, B.; Kuriyama, M.; Dobbyn, R. C.; Laor, U.; Larson, D.; Brown, M.</p> <p>1988-01-01</p> <p>Novel, streak-like disruption features restricted to the plane of diffraction have recently been <span class="hlt">observed</span> in <span class="hlt">images</span> obtained by synchrotron radiation diffraction from undoped, semi-insulating gallium arsenide crystals. These features were identified as ensembles of very thin platelets or interfaces lying in (110) planes, and a structural model consisting of antiphase domain boundaries was proposed. We report here the other principal features <span class="hlt">observed</span> in high resolution monochromatic synchrotron radiation diffraction <span class="hlt">images</span>: (quasi) cellular structure; linear, very low-angle subgrain boundaries in (110) directions, and surface stripes in a (110) direction. In addition, we report systematic differences in the acceptance angle for <span class="hlt">images</span> involving various diffraction vectors. When these <span class="hlt">observations</span> are considered together, a unifying picture emerges. The presence of ensembles of thin (110) antiphase platelet regions or boundaries is generally consistent not only with the streak-like diffraction features but with the other features reported here as well. For the formation of such regions we propose two mechanisms, operating in parallel, that appear to be consistent with the various defect features <span class="hlt">observed</span> by a variety of techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890058656&hterms=gallium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dgallium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890058656&hterms=gallium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dgallium"><span>Structural anomalies in undoped gallium arsenide <span class="hlt">observed</span> in high-resolution diffraction <span class="hlt">imaging</span> with monochromatic synchrotron radiation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Steiner, B.; Kuriyama, M.; Dobbyn, R. C.; Laor, U.; Larson, D.</p> <p>1989-01-01</p> <p>Novel, streak-like disruption features restricted to the plane of diffraction have recently been <span class="hlt">observed</span> in <span class="hlt">images</span> obtained by synchrotron radiation diffraction from undoped, semi-insulating gallium arsenide crystals. These features were identified as ensembles of very thin platelets or interfaces lying in (110) planes, and a structural model consisting of antiphase domain boundaries was proposed. We report here the other principal features <span class="hlt">observed</span> in high resolution monochromatic synchrotron radiation diffraction <span class="hlt">images</span>: (quasi) cellular structure; linear, very low-angle subgrain boundaries in (110) directions, and surface stripes in a (110) direction. In addition, we report systematic differences in the acceptance angle for <span class="hlt">images</span> involving various diffraction vectors. When these <span class="hlt">observations</span> are considered together, a unifying picture emerges. The presence of ensembles of thin (110) antiphase platelet regions or boundaries is generally consistent not only with the streak-like diffraction features but with the other features reported here as well. For the formation of such regions we propose two mechanisms, operating in parallel, that appear to be consistent with the various defect features <span class="hlt">observed</span> by a variety of techniques.</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('https://www.ncbi.nlm.nih.gov/pubmed/28983493','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28983493"><span>Use of a channelized Hotelling <span class="hlt">observer</span> to assess CT <span class="hlt">image</span> quality and optimize dose reduction for iteratively reconstructed <span class="hlt">images</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Favazza, Christopher P; Ferrero, Andrea; Yu, Lifeng; Leng, Shuai; McMillan, Kyle L; McCollough, Cynthia H</p> <p>2017-07-01</p> <p>The use of iterative reconstruction (IR) algorithms in CT generally decreases <span class="hlt">image</span> noise and enables dose reduction. However, the amount of dose reduction possible using IR without sacrificing diagnostic performance is difficult to assess with conventional <span class="hlt">image</span> quality metrics. Through this investigation, achievable dose reduction using a commercially available IR algorithm without loss of low contrast spatial resolution was determined with a channelized Hotelling <span class="hlt">observer</span> (CHO) model and used to optimize a clinical abdomen/pelvis exam protocol. A phantom containing 21 low contrast disks-three different contrast levels and seven different diameters-was <span class="hlt">imaged</span> at different dose levels. <span class="hlt">Images</span> were created with filtered backprojection (FBP) and IR. The CHO was tasked with detecting the low contrast disks. CHO performance indicated dose could be reduced by 22% to 25% without compromising low contrast detectability (as compared to full-dose FBP <span class="hlt">images</span>) whereas 50% or more dose reduction significantly reduced detection performance. Importantly, default settings for the scanner and protocol investigated reduced dose by upward of 75%. Subsequently, CHO-based protocol changes to the default protocol yielded <span class="hlt">images</span> of higher quality and doses more consistent with values from a larger, dose-optimized scanner fleet. CHO assessment provided objective data to successfully optimize a clinical CT acquisition protocol.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGRA..121.9204L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGRA..121.9204L"><span>Characteristics of mesospheric gravity waves over the southeastern Tibetan Plateau region</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Qinzeng; Xu, Jiyao; Liu, Xiao; Yuan, Wei; Chen, Jinsong</p> <p>2016-09-01</p> <p>The Tibetan Plateau (TP), known as "Third Pole" of the Earth, has important influences on global climates and local weather. An important objective in present study is to investigate how orographic features of the TP affect the geographical distributions of gravity wave (GW) sources. Three-year OH <span class="hlt">airglow</span> <span class="hlt">images</span> (November 2011 to October 2014) from Qujing (25.6°N, 103.7°E) were used to study the characteristics of GWs over the southeastern TP region. Along with the almost concurrent and collocated meteor radar wind measurements and temperature data from SABER/TIMED satellite, the propagation conditions of three types of GWs (freely propagating, ducted, or evanescent) were estimated. Most of GWs exhibited ducted or evanescent characteristics. Almost all GWs propagate southeastward in winter. The GW propagation directions in winter are significantly different from other <span class="hlt">airglow</span> <span class="hlt">imager</span> <span class="hlt">observations</span> at northern middle latitudes. Wind data and convective precipitation fields from the European Centre for Medium-Range Weather Forecasts reanalysis data are used to study the sources of GWs on the edge of the TP. Using backward ray-tracing analysis, we find that most of the mesospheric freely propagating GWs are located in or near the large wind shear intensity region ( 10 km- 17 km) on the southeastern edge of the TP in spring and winter. The averaged value of momentum flux is 11.6 ± 5.2 m2/s2 in winter and 7.5 ± 3.1 m2/s2 in summer. This work will provide valuable information for the GW parameterization schemes in general circulation models in TP region.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApJ...833...22C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApJ...833...22C"><span>Undercover EUV Solar Jets <span class="hlt">Observed</span> by the Interface Region <span class="hlt">Imaging</span> Spectrograph</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, N.-H.; Innes, D. E.</p> <p>2016-12-01</p> <p>It is well-known that extreme ultraviolet (EUV) emission emitted at the solar surface is absorbed by overlying cool plasma. Especially in active regions, dark lanes in EUV <span class="hlt">images</span> suggest that much of the surface activity is obscured. Simultaneous <span class="hlt">observations</span> from the Interface Region <span class="hlt">Imaging</span> Spectrograph, consisting of UV spectra and slit-jaw <span class="hlt">images</span> (SJI), give vital information with sub-arcsecond spatial resolution on the dynamics of jets not seen in EUV <span class="hlt">images</span>. We studied a series of small jets from recently formed bipole pairs beside the trailing spot of active region 11991, which occurred on 2014 March 5 from 15:02:21 UT to 17:04:07 UT. Collimated outflows with bright roots were present in SJI 1400 Å (transition region) and 2796 Å (upper chromosphere) that were mostly not seen in Atmospheric <span class="hlt">Imaging</span> Assembly (AIA) 304 Å (transition region) and AIA 171 Å (lower corona) <span class="hlt">images</span>. The Si IV spectra show a strong blue wing enhancement, but no red wing, in the line profiles of the ejecta for all recurrent jets, indicating outward flows without twists. We see two types of Mg II line profiles produced by the jets spires: reversed and non-reversed. Mg II lines remain optically thick, but turn optically thin in the highly Doppler shifted wings. The energy flux contained in each recurrent jet is estimated using a velocity differential emission measure technique that measures the emitting power of the plasma as a function of the line-of-sight velocity. We found that all the recurrent jets release similar energy (108 erg cm-2 s-1) toward the corona and the downward component is less than 3%.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AAS...22930303S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AAS...22930303S"><span><span class="hlt">Imaging</span> Protoplanets: <span class="hlt">Observing</span> Transition Disks with Non-Redundant Masking</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sallum, Stephanie</p> <p>2017-01-01</p> <p>Transition disks - protoplanetary disks with inner, solar system sized clearings - may be shaped by young planets. Directly <span class="hlt">imaging</span> protoplanets in these objects requires high contrast and resolution, making them promising targets for future extremely large telescopes. The interferometric technique of non-redundant masking (NRM) is well suited for these <span class="hlt">observations</span>, enabling companion detection for contrasts of 1:100 - 1:1000 at or within the diffraction limit. My dissertation focuses on searching for and characterizing companions in transition disk clearings using NRM. I will briefly describe the technique and present spatially resolved <span class="hlt">observations</span> of the T Cha and LkCa 15 transition disks. Both of these objects hosted posited substellar companions. However multi-epoch T Cha datasets cannot be explained by planets orbiting in the disk plane. Conversely, LkCa 15 data taken with the Large Binocular Telescope (LBT) in single-aperture mode reveal the presence of multiple forming planets. The dual aperture LBT will provide triple the angular resolution of these <span class="hlt">observations</span>, dramatically increasing the phase space for exoplanet detection. I will also present new results from the dual-aperture LBT, with similar resolution to that expected for next generation facilities like GMT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19900003993','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19900003993"><span>EIT: Solar corona synoptic <span class="hlt">observations</span> from SOHO with an Extreme-ultraviolet <span class="hlt">Imaging</span> Telescope</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Delaboudiniere, J. P.; Gabriel, A. H.; Artzner, G. E.; Michels, D. J.; Dere, K. P.; Howard, R. A.; Catura, R.; Stern, R.; Lemen, J.; Neupert, W.</p> <p>1988-01-01</p> <p>The Extreme-ultraviolet <span class="hlt">Imaging</span> Telescope (EIT) of SOHO (solar and heliospheric observatory) will provide full disk <span class="hlt">images</span> in emission lines formed at temperatures that map solar structures ranging from the chromospheric network to the hot magnetically confined plasma in the corona. <span class="hlt">Images</span> in four narrow bandpasses will be obtained using normal incidence multilayered optics deposited on quadrants of a Ritchey-Chretien telescope. The EIT is capable of providing a uniform one arc second resolution over its entire 50 by 50 arc min field of view. Data from the EIT will be extremely valuable for identifying and interpreting the spatial and temperature fine structures of the solar atmosphere. Temporal analysis will provide information on the stability of these structures and identify dynamical processes. EIT <span class="hlt">images</span>, issued daily, will provide the global corona context for aid in unifying the investigations and in forming the <span class="hlt">observing</span> plans for SOHO coronal instruments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9913E..4GT','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9913E..4GT"><span><span class="hlt">Image</span> processing improvement for optical <span class="hlt">observations</span> of space debris with the TAROT telescopes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thiebaut, C.; Theron, S.; Richard, P.; Blanchet, G.; Klotz, A.; Boër, M.</p> <p>2016-07-01</p> <p>CNES is involved in the Inter-Agency Space Debris Coordination Committee (IADC) and is <span class="hlt">observing</span> space debris with two robotic ground based fully automated telescopes called TAROT and operated by the CNRS. An <span class="hlt">image</span> processing algorithm devoted to debris detection in geostationary orbit is implemented in the standard pipeline. Nevertheless, this algorithm is unable to deal with debris tracking mode <span class="hlt">images</span>, this mode being the preferred one for debris detectability. We present an algorithm improvement for this mode and give results in terms of false detection rate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4651348','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4651348"><span>Pigeons (Columba livia) as Trainable <span class="hlt">Observers</span> of Pathology and Radiology Breast Cancer <span class="hlt">Images</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>Levenson, Richard M.; Krupinski, Elizabeth A.; Navarro, Victor M.; Wasserman, Edward A.</p> <p>2015-01-01</p> <p>Pathologists and radiologists spend years acquiring and refining their medically essential visual skills, so it is of considerable interest to understand how this process actually unfolds and what <span class="hlt">image</span> features and properties are critical for accurate diagnostic performance. Key insights into human behavioral tasks can often be obtained by using appropriate animal models. We report here that pigeons (Columba livia)—which share many visual system properties with humans—can serve as promising surrogate <span class="hlt">observers</span> of medical <span class="hlt">images</span>, a capability not previously documented. The birds proved to have a remarkable ability to distinguish benign from malignant human breast histopathology after training with differential food reinforcement; even more importantly, the pigeons were able to generalize what they had learned when confronted with novel <span class="hlt">image</span> sets. The birds’ histological accuracy, like that of humans, was modestly affected by the presence or absence of color as well as by degrees of <span class="hlt">image</span> compression, but these impacts could be ameliorated with further training. Turning to radiology, the birds proved to be similarly capable of detecting cancer-relevant microcalcifications on mammogram <span class="hlt">images</span>. However, when given a different (and for humans quite difficult) task—namely, classification of suspicious mammographic densities (masses)—the pigeons proved to be capable only of <span class="hlt">image</span> memorization and were unable to successfully generalize when shown novel examples. The birds’ successes and difficulties suggest that pigeons are well-suited to help us better understand human medical <span class="hlt">image</span> perception, and may also prove useful in performance assessment and development of medical <span class="hlt">imaging</span> hardware, <span class="hlt">image</span> processing, and <span class="hlt">image</span> analysis tools. PMID:26581091</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26581091','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26581091"><span>Pigeons (Columba livia) as Trainable <span class="hlt">Observers</span> of Pathology and Radiology Breast Cancer <span class="hlt">Images</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Levenson, Richard M; Krupinski, Elizabeth A; Navarro, Victor M; Wasserman, Edward A</p> <p>2015-01-01</p> <p>Pathologists and radiologists spend years acquiring and refining their medically essential visual skills, so it is of considerable interest to understand how this process actually unfolds and what <span class="hlt">image</span> features and properties are critical for accurate diagnostic performance. Key insights into human behavioral tasks can often be obtained by using appropriate animal models. We report here that pigeons (Columba livia)-which share many visual system properties with humans-can serve as promising surrogate <span class="hlt">observers</span> of medical <span class="hlt">images</span>, a capability not previously documented. The birds proved to have a remarkable ability to distinguish benign from malignant human breast histopathology after training with differential food reinforcement; even more importantly, the pigeons were able to generalize what they had learned when confronted with novel <span class="hlt">image</span> sets. The birds' histological accuracy, like that of humans, was modestly affected by the presence or absence of color as well as by degrees of <span class="hlt">image</span> compression, but these impacts could be ameliorated with further training. Turning to radiology, the birds proved to be similarly capable of detecting cancer-relevant microcalcifications on mammogram <span class="hlt">images</span>. However, when given a different (and for humans quite difficult) task-namely, classification of suspicious mammographic densities (masses)-the pigeons proved to be capable only of <span class="hlt">image</span> memorization and were unable to successfully generalize when shown novel examples. The birds' successes and difficulties suggest that pigeons are well-suited to help us better understand human medical <span class="hlt">image</span> perception, and may also prove useful in performance assessment and development of medical <span class="hlt">imaging</span> hardware, <span class="hlt">image</span> processing, and <span class="hlt">image</span> analysis tools.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSH51D..02L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSH51D..02L"><span>Solar Polar <span class="hlt">Imager</span>: <span class="hlt">Observing</span> Coronal Transients from a New Perspective (Invited)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liewer, P. C.</p> <p>2013-12-01</p> <p>The heliophysics community has long recognized the need for a mission to <span class="hlt">observe</span> the Sun and corona from a polar perspective. One mission concept, the Solar Polar <span class="hlt">Imager</span> (SPI), has been studied extensively (Liewer et al in NASA Space Science Vision Missions, 2008). In this concept, a solar sail is used to place a spacecraft in a circular 0.48-AU heliocentric orbit with an inclination of ~75 degrees. This orbit enables crucial <span class="hlt">observations</span> not possible from lower latitude perspectives. Magnetograph and Doppler <span class="hlt">observations</span> from a polar vantage point would revolutionize our understanding of the mechanism of solar activity cycles, polar magnetic field reversals, the internal structure and dynamics of the Sun and its atmosphere. The rapid 4-month polar orbit combined with both in situ and remote sensing instrumentation further enables unprecedented studies of the physical connection between the Sun, the solar wind, and solar energetic particles. From the polar perspective, white light <span class="hlt">imagers</span> could be used to track CMEs and predict their arrival at Earth (as demonstrated by STEREO). SPI is also well suited to study the relative roles of CME-driven shock versus flare-associated processes in solar energetic particle acceleration. With the circular 0.48 AU orbit, solar energetic particles could be more easily traced to their sources and their variation with latitude can be studied at a constant radius. This talk will discuss the science objectives, instrumentation and mission design for the SPI mission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ASPC..483..319S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ASPC..483..319S"><span><span class="hlt">Imaging</span> the Moon II: Webcam CCD <span class="hlt">Observations</span> and Analysis (a Two-Week Lab for Non-Majors)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sato, T.</p> <p>2014-07-01</p> <p><span class="hlt">Imaging</span> the Moon is a successful two-week lab involving real sky <span class="hlt">observations</span> of the Moon in which students make telescopic <span class="hlt">observations</span> and analyze their own <span class="hlt">images</span>. Originally developed around the 35 mm film camera, a common household object adapted for astronomical work, the lab now uses webcams as film photography has evolved into an obscure specialty technology and increasing numbers of students have little familiarity with it. The printed circuit board with the CCD is harvested from a commercial webcam and affixed to a tube to mount on a telescope in place of an eyepiece. <span class="hlt">Image</span> frames are compiled to form a lunar mosaic, and crater sizes are measured. Students also work through the logistical steps of telescope time assignment and scheduling. They learn to keep a schedule and work with uncertainties of weather in ways paralleling research <span class="hlt">observations</span>. Because there is no need for a campus observatory, this lab can be replicated at a wide variety of institutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SPIE.9652E..0KG','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SPIE.9652E..0KG"><span>Versatile illumination platform and fast optical switch to give standard <span class="hlt">observation</span> camera gated active <span class="hlt">imaging</span> capacity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grasser, R.; Peyronneaudi, Benjamin; Yon, Kevin; Aubry, Marie</p> <p>2015-10-01</p> <p>CILAS, subsidiary of Airbus Defense and Space, develops, manufactures and sales laser-based optronics equipment for defense and homeland security applications. Part of its activity is related to active systems for threat detection, recognition and identification. Active surveillance and active <span class="hlt">imaging</span> systems are often required to achieve identification capacity in case for long range <span class="hlt">observation</span> in adverse conditions. In order to ease the deployment of active <span class="hlt">imaging</span> systems often complex and expensive, CILAS suggests a new concept. It consists on the association of two apparatus working together. On one side, a patented versatile laser platform enables high peak power laser illumination for long range <span class="hlt">observation</span>. On the other side, a small camera add-on works as a fast optical switch to select photons with specific time of flight only. The association of the versatile illumination platform and the fast optical switch presents itself as an independent body, so called "flash module", giving to virtually any passive <span class="hlt">observation</span> systems gated active <span class="hlt">imaging</span> capacity in NIR and SWIR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29475259','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29475259"><span>Dual-polarized light-field <span class="hlt">imaging</span> micro-system via a liquid-crystal microlens array for direct three-dimensional <span class="hlt">observation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xin, Zhaowei; Wei, Dong; Xie, Xingwang; Chen, Mingce; Zhang, Xinyu; Liao, Jing; Wang, Haiwei; Xie, Changsheng</p> <p>2018-02-19</p> <p>Light-field <span class="hlt">imaging</span> is a crucial and straightforward way of measuring and analyzing surrounding light worlds. In this paper, a dual-polarized light-field <span class="hlt">imaging</span> micro-system based on a twisted nematic liquid-crystal microlens array (TN-LCMLA) for direct three-dimensional (3D) <span class="hlt">observation</span> is fabricated and demonstrated. The prototyped camera has been constructed by integrating a TN-LCMLA with a common CMOS sensor array. By switching the working state of the TN-LCMLA, two orthogonally polarized light-field <span class="hlt">images</span> can be remapped through the functioned <span class="hlt">imaging</span> sensors. The <span class="hlt">imaging</span> micro-system in conjunction with the electric-optical microstructure can be used to perform polarization and light-field <span class="hlt">imaging</span>, simultaneously. Compared with conventional plenoptic cameras using liquid-crystal microlens array, the polarization-independent light-field <span class="hlt">images</span> with a high <span class="hlt">image</span> quality can be obtained in the arbitrary polarization state selected. We experimentally demonstrate characters including a relatively wide operation range in the manipulation of incident beams and the multiple <span class="hlt">imaging</span> modes, such as conventional two-dimensional <span class="hlt">imaging</span>, light-field <span class="hlt">imaging</span>, and polarization <span class="hlt">imaging</span>. Considering the obvious features of the TN-LCMLA, such as very low power consumption, providing multiple <span class="hlt">imaging</span> modes mentioned, simple and low-cost manufacturing, the <span class="hlt">imaging</span> micro-system integrated with this kind of liquid-crystal microstructure driven electrically presents the potential capability of directly <span class="hlt">observing</span> a 3D object in typical scattering media.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29708194','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29708194"><span>A simple method for low-contrast detectability, <span class="hlt">image</span> quality and dose optimisation with CT iterative reconstruction algorithms and model <span class="hlt">observers</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bellesi, Luca; Wyttenbach, Rolf; Gaudino, Diego; Colleoni, Paolo; Pupillo, Francesco; Carrara, Mauro; Braghetti, Antonio; Puligheddu, Carla; Presilla, Stefano</p> <p>2017-01-01</p> <p>The aim of this work was to evaluate detection of low-contrast objects and <span class="hlt">image</span> quality in computed tomography (CT) phantom <span class="hlt">images</span> acquired at different tube loadings (i.e. mAs) and reconstructed with different algorithms, in order to find appropriate settings to reduce the dose to the patient without any <span class="hlt">image</span> detriment. <span class="hlt">Images</span> of supraslice low-contrast objects of a CT phantom were acquired using different mAs values. <span class="hlt">Images</span> were reconstructed using filtered back projection (FBP), hybrid and iterative model-based methods. <span class="hlt">Image</span> quality parameters were evaluated in terms of modulation transfer function; noise, and uniformity using two software resources. For the definition of low-contrast detectability, studies based on both human (i.e. four-alternative forced-choice test) and model <span class="hlt">observers</span> were performed across the various <span class="hlt">images</span>. Compared to FBP, <span class="hlt">image</span> quality parameters were improved by using iterative reconstruction (IR) algorithms. In particular, IR model-based methods provided a 60% noise reduction and a 70% dose reduction, preserving <span class="hlt">image</span> quality and low-contrast detectability for human radiological evaluation. According to the model <span class="hlt">observer</span>, the diameters of the minimum detectable detail were around 2 mm (up to 100 mAs). Below 100 mAs, the model <span class="hlt">observer</span> was unable to provide a result. IR methods improve CT protocol quality, providing a potential dose reduction while maintaining a good <span class="hlt">image</span> detectability. Model <span class="hlt">observer</span> can in principle be useful to assist human performance in CT low-contrast detection tasks and in dose optimisation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010003574','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010003574"><span><span class="hlt">Observations</span> of Leonid Meteors Using a Mid-Wave Infrared <span class="hlt">Imaging</span> Spectrograph</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rossano, G. S.; Russell, R. W.; Lynch, D. K.; Tessensohn, T. K.; Warren, D.; Jenniskens, P.; DeVincenzi, Donald L. (Technical Monitor)</p> <p>2000-01-01</p> <p>We report broadband 3-5.5 micrometer detections of two Leonid meteors <span class="hlt">observed</span> during the 1998 Leonid Multi-Instrument Aircraft Campaign. Each meteor was detected at only one position along their trajectory just prior to the point of maximum light emission. We describe the particular aspects of the Aerospace Corp. Mid-wave Infra-Red <span class="hlt">Imaging</span> Spectrograph (MIRIS) developed for the <span class="hlt">observation</span> of short duration transient events that impact its ability to detect Leonid meteors. This instrument had its first deployment during the 1998 Leonid MAC. We infer from our <span class="hlt">observations</span> that the mid-infrared light curves of two Leonid meteors differed from the visible light curve. At the points of detection, the infrared emission in the MIRIS passband was 25 +/- 4 times that at optical wavelengths for both meteors. In addition, we find an upper limit of 800 K for the solid body temperature of the brighter meteor we <span class="hlt">observed</span>, at the point in the trajectory where we made our mid-wave infrared detection.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19810065306&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810065306&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dtwilight"><span><span class="hlt">Observations</span> of the Ca/+/ twilight <span class="hlt">airglow</span> from intermediate layers of ionization</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tepley, C. A.; Meriwether, J. W., Jr.; Walker, J. C. G.; Mathews, J. D.</p> <p>1981-01-01</p> <p>Optical and incoherent scatter radar techniques are applied to detect the presence of Ca(+) in lower thermospheric intermediate layers over Arecibo. The Arecibo 430 MHz radar is used to measure electron densities, and the altitude distribution and density of the calcium ion is inferred from the variation of twilight resonant scattering with solar depression angle. Ca(+) and electron column densities are compared, and results indicate that the composition of low-altitude intermediate layers is 2% Ca(+), which is consistent with rocket mass spectrometer measurements. Fe(+) and Mg(+) ultraviolet resonance lines are not detected from the ground due to ozone absorbing all radiation short of 3000 A, and measurements of the neutral iron resonance line at 3860 A show that an atmospheric continuum may result in overestimations of emission rates at high solar depression angles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22660958-undercover-euv-solar-jets-observed-interface-region-imaging-spectrograph','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22660958-undercover-euv-solar-jets-observed-interface-region-imaging-spectrograph"><span>UNDERCOVER EUV SOLAR JETS <span class="hlt">OBSERVED</span> BY THE INTERFACE REGION <span class="hlt">IMAGING</span> SPECTROGRAPH</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>Chen, N.-H.; Innes, D. E.</p> <p></p> <p>It is well-known that extreme ultraviolet (EUV) emission emitted at the solar surface is absorbed by overlying cool plasma. Especially in active regions, dark lanes in EUV <span class="hlt">images</span> suggest that much of the surface activity is obscured. Simultaneous <span class="hlt">observations</span> from the Interface Region <span class="hlt">Imaging</span> Spectrograph, consisting of UV spectra and slit-jaw <span class="hlt">images</span> (SJI), give vital information with sub-arcsecond spatial resolution on the dynamics of jets not seen in EUV <span class="hlt">images</span>. We studied a series of small jets from recently formed bipole pairs beside the trailing spot of active region 11991, which occurred on 2014 March 5 from 15:02:21 UT tomore » 17:04:07 UT. Collimated outflows with bright roots were present in SJI 1400 Å (transition region) and 2796 Å (upper chromosphere) that were mostly not seen in Atmospheric <span class="hlt">Imaging</span> Assembly (AIA) 304 Å (transition region) and AIA 171 Å (lower corona) <span class="hlt">images</span>. The Si iv spectra show a strong blue wing enhancement, but no red wing, in the line profiles of the ejecta for all recurrent jets, indicating outward flows without twists. We see two types of Mg ii line profiles produced by the jets spires: reversed and non-reversed. Mg ii lines remain optically thick, but turn optically thin in the highly Doppler shifted wings. The energy flux contained in each recurrent jet is estimated using a velocity differential emission measure technique that measures the emitting power of the plasma as a function of the line-of-sight velocity. We found that all the recurrent jets release similar energy (10{sup 8} erg cm{sup −2} s{sup −1}) toward the corona and the downward component is less than 3%.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9147E..83W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9147E..83W"><span>Gemini planet <span class="hlt">imager</span> <span class="hlt">observational</span> calibrations VII: on-sky polarimetric performance of the Gemini planet <span class="hlt">imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wiktorowicz, Sloane J.; Millar-Blanchaer, Max; Perrin, Marshall D.; Graham, James R.; Fitzgerald, Michael P.; Maire, Jérôme; Ingraham, Patrick; Savransky, Dmitry; Macintosh, Bruce A.; Thomas, Sandrine J.; Chilcote, Jeffrey K.; Draper, Zachary H.; Song, Inseok; Cardwell, Andrew; Goodsell, Stephen J.; Hartung, Markus; Hibon, Pascale; Rantakyrö, Fredrik; Sadakuni, Naru</p> <p>2014-07-01</p> <p>We present on-sky polarimetric <span class="hlt">observations</span> with the Gemini Planet <span class="hlt">Imager</span> (GPI) obtained at straight Cassegrain focus on the Gemini South 8-m telescope. <span class="hlt">Observations</span> of polarimetric calibrator stars, ranging from nearly un- polarized to strongly polarized, enable determination of the combined telescope and instrumental polarization. We find the conversion of Stokes I to linear and circular instrumental polarization in the instrument frame to be I --> (QIP, UIP, PIP, VIP) = (-0.037 +/- 0.010%, +0.4338 +/- 0.0075%, 0.4354 +/- 0.0075%, -6.64 +/- 0.56%). Such precise measurement of instrumental polarization enables ~0.1% absolute accuracy in measurements of linear polarization, which together with GPI's high contrast will allow GPI to explore scattered light from circumstellar disk in unprecedented detail, conduct <span class="hlt">observations</span> of a range of other astronomical bodies, and potentially even study polarized thermal emission from young exoplanets. <span class="hlt">Observations</span> of unpolarized standard stars also let us quantify how well GPI's differential polarimetry mode can suppress the stellar PSF halo. We show that GPI polarimetry achieves cancellation of unpolarized starlight by factors of 100-200, reaching the photon noise limit for sensitivity to circumstellar scattered light for all but the smallest separations at which the calibration for instrumental polarization currently sets the limit.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ASInC...4..189O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ASInC...4..189O"><span>First light <span class="hlt">observations</span> with TIFR Near Infrared <span class="hlt">Imaging</span> Camera (TIRCAM-II)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ojha, D. K.; Ghosh, S. K.; D'Costa, S. L. A.; Naik, M. B.; Sandimani, P. R.; Poojary, S. S.; Bhagat, S. B.; Jadhav, R. B.; Meshram, G. S.; Bakalkar, C. B.; Ramaprakash, A. N.; Mohan, V.; Joshi, J.</p> <p></p> <p>TIFR near infrared <span class="hlt">imaging</span> camera (TIRCAM-II) is based on the Aladdin III Quadrant InSb focal plane array (512×512 pixels; 27.6 μm pixel size; sensitive between 1 - 5.5 μm). TIRCAM-II had its first engineering run with the 2 m IUCAA telescope at Girawali during February - March 2011. The first light <span class="hlt">observations</span> with TIRCAM-II were quite successful. Several infrared standard with TIRCAM-II were quite successful. Several infrared standard stars, the Trapezium Cluster in Orion region, McNeil's nebula, etc., were <span class="hlt">observed</span> in the J, K and in a narrow-band at 3.6 μm (nbL). In the nbL band, some bright stars could be detected from the Girawali site. The performance of TIRCAM-II is discussed in the light of preliminary <span class="hlt">observations</span> in near infrared bands.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1993JGR....9817183G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1993JGR....9817183G"><span>Galileo <span class="hlt">imaging</span> <span class="hlt">observations</span> of Lunar Maria and related deposits</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Greeley, Ronald; Kadel, Steven D.; Williams, David A.; Gaddis, Lisa R.; Head, James W.; McEwen, Alfred S.; Murchie, Scott L.; Nagel, Engelbert; Neukum, Gerhard; Pieters, Carle M.; Sunshine, Jessica M.; Wagner, Roland; Belton, Michael J. S.</p> <p></p> <p>The Galileo spacecraft <span class="hlt">imaged</span> parts of the western limb and far side of the Moon in December 1990. Ratios of 0.41/0.56 μm filter <span class="hlt">images</span> from the Solid State <span class="hlt">Imaging</span> (SSI) experiment provided information on the titanium content of mare deposits; ratios of the 0.76/0.99 μm <span class="hlt">images</span> indicated 1 μm absorptions associated with Fe2+ in mafic minerals. Mare ages were derived from crater statistics obtained from Lunar Orbiter <span class="hlt">images</span>. Results on mare compositions in western Oceanus Procellarum and the Humorum basin are consistent with previous Earth-based <span class="hlt">observations</span>, thus providing confidence in the use of Galileo data to extract compositional information. Mare units in the Grimaldi and Riccioli basins range in age from 3.25 to 3.48 Ga and consist of medium- to medium-high titanium (<4 to 7% TiO2) content lavas. The Schiller-Zucchius basin shows a higher 0.76/0.99 μm ratio than the surrounding highlands, indicating a potentially higher mafic mineral content consistent with previous interpretations that the area includes mare deposits blanketed by highland ejecta and light plains materials. The oldest mare materials in the Orientale basin occur in south-central Mare Orientale and are 3.7 Ga old; youngest mare materials are in Lacus Autumni and are 2.85 Ga old; these units are medium- to medium-high titanium (<4 to 7% TiO2) basalts. Thus, volcanism was active in Orientale for 0.85 Ga, but lavas were relatively constant in composition. Galileo data suggest that Mendel-Rydberg mare is similar to Mare Orientale; cryptomare are present as well. Thus, the mare lavas on the western limb and far side (to 178°E) are remarkably uniform in composition, being generally of medium- to medium-high titanium content and having relatively low 0.76/0.99 μm ratios. This region of the Moon is between two postulated large impact structures, the Procellarum and the South Pole-Aitken basins, and may have a relatively thick crust. In areas underlain by an inferred thinner crust, i.e., zones</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1324517-gemini-planet-imager-observations-au-microscopii-debris-disk-asymmetries-within-one-arcsecond','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1324517-gemini-planet-imager-observations-au-microscopii-debris-disk-asymmetries-within-one-arcsecond"><span>Gemini Planet <span class="hlt">Imager</span> <span class="hlt">observations</span> of the AU Microscopii debris disk: Asymmetries within one arcsecond</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Wang, Jason J.; Graham, James R.; Pueyo, Laurent; ...</p> <p>2015-09-23</p> <p>We present Gemini Planet <span class="hlt">Imager</span> (GPI) <span class="hlt">observations</span> of AU Microscopii, a young M dwarf with an edge-on, dusty debris disk. Integral field spectroscopy and broadband <span class="hlt">imaging</span> polarimetry were obtained during the commissioning of GPI. In our broadband <span class="hlt">imaging</span> polarimetry <span class="hlt">observations</span>, we detect the disk only in total intensity and find asymmetries in the morphology of the disk between the southeast (SE) and northwest (NW) sides. The SE side of the disk exhibits a bump at 1'' (10 AU projected separation) that is three times more vertically extended and three times fainter in peak surface brightness than the NW side atmore » similar separations. This part of the disk is also vertically offset by 69 ± 30 mas to the northeast at 1'' when compared to the established disk midplane and is consistent with prior Atacama Large Millimeter/submillimeter Array and Hubble Space Telescope/Space Telescope <span class="hlt">Imaging</span> Spectrograph <span class="hlt">observations</span>. We see hints that the SE bump might be a result of detecting a horizontal sliver feature above the main disk that could be the disk backside. Alternatively, when including the morphology of the NW side, where the disk midplane is offset in the opposite direction ~50 mas between 0farcs4 and 1farcs2, the asymmetries suggest a warp-like feature. Using our integral field spectroscopy data to search for planets, we are 50% complete for ~4 MJup planets at 4 AU. Lastly, we detect a source, resolved only along the disk plane, that could either be a candidate planetary mass companion or a compact clump in the disk.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SPIE.9463E..0CH','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SPIE.9463E..0CH"><span>Collaborative real-time motion video analysis by human <span class="hlt">observer</span> and <span class="hlt">image</span> exploitation algorithms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hild, Jutta; Krüger, Wolfgang; Brüstle, Stefan; Trantelle, Patrick; Unmüßig, Gabriel; Heinze, Norbert; Peinsipp-Byma, Elisabeth; Beyerer, Jürgen</p> <p>2015-05-01</p> <p>Motion video analysis is a challenging task, especially in real-time applications. In most safety and security critical applications, a human <span class="hlt">observer</span> is an obligatory part of the overall analysis system. Over the last years, substantial progress has been made in the development of automated <span class="hlt">image</span> exploitation algorithms. Hence, we investigate how the benefits of automated video analysis can be integrated suitably into the current video exploitation systems. In this paper, a system design is introduced which strives to combine both the qualities of the human <span class="hlt">observer</span>'s perception and the automated algorithms, thus aiming to improve the overall performance of a real-time video analysis system. The system design builds on prior work where we showed the benefits for the human <span class="hlt">observer</span> by means of a user interface which utilizes the human visual focus of attention revealed by the eye gaze direction for interaction with the <span class="hlt">image</span> exploitation system; eye tracker-based interaction allows much faster, more convenient, and equally precise moving target acquisition in video <span class="hlt">images</span> than traditional computer mouse selection. The system design also builds on prior work we did on automated target detection, segmentation, and tracking algorithms. Beside the system design, a first pilot study is presented, where we investigated how the participants (all non-experts in video analysis) performed in initializing an object tracking subsystem by selecting a target for tracking. Preliminary results show that the gaze + key press technique is an effective, efficient, and easy to use interaction technique when performing selection operations on moving targets in videos in order to initialize an object tracking function.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170003779&hterms=hodge&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dhodge','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170003779&hterms=hodge&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dhodge"><span>The HEXITEC Hard X-Ray Pixelated CdTe <span class="hlt">Imager</span> for Fast Solar <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>Baumgartner, Wayne H.; Christe, Steven D.; Ryan, Daniel; Inglis, Andrew R.; Shih, Albert Y.; Gregory, Kyle; Wilson, Matt; Seller, Paul; Gaskin, Jessica; Wilson-Hodge, Colleen</p> <p>2016-01-01</p> <p>There is an increasing demand in solar and astrophysics for high resolution X-ray spectroscopic <span class="hlt">imaging</span>. Such <span class="hlt">observations</span> would present ground breaking opportunities to study the poorly understood high energy processes in our solar system and beyond, such as solar flares, X-ray binaries, and active galactic nuclei. However, such <span class="hlt">observations</span> require a new breed of solid state detectors sensitive to high energy X-rays with fine independent pixels to sub-sample the point spread function (PSF) of the X-ray optics. For solar <span class="hlt">observations</span> in particular, they must also be capable of handling very high count rates as photon fluxes from solar flares often cause pile up and saturation in present generation detectors. The Rutherford Appleton Laboratory (RAL) has recently developed a new cadmium telluride (CdTe) detector system, called HEXITEC (High Energy X-ray <span class="hlt">Imaging</span> Technology). It is an 80 x 80 array of 250 micron independent pixels sensitive in the 2-200 keV band and capable of a high full frame read out rate of 10 kHz. HEXITEC provides the smallest independently read out CdTe pixels currently available, and are well matched to the few arcsecond PSF produced by current and next generation hard X-ray focusing optics. NASA's Goddard and Marshall Space Flight Centers are collaborating with RAL to develop these detectors for use on future space borne hard X-ray focusing telescopes. We show the latest results on HEXITEC's <span class="hlt">imaging</span> capability, energy resolution, high read out rate, and reveal it to be ideal for such future instruments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018LPICo2063.3139T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018LPICo2063.3139T"><span>Tunable Light-Guide <span class="hlt">Image</span> Processing Snapshot Spectrometer (TuLIPSS) for Earth and Moon <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>Tkaczyk, T. S.; Alexander, D.; Luvall, J. C.; Wang, Y.; Dwight, J. G.; Pawlowsk, M. E.; Howell, B.; Tatum, P. F.; Stoian, R.-I.; Cheng, S.; Daou, A.</p> <p>2018-02-01</p> <p>A tunable light-guide <span class="hlt">image</span> processing snapshot spectrometer (TuLIPSS) for Earth science research and <span class="hlt">observation</span> is being developed through a NASA instrument incubator project with Rice University and Marshall Space Flight Center.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22522186-molecular-ionized-hydrogen-doradus-imaging-observations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22522186-molecular-ionized-hydrogen-doradus-imaging-observations"><span>MOLECULAR AND IONIZED HYDROGEN IN 30 DORADUS. I. <span class="hlt">IMAGING</span> <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>Yeh, Sherry C. C.; Seaquist, Ernest R.; Matzner, Christopher D.</p> <p>2015-07-10</p> <p>We present the first fully calibrated H{sub 2} 1–0 S(1) <span class="hlt">image</span> of the entire 30 Doradus nebula. The <span class="hlt">observations</span> were conducted using the NOAO Extremely Wide-field Infrared <span class="hlt">Imager</span> (NEWFIRM) on the CTIO 4 m Blanco Telescope. Together with a NEWFIRM Brγ <span class="hlt">image</span> of 30 Doradus, our data reveal the morphologies of the warm molecular gas and ionized gas in 30 Doradus. The brightest H{sub 2}-emitting area, which extends from the northeast to the southwest of R136, is a photodissociation region (PDR) viewed face-on, while many clumps and pillar features located at the outer shells of 30 Doradus are PDRs viewedmore » edge-on. Based on the morphologies of H{sub 2}, Brγ, CO, and 8 μm emission, the H{sub 2} to Brγ line ratio, and Cloudy models, we find that the H{sub 2} emission is formed inside the PDRs of 30 Doradus, 2–3 pc to the ionization front of the H ii region, in a relatively low-density environment <10{sup 4} cm{sup −3}. Comparisons with Brγ, 8 μm, and CO emission indicate that H{sub 2} emission is due to fluorescence, and provide no evidence for shock excited emission of this line.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40831&hterms=oil+drilling&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Doil%2Bdrilling','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40831&hterms=oil+drilling&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Doil%2Bdrilling"><span>Earth <span class="hlt">observation</span> photo taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Earth <span class="hlt">observation</span> photo taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). <span class="hlt">Image</span> of California's coast from Point Concepcion (far left) to Ventura (right). The city of Santa Barbara is visible as a bright region (center). The row of bright spots in the ocean are oil drilling platforms in the Santa Barbara Channel, while the random points of brightness in the channel are vessels. Lakes Cachuma (left) and Casitas (right) are seen as large dark areas. Folded sedimentary rock layers are visible in the Santa Ynez Mountain Range which stretches down the coastline; the stratification terminates at the Santa Ynez fault on the island side of the mountains.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFMSA11A2130S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFMSA11A2130S"><span>MOOSE: A Multi-Spectral Observatory Of Sensitive EMCCDs for innovative research in space physics and aeronomy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Samara, M.; Michell, R. G.; Hampton, D. L.; Trondsen, T.</p> <p>2012-12-01</p> <p>The Multi-Spectral Observatory Of Sensitive EMCCDs (MOOSE) consists of 5 <span class="hlt">imaging</span> systems and is the result of an NSF-funded Major Research Instrumentation project. The main objective of MOOSE is to provide a resource to all members of the scientific community that have interests in <span class="hlt">imaging</span> low-light-level phenomena, such as aurora, <span class="hlt">airglow</span>, and meteors. Each <span class="hlt">imager</span> consists of an Andor DU-888 Electron Multiplying CCD (EMCCD), combined with a telecentric optics section, made by Keo Scientific Ltd., with a selection of available angular fields of view. During the northern hemisphere winter the system is typically based and operated at Poker Flat Research Range in Alaska, but any or all <span class="hlt">imagers</span> can be shipped anywhere in individual stand-alone cases. We will discuss the main components of the MOOSE project, including the <span class="hlt">imagers</span>, optics, lenses and filters, as well as the Linux-based control software that enables remote operation. We will also discuss the calibration of the <span class="hlt">imagers</span> along with the initial deployments and testing done. We are requesting community input regarding operational modes, such as filter and field of view combinations, frame rates, and potentially moving some <span class="hlt">imagers</span> to other locations, either for tomography or for larger spatial coverage. In addition, given the large volume of auroral <span class="hlt">image</span> data already available, we are encouraging collaborations for which we will freely distribute the data and any analysis tools already developed. Most significantly, initial science highlights relating to aurora, <span class="hlt">airglow</span> and meteors will be discussed in the context of the creative and innovative ways that the MOOSE observatory can be used in order to address a new realm of science topics, previously unachievable with traditional single <span class="hlt">imager</span> systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JBO....20j7001M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JBO....20j7001M"><span>Nondestructive <span class="hlt">observation</span> of teeth post core-space using optical coherence tomography: comparison with microcomputed tomography and live <span class="hlt">images</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Minamino, Takuya; Mine, Atsushi; Matsumoto, Mariko; Sugawa, Yoshihiko; Kabetani, Tomoshige; Higashi, Mami; Kawaguchi, Asuka; Ohmi, Masato; Awazu, Kunio; Yatani, Hirofumi</p> <p>2015-10-01</p> <p>No previous reports have <span class="hlt">observed</span> inside the root canal using both optical coherence tomography (OCT) and x-ray microcomputed tomography (μCT) for the same sample. The purpose of this study was to clarify both OCT and μCT <span class="hlt">image</span> properties from <span class="hlt">observations</span> of the same root canal after resin core build-up treatment. As OCT allows real-time <span class="hlt">observation</span> of samples, gap formation may be able to be shown in real time. A dual-cure, one-step, self-etch adhesive system bonding agent, and dual-cure resin composite core material were used in root canals in accordance with instructions from the manufacturer. The resulting OCT <span class="hlt">images</span> were superior for identifying gap formation at the interface, while μCT <span class="hlt">images</span> were better to grasp the tooth form. Continuous tomographic <span class="hlt">images</span> from real-time OCT <span class="hlt">observation</span> allowed successful construction of a video of the resin core build-up procedure. After 10 to 12 s of light curing, a gap with a clear new signal occurred at the root-core material interface, proceeding from the coronal side (6 mm from the cemento-enamel junction) to the apical side of the root.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SPIE.9416E..06B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SPIE.9416E..06B"><span>Computational assessment of mammography accreditation phantom <span class="hlt">images</span> and correlation with human <span class="hlt">observer</span> analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Barufaldi, Bruno; Lau, Kristen C.; Schiabel, Homero; Maidment, D. A.</p> <p>2015-03-01</p> <p>Routine performance of basic test procedures and dose measurements are essential for assuring high quality of mammograms. International guidelines recommend that breast care providers ascertain that mammography systems produce a constant high quality <span class="hlt">image</span>, using as low a radiation dose as is reasonably achievable. The main purpose of this research is to develop a framework to monitor radiation dose and <span class="hlt">image</span> quality in a mixed breast screening and diagnostic <span class="hlt">imaging</span> environment using an automated tracking system. This study presents a module of this framework, consisting of a computerized system to measure the <span class="hlt">image</span> quality of the American College of Radiology mammography accreditation phantom. The methods developed combine correlation approaches, matched filters, and data mining techniques. These methods have been used to analyze radiological <span class="hlt">images</span> of the accreditation phantom. The classification of structures of interest is based upon reports produced by four trained readers. As previously reported, human <span class="hlt">observers</span> demonstrate great variation in their analysis due to the subjectivity of human visual inspection. The software tool was trained with three sets of 60 phantom <span class="hlt">images</span> in order to generate decision trees using the software WEKA (Waikato Environment for Knowledge Analysis). When tested with 240 <span class="hlt">images</span> during the classification step, the tool correctly classified 88%, 99%, and 98%, of fibers, speck groups and masses, respectively. The variation between the computer classification and human reading was comparable to the variation between human readers. This computerized system not only automates the quality control procedure in mammography, but also decreases the subjectivity in the expert evaluation of the phantom <span class="hlt">images</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10567E..2QP','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10567E..2QP"><span>Variable optical filters for earth-<span class="hlt">observation</span> <span class="hlt">imaging</span> minispectrometers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Piegari, A.; Bulir, J.; Krasilnikova, A.; Dami, M.; Harnisch, B.</p> <p>2017-11-01</p> <p>Small-dimension, low-mass spectrometers are useful for both Earth <span class="hlt">observation</span> and planetary missions. A very compact multi-spectral mini-spectrometer that contains no moving parts, can be constructed combining a graded-thickness filter, having a spatially variable narrow-band transmission, to a CCD array detector. The peak wavelength of the transmission filter is moving along one direction of the filter surface, such that each line of a two-dimensional array detector, equipped with this filter, will detect radiation in a different pass band. The spectrum of interest for <span class="hlt">image</span> spectrometry of the Earth surface is very wide, 400-1000nm. This requirement along with the need of a very small dimension, makes this filter very difficult to manufacture. Preliminary results on metal-dielectric wedge filters, with a gradient of the transmission peak wavelength equal to 60nm/mm, are reported.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SPIE10577E..0NH','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SPIE10577E..0NH"><span>Lesion detection performance of cone beam CT <span class="hlt">images</span> with anatomical background noise: single-slice vs. multi-slice human and model <span class="hlt">observer</span> study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Han, Minah; Jang, Hanjoo; Baek, Jongduk</p> <p>2018-03-01</p> <p>We investigate lesion detectability and its trends for different noise structures in single-slice and multislice CBCT <span class="hlt">images</span> with anatomical background noise. Anatomical background noise is modeled using a power law spectrum of breast anatomy. Spherical signal with a 2 mm diameter is used for modeling a lesion. CT projection data are acquired by the forward projection and reconstructed by the Feldkamp-Davis-Kress algorithm. To generate different noise structures, two types of reconstruction filters (Hanning and Ram-Lak weighted ramp filters) are used in the reconstruction, and the transverse and longitudinal planes of reconstructed volume are used for detectability evaluation. To evaluate single-slice <span class="hlt">images</span>, the central slice, which contains the maximum signal energy, is used. To evaluate multislice <span class="hlt">images</span>, central nine slices are used. Detectability is evaluated using human and model <span class="hlt">observer</span> studies. For model <span class="hlt">observer</span>, channelized Hotelling <span class="hlt">observer</span> (CHO) with dense difference-of-Gaussian (D-DOG) channels are used. For all noise structures, detectability by a human <span class="hlt">observer</span> is higher for multislice <span class="hlt">images</span> than single-slice <span class="hlt">images</span>, and the degree of detectability increase in multislice <span class="hlt">images</span> depends on the noise structure. Variation in detectability for different noise structures is reduced in multislice <span class="hlt">images</span>, but detectability trends are not much different between single-slice and multislice <span class="hlt">images</span>. The CHO with D-DOG channels predicts detectability by a human <span class="hlt">observer</span> well for both single-slice and multislice <span class="hlt">images</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013EGUGA..15.1951A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013EGUGA..15.1951A"><span>Initial <span class="hlt">Observations</span> and Activities of Curiosity's Mars Hand Lens <span class="hlt">Imager</span> (MAHLI) at the Gale Field Site</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aileen Yingst, R.; Edgett, Kenneth; MSL Science Team</p> <p>2013-04-01</p> <p> the dust and sand obscuration, the <span class="hlt">observables</span> are unclear —grains 300-500 µm size in the Bathurst Inlet <span class="hlt">images</span> and 300-500 µm-sized rhombus-shaped crystals in the rock, Jake Matijevic have been <span class="hlt">observed</span> by some workers. Sand and granules (as well as dust), exhibiting a variety of colors, shapes, and other grain attributes, were deposited on rover hardware during descent. As noted above, sand as well as dust also mantles the rocks <span class="hlt">observed</span> by MAHLI; in one case the cohesive properties of this material was demonstrated by the presence of a "micro landslide" on a rock named Burwash. At the Rocknest sand shadow, a variety of coarse to very coarse sand grains of differing color, shape, luster, angularity, and roundness were <span class="hlt">observed</span>, including glassy spheroids and ellipsoids (perhaps formed from impact melt droplets) and clear, translucent grains. The fine to very fine sands sieved (≤ 150 µm) and delivered to the rover's <span class="hlt">observation</span> tray exhibited at least four distinct grain types, including clear, translucent crystal fragments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180001849','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180001849"><span>The Latest on the Venus Thermospheric General Circulation Model: Capabilities and Simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brecht, A. S.; Bougher, S. W.; Parkinson, C. D.</p> <p>2017-01-01</p> <p>Venus has a complex and dynamic upper atmosphere. This has been <span class="hlt">observed</span> many times by ground-based, orbiters, probes, and fly-by missions going to other planets. Two over-arching questions are generally asked when examining the Venus upper atmosphere: (1) what creates the complex structure in the atmosphere, and (2) what drives the varying dynamics. A great way to interpret and connect <span class="hlt">observations</span> to address these questions utilizes numerical modeling; and in the case of the middle and upper atmosphere (above the cloud tops), a 3D hydrodynamic numerical model called the Venus Thermospheric General Circulation Model (VTGCM) can be used. The VTGCM can produce climatological averages of key features in comparison to <span class="hlt">observations</span> (i.e. nightside temperature, O2 IR nightglow emission). More recently, the VTGCM has been expanded to include new chemical constituents and <span class="hlt">airglow</span> emissions, as well as new parameterizations to address waves and their impact on the varying global circulation and corresponding <span class="hlt">airglow</span> distributions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990013971','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990013971"><span>A Novel, Poly-Etalon, Fabry-Perot for Planetary Research</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kerr, Robert B.; Doe, Richard; Noto, John</p> <p>1997-01-01</p> <p>In an effort to develop a mechanically robust, high throughput and solid state spectrometer several liquid crystal Fabry-Perot etalons were constructed. The etalons were tested for spectral response, radiation resistance and optical transmission. The first year of this project was spent developing and understanding the properties of the liquid crystal etalons; in the second year an intensified all-sky <span class="hlt">imaging</span> system was developed around a pair of LC etalons. The <span class="hlt">imaging</span> system, developed jointly with SRI International represents a unique brassboard to demonstrate the use of LC etalons as tunable filters. The first set of etalons constructed in year one of this project were tested for spectral response and throughput while etalon surrogates were exposed to proton radiation simulating the exposure of an object in Low Earth Orbit (LEO). The 2" diameter etalons had a measure finesse of approximately 10 and were tunable over five orders. Liquid crystals exposed to proton irradiation showed no signs of damage. In year two two larger diameter (3") etalons were constructed with gaps of 3 and 5 microns. This pair of etalons is for use in a high resolution, all-sky spectral <span class="hlt">imager</span>. The WATUMI <span class="hlt">imager</span> system follows the heritage of all sky, narrow band, intensified <span class="hlt">imagers</span> however it includes two LC Fabry-Perot etalons to provide tunability and the ability to switch wavelengths rapidly, an import consideration in auroral <span class="hlt">airglow</span> <span class="hlt">imaging</span>. This work also resulted in two publications and one poster presentation. The instrument will be uniquely capable, with superior throughput and speed, to measure optical <span class="hlt">airglow</span> of multiple emission lines in harsh conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22685142-clinically-observed-discrepancy-between-image-based-log-based-mlc-positions','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22685142-clinically-observed-discrepancy-between-image-based-log-based-mlc-positions"><span>A clinically <span class="hlt">observed</span> discrepancy between <span class="hlt">image</span>-based and log-based MLC positions</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>Neal, Brian, E-mail: bpn2p@virginia.edu; Ahmed, Mahmoud; Kathuria, Kunal</p> <p>2016-06-15</p> <p>Purpose: To present a clinical case in which real-time intratreatment <span class="hlt">imaging</span> identified an multileaf collimator (MLC) leaf to be consistently deviating from its programmed and logged position by >1 mm. Methods: An EPID-based exit-fluence dosimetry system designed to prevent gross delivery errors was used to capture cine during treatment <span class="hlt">images</span>. The author serendipitously visually identified a suspected MLC leaf displacement that was not otherwise detected. The leaf position as recorded on the EPID <span class="hlt">images</span> was measured and log-files were analyzed for the treatment in question, the prior day’s treatment, and for daily MLC test patterns acquired on those treatment days.more » Additional standard test patterns were used to quantify the leaf position. Results: Whereas the log-file reported no difference between planned and recorded positions, <span class="hlt">image</span>-based measurements showed the leaf to be 1.3 ± 0.1 mm medial from the planned position. This offset was confirmed with the test pattern irradiations. Conclusions: It has been clinically <span class="hlt">observed</span> that log-file derived leaf positions can differ from their actual position by >1 mm, and therefore cannot be considered to be the actual leaf positions. This cautions the use of log-based methods for MLC or patient quality assurance without independent confirmation of log integrity. Frequent verification of MLC positions through independent means is a necessary precondition to trust log-file records. Intratreatment EPID <span class="hlt">imaging</span> provides a method to capture departures from MLC planned positions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFMSH33A2035S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFMSH33A2035S"><span>Radio <span class="hlt">Imaging</span> <span class="hlt">Observations</span> of Solar Activity Cycle and Its Anomaly</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shibasaki, K.</p> <p>2011-12-01</p> <p>The 24th solar activity cycle has started and relative sunspot numbers are increasing. However, their rate of increase is rather slow compared to previous cycles. Active region sizes are small, lifetime is short, and big (X-class) flares are rare so far. We study this anomalous situation using data from Nobeyama Radioheliograph (NoRH). Radio <span class="hlt">imaging</span> <span class="hlt">observations</span> have been done by NoRH since 1992. Nearly 20 years of daily radio <span class="hlt">images</span> of the Sun at 17 GHz are used to synthesize a radio butterfly diagram. Due to stable operation of the instrument and a robust calibration method, uniform datasets are available covering the whole period of <span class="hlt">observation</span>. The radio butterfly diagram shows bright features corresponding to active region belts and their migration toward low latitude as the solar cycle progresses. In the present solar activity cycle (24), increase of radio brightness is delayed and slow. There are also bright features around both poles (polar brightening). Their brightness show solar cycle dependence but peaks around solar minimum. Comparison between the last minimum and the previous one shows decrease of its brightness. This corresponds to weakening of polar magnetic field activity between them. In the northern pole, polar brightening is already weakened in 2011, which means it is close to solar maximum in the northern hemisphere. Southern pole does not show such feature yet. Slow rise of activity in active region belt, weakening of polar activity during the minimum, and large north-south asymmetry in polar activity imply that global solar activity and its synchronization are weakening.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014acm..conf..321L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014acm..conf..321L"><span>Asteroid (4179) Toutatis size determination via optical <span class="hlt">images</span> <span class="hlt">observed</span> by the Chang'e-2 probe</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, P.; Huang, J.; Zhao, W.; Wang, X.; Meng, L.; Tang, X.</p> <p>2014-07-01</p> <p>This work is a physical and statistical study of the asteroid (4179) Toutatis using the optical <span class="hlt">images</span> obtained by a solar panel monitor of the Chang'e-2 probe on Dec. 13, 2012 [1]. In the <span class="hlt">imaging</span> strategy, the camera is focused at infinity. This is specially designed for the probe with its solar panels monitor's principle axis pointing to the relative velocity direction of the probe and Toutatis. The <span class="hlt">imaging</span> strategy provides a dedicated way to resolve the size by multi-frame optical <span class="hlt">images</span>. The inherent features of the data are: (1) almost no rotation was recorded because of the 5.41-7.35 Earth-day rotation period and the small amount of elapsed <span class="hlt">imaging</span> time, only minutes, make the object stay in the <span class="hlt">images</span> in a fixed position and orientation; (2) the sharpness of the upper left boundary and the vagueness of lower right boundary resulting from the direction of SAP (Sun-Asteroid-Probe angle) cause a varying accuracy in locating points at different parts of Toutatis. A common view is that direct, accurate measurements of asteroid shapes, sizes, and pole positions are now possible for larger asteroids that can be spatially resolved using the Hubble Space Telescope or large ground-based telescopes equipped with adaptive optics. For a quite complex planetary/asteroid probe study, these measurements certainly need continuous validation via a variety of ways [2]. Based on engineering parameters of the probe during the fly-by, the target spatial resolving and measuring procedures are described in the paper. Results estimated are optical perceptible size on the flyby epoch under the solar phase angles during the <span class="hlt">imaging</span>. It is found that the perceptible size measured using the optical <span class="hlt">observations</span> and the size derived from the radar <span class="hlt">observations</span> by Ostro et al.~in 1995 [3], are close to one another.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26158086','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26158086"><span>Implementation of a channelized Hotelling <span class="hlt">observer</span> model to assess <span class="hlt">image</span> quality of x-ray angiography systems.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Favazza, Christopher P; Fetterly, Kenneth A; Hangiandreou, Nicholas J; Leng, Shuai; Schueler, Beth A</p> <p>2015-01-01</p> <p>Evaluation of flat-panel angiography equipment through conventional <span class="hlt">image</span> quality metrics is limited by the scope of standard spatial-domain <span class="hlt">image</span> quality metric(s), such as contrast-to-noise ratio and spatial resolution, or by restricted access to appropriate data to calculate Fourier domain measurements, such as modulation transfer function, noise power spectrum, and detective quantum efficiency. <span class="hlt">Observer</span> models have been shown capable of overcoming these limitations and are able to comprehensively evaluate medical-<span class="hlt">imaging</span> systems. We present a spatial domain-based channelized Hotelling <span class="hlt">observer</span> model to calculate the detectability index (DI) of our different sized disks and compare the performance of different <span class="hlt">imaging</span> conditions and angiography systems. When appropriate, changes in DIs were compared to expectations based on the classical Rose model of signal detection to assess linearity of the model with quantum signal-to-noise ratio (SNR) theory. For these experiments, the estimated uncertainty of the DIs was less than 3%, allowing for precise comparison of <span class="hlt">imaging</span> systems or conditions. For most experimental variables, DI changes were linear with expectations based on quantum SNR theory. DIs calculated for the smallest objects demonstrated nonlinearity with quantum SNR theory due to system blur. Two angiography systems with different detector element sizes were shown to perform similarly across the majority of the detection tasks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20170007520&hterms=colors&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcolors','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20170007520&hterms=colors&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcolors"><span>An Investigation of Dust Storms <span class="hlt">Observed</span> with the Mars Color <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Guzewich, Scott D.; Toigo, Anthony D.; Wang, Huiqun</p> <p>2017-01-01</p> <p>Daily global <span class="hlt">imaging</span> by the Mars Color <span class="hlt">Imager</span> (MARCI) continues the record of the Mars Orbiter Camera (MOC) and has allowed creation of a long-duration record of Martian dust storms. We <span class="hlt">observe</span> dust storms over the first two Mars years of the MARCI record, including tracking individual storms over multiple sols, as well as tracking the growth and recession of the seasonal polar caps. Using the combined 6 Mars year record of textured dust storms (storms with visible textures on the <span class="hlt">observed</span> dust cloud tops), we study the relationship between textured dust storm activity and meteorology (as simulated by the MarsWRF general circulation model) and surface properties. We find that textured dust storms preferentially occur in places and seasons with above average surface wind stress. Textured dust storm occurrence also has a modest linear anti-correlation with surface albedo (0.43) and topography (0.40). Lastly, we perform an empirical orthogonal function (EOF) analysis on the distribution of occurrence of textured dust storms and find that over 50 of the variance in textured dust storm activity can be explained by two EOF modes. We associate the first EOF mode with cap-edge storms just before Ls = 180deg and the second EOF mode with flushing dust storms that occur from Ls = 180-210deg and again near Ls = 320deg.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AMT....11..473Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AMT....11..473Z"><span>Retrieval of O2(1Σ) and O2(1Δ) volume emission rates in the mesosphere and lower thermosphere using SCIAMACHY MLT limb scans</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zarboo, Amirmahdi; Bender, Stefan; Burrows, John P.; Orphal, Johannes; Sinnhuber, Miriam</p> <p>2018-01-01</p> <p>We present the retrieved volume emission rates (VERs) from the <span class="hlt">airglow</span> of both the daytime and twilight O2(1Σ) band and O2(1Δ) band emissions in the mesosphere and lower thermosphere (MLT). The SCanning <span class="hlt">Imaging</span> Absorption SpectroMeter for Atmospheric CHartographY (SCIAMACHY) onboard the European Space Agency Envisat satellite <span class="hlt">observes</span> upwelling radiances in limb-viewing geometry during its special MLT mode over the range 50-150 km. In this study we use the limb <span class="hlt">observations</span> in the visible (595-811 nm) and near-infrared (1200-1360 nm) bands. We have investigated the daily mean latitudinal distributions and the time series of the retrieved VER in the altitude range from 53 to 149 km. The maximal <span class="hlt">observed</span> VERs of O2(1Δ) during daytime are typically 1 to 2 orders of magnitude larger than those of O2(1Σ). The latter peaks at around 90 km, whereas the O2(1Δ) emissivity decreases with altitude, with the largest values at the lower edge of the <span class="hlt">observations</span> (about 53 km). The VER values in the upper mesosphere (above 80 km) are found to depend on the position of the sun, with pronounced high values occurring during summer for O2(1Δ). O2(1Σ) emissions show additional high values at polar latitudes during winter and spring. These additional high values are presumably related to the downwelling of atomic oxygen after large sudden stratospheric warmings (SSWs). Accurate measurements of the O2(1Σ) and O2(1Δ) <span class="hlt">airglow</span>, provided that the mechanism of their production is understood, yield valuable information about both the chemistry and dynamics in the MLT. For example, they can be used to infer the amounts and distribution of ozone, solar heating rates, and temperature in the MLT.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JKPS...65..786K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JKPS...65..786K"><span>A comparison of FUV dayglows measured by STSAT-1/FIMS with the AURIC model in a geomagnetic quiet condition</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kam, Hosik; Kim, Yong Ha; Hong, Jun-Seok; Lee, Joon-Chan; Choi, Yeon-Ju; Min, Kyung Wook</p> <p>2014-09-01</p> <p>The Korea scientific microsatellite, STSAT-1 (Science and Technology Satellite-1), was launched in 2003 and <span class="hlt">observed</span> far ultraviolet (FUV) <span class="hlt">airglow</span> from the upper atmosphere with a Far-ultraviolet <span class="hlt">IMaging</span> Spectrograph (FIMS) at an altitude of 690 km. The FIMS consists of a dual-band <span class="hlt">imaging</span> spectrograph of 900-1150 Å (S-band) and 1340-1715 Å (L-band). Limb scanning <span class="hlt">observations</span> were performed only at the S-band, resulting in intensity profiles of OI 989 Å, OI 1026 Å, NII 1085 Å and NI 1134 Å emission lines near the horizon. We compare these emission intensities with those computed by using a theoretical model, the AURIC (Atmospheric Ultraviolet Radiance Integrated Code). The intensities of the OI 1026 Å, NII 1085 Å and NI 1134 Å emissions measured by using the FIMS are overall consistent with the values computed by using AURIC under the thermospheric and solar activity conditions on August 6, 1984, which is close to the FIMS's <span class="hlt">observation</span> condition. We find that the FIMS dayglow intensity profiles match reasonably well with AURIC intensity profiles for the MSIS90 oxygen atom density profiles within factors of 0.5 and 2. However, the FIMS intensities of the OI 989 Å line are about 2 ˜ 4 times stronger than the AURIC intensities, which is expected because AURIC does not properly simulate resonance scattering of <span class="hlt">airglow</span> and solar photons at 989 Å by atomic oxygen in the thermosphere. We also find that the maximum tangential altitudes of the oxygen bearing dayglows (OI 989 Å, OI 1026 Å) are higher than those of the nitrogen-bearing dayglows (NII 1085 Å, NI 1134 Å), which is confirmed by using AURIC model calculations. This is expected because the oxygen atoms are distributed at higher altitudes in the thermosphere than the nitrogen molecules. Validations of the qualities of both the FIMS instrument and the AURIC model indicate that AURIC should be updated with improved thermospheric models and with measured solar FUV spectra for better agreement with the</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/2010JGRE..115.9002V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JGRE..115.9002V"><span><span class="hlt">Observations</span> from the High Resolution <span class="hlt">Imaging</span> Science Experiment (HiRISE): Martian dust devils in Gusev and Russell craters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Verba, Circe A.; Geissler, Paul E.; Titus, Timothy N.; Waller, Devin</p> <p>2010-09-01</p> <p>Two areas targeted for repeated <span class="hlt">imaging</span> by detailed High Resolution <span class="hlt">Imaging</span> Science Experiment (HiRISE) <span class="hlt">observations</span> allow us to examine morphological differences and monitor seasonal variations of Martian dust devil tracks at two quite different locations. Russell crater (53.3°S, 12.9°E) is regularly <span class="hlt">imaged</span> to study seasonal processes including deposition and sublimation of CO2 frost. Gusev crater (14.6°S, 175.4°E) has been frequently <span class="hlt">imaged</span> in support of the Mars Exploration Rover mission. Gusev crater provides the first opportunity to compare “ground truth” orbital <span class="hlt">observations</span> of dust devil tracks to surface <span class="hlt">observations</span> of active dust plumes. Orbital <span class="hlt">observations</span> show that dust devil tracks are rare, forming at a rate <1/110 that of the occurrence of active dust plumes estimated from Spirit's surface <span class="hlt">observations</span>. Furthermore, the tracks <span class="hlt">observed</span> from orbit are wider than typical plume diameters <span class="hlt">observed</span> by Spirit. We conclude that the tracks in Gusev are primarily formed by rare, large dust devils. Smaller dust devils fail to leave tracks that are visible from orbit, perhaps because of limited surface excavation depths. Russell crater displays more frequent, smaller sinuous tracks than Gusev. This may be due to the thin dust cover in Russell, allowing smaller dust devils to penetrate through the bright dust layer and leave conspicuous tracks. The start of the dust devil season and peak activity are delayed in Russell in comparison to Gusev, likely because of its more southerly location. Dust devils in both sites travel in directions consistent with general circulation model (GCM)-predicted winds, confirming a laboratory-derived approach to determining dust devil travel directions based on track morphology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9143E..2BZ','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9143E..2BZ"><span>BATMAN flies: a compact spectro-<span class="hlt">imager</span> for space <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zamkotsian, Frederic; Ilbert, Olivier; Zoubian, Julien; Delsanti, Audrey; Boissier, Samuel; Lancon, Ariane</p> <p>2014-08-01</p> <p>BATMAN flies is a compact spectro-<span class="hlt">imager</span> based on MOEMS for generating reconfigurable slit masks, and feeding two arms in parallel. The FOV is 25 x 12 arcmin2 for a 1m telescope, in infrared (0.85-1.7μm) and 500-1000 spectral resolution. Unique science cases for Space <span class="hlt">Observation</span> are reachable with this deep spectroscopic multi-survey instrument: deep survey of high-z galaxies down to H=25 on 5 deg2 with continuum detection and all z>7 candidates at H=26.2 over 5 deg2; deep survey of young stellar clusters in nearby galaxies; deep survey of the Kuiper Belt of ALL known objects down to H=22. Pathfinder towards BATMAN in space is already running with ground-based demonstrators.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970026745','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970026745"><span>Nocturnal <span class="hlt">Observations</span> of the Semidiurnal Tide at a Midlatitude Site</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Niciejewski, R. J.; Killeen, T. L.</p> <p>1995-01-01</p> <p>Fabry-Perot interferometer <span class="hlt">observations</span> of the mesospheric hydroxyl emission and the lower thermospheric OI (5577A) emission have been conducted from an <span class="hlt">airglow</span> observatory at a dark field site in southeastern Michigan for the past several years. The primary functions of the observatory are to provide a database for correlative <span class="hlt">observations</span> with the UARS satellite and to provide a synoptic measurement program for the coupling energetics and dynamics of atmospheric regions effort, An intensive operational effort between May 1993 and July 1994 has resulted in a substantial data set from which neutral winds have been determined from the bifilter acquisition sequence. A 'best fit' analysis in the least squares sense of the simultaneous measurements of the neutral winds to a 12-hour periodicity has provided amplitude and phase parameters for the semidiurnal tide as well as a measure of the mean wind. The measured tidal amplitude is greater at the higher altitude, though the seasonal behavior at both altitudes is similar with greater amplitudes during August/September and April/May. Both meridional and zonal wind components are consistent with a semidiurnal tidal description during the entire <span class="hlt">observational</span> sequence except for the May to July 1993 period. The mean winds show annual variation in the meridional flow, being equatorward from May to October and poleward during the winter. The zonal flow is primarily eastward during the entire <span class="hlt">observational</span> window with higher speed flows during May/June at the higher attitude and June/July at the lower altitude. A comparison with a semidiurnal tidal model indicates that the measured tidal amplitudes are a factor of 2 times greater, while the phases show similar equinoctial transitions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OcDyn..67..875D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OcDyn..67..875D"><span>Exploring <span class="hlt">image</span> data assimilation in the prospect of high-resolution satellite oceanic <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>Durán Moro, Marina; Brankart, Jean-Michel; Brasseur, Pierre; Verron, Jacques</p> <p>2017-07-01</p> <p>Satellite sensors increasingly provide high-resolution (HR) <span class="hlt">observations</span> of the ocean. They supply <span class="hlt">observations</span> of sea surface height (SSH) and of tracers of the dynamics such as sea surface salinity (SSS) and sea surface temperature (SST). In particular, the Surface Water Ocean Topography (SWOT) mission will provide measurements of the surface ocean topography at very high-resolution (HR) delivering unprecedented information on the meso-scale and submeso-scale dynamics. This study investigates the feasibility to use these measurements to reconstruct meso-scale features simulated by numerical models, in particular on the vertical dimension. A methodology to reconstruct three-dimensional (3D) multivariate meso-scale scenes is developed by using a HR numerical model of the Solomon Sea region. An inverse problem is defined in the framework of a twin experiment where synthetic <span class="hlt">observations</span> are used. A true state is chosen among the 3D multivariate states which is considered as a reference state. In order to correct a first guess of this true state, a two-step analysis is carried out. A probability distribution of the first guess is defined and updated at each step of the analysis: (i) the first step applies the analysis scheme of a reduced-order Kalman filter to update the first guess probability distribution using SSH <span class="hlt">observation</span>; (ii) the second step minimizes a cost function using <span class="hlt">observations</span> of HR <span class="hlt">image</span> structure and a new probability distribution is estimated. The analysis is extended to the vertical dimension using 3D multivariate empirical orthogonal functions (EOFs) and the probabilistic approach allows the update of the probability distribution through the two-step analysis. Experiments show that the proposed technique succeeds in correcting a multivariate state using meso-scale and submeso-scale information contained in HR SSH and <span class="hlt">image</span> structure <span class="hlt">observations</span>. It also demonstrates how the surface information can be used to reconstruct the ocean state below</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840049457&hterms=astronomia+espacio&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dastronomia%2By%2Bespacio','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840049457&hterms=astronomia+espacio&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dastronomia%2By%2Bespacio"><span>Microwave, soft and hard X-ray <span class="hlt">imaging</span> <span class="hlt">observations</span> of two solar flares</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.; Erskine, F. T.; Schmahl, E. J.; Machado, M. E.; Rovira, M. G.</p> <p>1984-01-01</p> <p>A set of microwave and hard X-ray <span class="hlt">observations</span> of two flares <span class="hlt">observed</span> simultaneously with the Very Large Array (VLA) and the Solar Maximum Mission Hard X-ray <span class="hlt">Imaging</span> Spectrometer (SMM-HXIS) are presented. The LVA was used at 6 cm to map the slowly varying and burst components in three neighboring solar active regions (Boulder Nos. 2522, 2530, and 2519) from approximately 14:00 UT until 01:00 UT on June 24-25, 1980. Six microwave bursts less than 30 sfu were <span class="hlt">observed</span>, and for the strongest of these, two-dimensional 'snapshot' (10 s) maps with spatial resolution of 5 in. were synthesized. HXIS data show clear interconnections between regions 2522 and 2530. The X-ray <span class="hlt">observations</span> present a global picture of flaring activity, while the VLA data show the complexity of the small magnetic structures associated with the impulsive phase phenomena. It is seen that energy release did not occur in a single isolated magnetic structure, but over a large area of intermingled loop structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=GL-2002-001170&hterms=cosmology&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dcosmology','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=GL-2002-001170&hterms=cosmology&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dcosmology"><span>False-color <span class="hlt">images</span> from <span class="hlt">observations</span> by the Supernova Cosmology Project of one of the two most dista</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2002-01-01</p> <p>TFalse-color <span class="hlt">images</span> from <span class="hlt">observations</span> by the Supernova Cosmology Project of one of the two most distant spectroscopically confirmed supernova. From the left: the first two <span class="hlt">images</span>, from the Cerro Tololo Interamerican Observatory 4-meter telescope, show a small region of sky just before and just after the the appearance of a type-Ia supernova that exploded when the universe was about half its present age. The third <span class="hlt">image</span> shows the same supernova as <span class="hlt">observed</span> with the Hubble Space Telescope. This much sharper picture allows a much better measurement of the apparent brightness and hence the distance of this supernova. Because their intrinsic brightness is predictable, such supernovae help to determine the deceleration, and so the eventual fate, of the universe. Credit: Perlmutter et al., The Supernova Cosmology Project</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10136E..0MB','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10136E..0MB"><span>Signal template generation from acquired mammographic <span class="hlt">images</span> for the non-prewhitening model <span class="hlt">observer</span> with eye-filter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Balta, Christiana; Bouwman, Ramona W.; Sechopoulos, Ioannis; Broeders, Mireille J. M.; Karssemeijer, Nico; van Engen, Ruben E.; Veldkamp, Wouter J. H.</p> <p>2017-03-01</p> <p>Model <span class="hlt">observers</span> (MOs) are being investigated for <span class="hlt">image</span> quality assessment in full-field digital mammography (FFDM). Signal templates for the non-prewhitening MO with eye filter (NPWE) were formed using acquired FFDM <span class="hlt">images</span>. A signal template was generated from acquired <span class="hlt">images</span> by averaging multiple exposures resulting in a low noise signal template. Noise elimination while preserving the signal was investigated and a methodology which results in a noise-free template is proposed. In order to deal with signal location uncertainty, template shifting was implemented. The procedure to generate the template was evaluated on <span class="hlt">images</span> of an anthropomorphic breast phantom containing microcalcification-related signals. Optimal reduction of the background noise was achieved without changing the signal. Based on a validation study in simulated <span class="hlt">images</span>, the difference (bias) in MO performance from the ground truth signal was calculated and found to be <1%. As template generation is a building stone of the entire <span class="hlt">image</span> quality assessment framework, the proposed method to construct templates from acquired <span class="hlt">images</span> facilitates the use of the NPWE MO in acquired <span class="hlt">images</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.P11B1271S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.P11B1271S"><span>Latitudinal Variations In Vertical Cloud Structure Of Jupiter As Determined By Ground- based <span class="hlt">Observation</span> With Multispectral <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sato, T.; Kasaba, Y.; Takahashi, Y.; Murata, I.; Uno, T.; Tokimasa, N.; Sakamoto, M.</p> <p>2008-12-01</p> <p>We conducted ground-based <span class="hlt">observation</span> of Jupiter with the liquid crystal tunable filter (LCTF) and EM-CCD camera in two methane absorption bands (700-757nm, 872-950nm at 3 nm step: total of 47 wavelengths) to derive detailed Jupiter's vertical cloud structure. The 2-meter reflector telescope at Nishi-Harima astronomical observatory in Japan was used for our <span class="hlt">observation</span> on 26-30 May, 2008. After a series of <span class="hlt">image</span> processing (composition of high quality <span class="hlt">images</span> in each wavelength and geometry calibration), we converted <span class="hlt">observed</span> intensity to absolute reflectivity at each pixel using standard star. As a result, we acquired Jupiter's data cubes with high-spatial resolution (about 1") and narrow band <span class="hlt">imaging</span> (typically 7nm) in each methane absorption band by superimposing 30 Jupiter's <span class="hlt">images</span> obtained in short exposure time (50 ms per one <span class="hlt">image</span>). These data sets enable us to probe different altitudes of Jupiter from 100 mbar down to 1bar level with higher vertical resolution than using convectional interference filters. To interpret <span class="hlt">observed</span> center-limb profiles, we developed radiative transfer code based on layer adding doubling algorithm to treat multiple scattering of solar light theoretically and extracted information on aerosol altitudes and optical properties using two-cloud model. First, we fit 5 different profiles simultaneously in continuum data (745-757 nm) to retrieve information on optical thickness of haze and single scattering albedo of cloud. Second, we fit 15 different profiles around 727nm methane absorption band and 13 different profiles around 890 nm methane absorption band to retrieve information on the aerosol altitude location and optical thickness of cloud. In this presentation, we present the results of these modeling simulations and discuss the latitudinal variations of Jupiter's vertical cloud structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSM21E..07G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSM21E..07G"><span>Ultraviolet <span class="hlt">Observations</span> of the Earth and Moon during the Juno Flyby</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gladstone, R.; Versteeg, M. H.; Davis, M.; Greathouse, T. K.; Gerard, J. M.; Grodent, D. C.; Bonfond, B.</p> <p>2013-12-01</p> <p>We present the initial results from Juno-UVS <span class="hlt">observations</span> of the Earth and Moon obtained during the flyby of the Juno spacecraft on 9 October 2013. Juno-UVS is an <span class="hlt">imaging</span> spectrograph with a bandpass of 70<λ<205 nm. This wavelength range includes all important ultraviolet (UV) emissions from the H2 bands and the H Lyman series which are produced in Jupiter's auroras, and also the absorption signatures of aurorally-produced hydrocarbons. The Juno-UVS instrument consists of two separate sections: a dedicated telescope/spectrograph assembly and a vault electronics box. The telescope/spectrograph assembly contains a telescope which feeds a 0.15-m Rowland circle spectrograph. The telescope has a 4 x 4 cm2 input aperture and uses an off-axis parabolic (OAP) primary mirror. A flat scan mirror situated at the front end of the telescope (used to <span class="hlt">observe</span> at up to ×30° perpendicular to the Juno spin plane) directs incoming light to the OAP. The light is focused onto the spectrograph entrance slit, which has a 'dog-bone' shape 7.2° long, in three sections of 0.2°, 0.025°, and 0.2° width (as projected onto the sky). Light entering the slit is dispersed by a toroidal grating which focuses UV light onto a curved microchannel plate cross delay line detector with a solar blind UV-sensitive CsI photocathode, which makes up the instrument's focal plane. Tantalum surrounds the detector assembly to shield it from high-energy electrons. The detector electronics are located behind the detector. All other electronics are located in a box inside Juno's spacecraft vault, including redundant low-voltage and high-voltage power supplies, command and data handling electronics, heater/actuator electronics, scan mirror electronics, and event processing electronics. The purpose of Juno-UVS is to remotely sense Jupiter's auroral morphology and brightness to provide context for in situ measurements by Juno's particle instruments. The recent Earth flyby provided an opportunity to: 1) use</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4478895','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4478895"><span>Implementation of a channelized Hotelling <span class="hlt">observer</span> model to assess <span class="hlt">image</span> quality of x-ray angiography systems</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Favazza, Christopher P.; Fetterly, Kenneth A.; Hangiandreou, Nicholas J.; Leng, Shuai; Schueler, Beth A.</p> <p>2015-01-01</p> <p>Abstract. Evaluation of flat-panel angiography equipment through conventional <span class="hlt">image</span> quality metrics is limited by the scope of standard spatial-domain <span class="hlt">image</span> quality metric(s), such as contrast-to-noise ratio and spatial resolution, or by restricted access to appropriate data to calculate Fourier domain measurements, such as modulation transfer function, noise power spectrum, and detective quantum efficiency. <span class="hlt">Observer</span> models have been shown capable of overcoming these limitations and are able to comprehensively evaluate medical-<span class="hlt">imaging</span> systems. We present a spatial domain-based channelized Hotelling <span class="hlt">observer</span> model to calculate the detectability index (DI) of our different sized disks and compare the performance of different <span class="hlt">imaging</span> conditions and angiography systems. When appropriate, changes in DIs were compared to expectations based on the classical Rose model of signal detection to assess linearity of the model with quantum signal-to-noise ratio (SNR) theory. For these experiments, the estimated uncertainty of the DIs was less than 3%, allowing for precise comparison of <span class="hlt">imaging</span> systems or conditions. For most experimental variables, DI changes were linear with expectations based on quantum SNR theory. DIs calculated for the smallest objects demonstrated nonlinearity with quantum SNR theory due to system blur. Two angiography systems with different detector element sizes were shown to perform similarly across the majority of the detection tasks. PMID:26158086</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29674564','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29674564"><span><span class="hlt">Observing</span> the cell in its native state: <span class="hlt">Imaging</span> subcellular dynamics in multicellular organisms.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Tsung-Li; Upadhyayula, Srigokul; Milkie, Daniel E; Singh, Ved; Wang, Kai; Swinburne, Ian A; Mosaliganti, Kishore R; Collins, Zach M; Hiscock, Tom W; Shea, Jamien; Kohrman, Abraham Q; Medwig, Taylor N; Dambournet, Daphne; Forster, Ryan; Cunniff, Brian; Ruan, Yuan; Yashiro, Hanako; Scholpp, Steffen; Meyerowitz, Elliot M; Hockemeyer, Dirk; Drubin, David G; Martin, Benjamin L; Matus, David Q; Koyama, Minoru; Megason, Sean G; Kirchhausen, Tom; Betzig, Eric</p> <p>2018-04-20</p> <p>True physiological <span class="hlt">imaging</span> of subcellular dynamics requires studying cells within their parent organisms, where all the environmental cues that drive gene expression, and hence the phenotypes that we actually <span class="hlt">observe</span>, are present. A complete understanding also requires volumetric <span class="hlt">imaging</span> of the cell and its surroundings at high spatiotemporal resolution, without inducing undue stress on either. We combined lattice light-sheet microscopy with adaptive optics to achieve, across large multicellular volumes, noninvasive aberration-free <span class="hlt">imaging</span> of subcellular processes, including endocytosis, organelle remodeling during mitosis, and the migration of axons, immune cells, and metastatic cancer cells in vivo. The technology reveals the phenotypic diversity within cells across different organisms and developmental stages and may offer insights into how cells harness their intrinsic variability to adapt to different physiological environments. Copyright © 2018 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6921407-imaging-observations-sn1987a-gamma-ray-energies','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6921407-imaging-observations-sn1987a-gamma-ray-energies"><span><span class="hlt">Imaging</span> <span class="hlt">observations</span> of SN1987A at gamma-ray energies</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>Cook, W.R.; Palmer, D.M.; Prince, T.A.</p> <p>1988-09-25</p> <p>The Caltech <span class="hlt">imaging</span> ..gamma..-ray telescope was launched by balloon from Alice Springs, NT, Australia for <span class="hlt">observations</span> of SN1987A during the period 18.60--18.87 November 1987 UT. The preliminary results presented here are derived from 8200 seconds of instrument livetime on the supernova and 2500 seconds on the Crab Nebula and pulsar at a float altitude of 37 km. We have obtained the first <span class="hlt">images</span> of the SN1987A region at ..gamma..-ray energies confirming that the bulk of the ..gamma..-ray emission comes from the supernova and not from LMC X-1. A count excess is detected between 300 and 1300 keV from the directionmore » of the supernova, one third of which comes from energy bands of width 80 and 92 keV centered on 847 and 1238 keV, respectively. The excess can be interpreted as a line photon flux plus scattered photon continuum from the radioactive decay of /sup 56/Co synthesized in the supernova explosion. We compare our data to recent predictions and find it to be consistent with models invoking moderate mixing of core material into the envelope.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988AIPC..170...60C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988AIPC..170...60C"><span><span class="hlt">Imaging</span> <span class="hlt">observations</span> of SN1987A at gamma-ray energies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cook, W. R.; Palmer, D. M.; Prince, T. A.; Schindler, S. M.; Starr, C. H.; Stone, E. C.</p> <p>1988-09-01</p> <p>The Caltech <span class="hlt">imaging</span> γ-ray telescope was launched by balloon from Alice Springs, NT, Australia for <span class="hlt">observations</span> of SN1987A during the period 18.60-18.87 November 1987 UT. The preliminary results presented here are derived from 8200 seconds of instrument livetime on the supernova and 2500 seconds on the Crab Nebula and pulsar at a float altitude of 37 km. We have obtained the first <span class="hlt">images</span> of the SN1987A region at γ-ray energies confirming that the bulk of the γ-ray emission comes from the supernova and not from LMC X-1. A count excess is detected between 300 and 1300 keV from the direction of the supernova, one third of which comes from energy bands of width 80 and 92 keV centered on 847 and 1238 keV, respectively. The excess can be interpreted as a line photon flux plus scattered photon continuum from the radioactive decay of 56Co synthesized in the supernova explosion. We compare our data to recent predictions and find it to be consistent with models invoking moderate mixing of core material into the envelope.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20020068003&hterms=cosmic+microwave+background&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcosmic%2Bmicrowave%2Bbackground','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20020068003&hterms=cosmic+microwave+background&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcosmic%2Bmicrowave%2Bbackground"><span>The Anisotropy of the Microwave Background to l=3500: Mosaic <span class="hlt">Observations</span> with the Cosmic Background <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pearson, T. J.; Mason, B. S.; Readhead, A. C. S.; Shepherd, M. C.; Sievers, J. L.; Udomprasert, P. S.; Cartwright, J. K.; Farmer, A. J.; Padin, S.; Myers, S. T.; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20020068003'); toggleEditAbsImage('author_20020068003_show'); toggleEditAbsImage('author_20020068003_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20020068003_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20020068003_hide"></p> <p>2002-01-01</p> <p>Using the Cosmic Background <span class="hlt">Imager</span>, a 13-element interferometer array operating in the 26-36 GHz frequency band, we have <span class="hlt">observed</span> 40 deg (sup 2) of sky in three pairs of fields, each approximately 145 feet x 165 feet, using overlapping pointings: (mosaicing). We present <span class="hlt">images</span> and power spectra of the cosmic microwave background radiation in these mosaic fields. We remove ground radiation and other low-level contaminating signals by differencing matched <span class="hlt">observations</span> of the fields in each pair. The primary foreground contamination is due to point sources (radio galaxies and quasars). We have subtracted the strongest sources from the data using higher-resolution measurements, and we have projected out the response to other sources of known position in the power-spectrum analysis. The <span class="hlt">images</span> show features on scales approximately 6 feet-15 feet, corresponding to masses approximately 5-80 x 10(exp 14) solar mass at the surface of last scattering, which are likely to be the seeds of clusters of galaxies. The power spectrum estimates have a resolution delta l approximately 200 and are consistent with earlier results in the multipole range l approximately less than 1000. The power spectrum is detected with high signal-to-noise ratio in the range 300 approximately less than l approximately less than 1700. For 1700 approximately less than l approximately less than 3000 the <span class="hlt">observations</span> are consistent with the results from more sensitive CBI deep-field <span class="hlt">observations</span>. The results agree with the extrapolation of cosmological models fitted to <span class="hlt">observations</span> at lower l, and show the predicted drop at high l (the "damping tail").</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/962373','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/962373"><span>Detection of facilities in satellite imagery using semi-supervised <span class="hlt">image</span> classification and auxiliary contextual <span class="hlt">observables</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>Harvey, Neal R; Ruggiero, Christy E; Pawley, Norma H</p> <p>2009-01-01</p> <p>Detecting complex targets, such as facilities, in commercially available satellite imagery is a difficult problem that human analysts try to solve by applying world knowledge. Often there are known <span class="hlt">observables</span> that can be extracted by pixel-level feature detectors that can assist in the facility detection process. Individually, each of these <span class="hlt">observables</span> is not sufficient for an accurate and reliable detection, but in combination, these auxiliary <span class="hlt">observables</span> may provide sufficient context for detection by a machine learning algorithm. We describe an approach for automatic detection of facilities that uses an automated feature extraction algorithm to extract auxiliary <span class="hlt">observables</span>, and a semi-supervisedmore » assisted target recognition algorithm to then identify facilities of interest. We illustrate the approach using an example of finding schools in Quickbird <span class="hlt">image</span> data of Albuquerque, New Mexico. We use Los Alamos National Laboratory's Genie Pro automated feature extraction algorithm to find a set of auxiliary features that should be useful in the search for schools, such as parking lots, large buildings, sports fields and residential areas and then combine these features using Genie Pro's assisted target recognition algorithm to learn a classifier that finds schools in the <span class="hlt">image</span> data.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006cosp...36.2801K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006cosp...36.2801K"><span>Seasonal variation of upper mesospheric temperatures from the OH and O2 nightglow over King Sejong Station, Antarctica</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, J.-H.; Kim, Y. H.; Moon, B.-K.; Chung, J.-K.; Won, Y.-I.</p> <p></p> <p>A spectral <span class="hlt">airglow</span> temperature <span class="hlt">imager</span> SATI was operated at King Sejong Station 62 22 r S 301 2 r E Korea Antarctic Research Station during a period of 2002 - 2005 Rotational temperatures from the OH 6-2 and O 2 0-1 band <span class="hlt">airglow</span> were obtained for more than 600 nights during the 4 year operation Both the OH and O 2 temperatures show similar seasonal variations which change significantly year by year A maximum temperature occurred early May in 2003 and 2004 whereas two maxima appeared in April and August in 2002 The 2005 data show only a broad and weak maximum during months of April and May The data also show oscillations with periods of hours that seem to relate to tides and gravity waves and fluctuations with timescales of days that could be due to planetary waves Detailed analysis will be performed to the data set to identify major atmospheric oscillations or variation over hours days and seasons</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3465370','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3465370"><span>Suomi satellite brings to light a unique frontier of nighttime environmental sensing capabilities</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Miller, Steven D.; Mills, Stephen P.; Elvidge, Christopher D.; Lindsey, Daniel T.; Lee, Thomas F.; Hawkins, Jeffrey D.</p> <p>2012-01-01</p> <p>Most environmental satellite radiometers use solar reflectance information when it is available during the day but must resort at night to emission signals from infrared bands, which offer poor sensitivity to low-level clouds and surface features. A few sensors can take advantage of moonlight, but the inconsistent availability of the lunar source limits measurement utility. Here we show that the Day/Night Band (DNB) low-light visible sensor on the recently launched Suomi National Polar-orbiting Partnership (NPP) satellite has the unique ability to <span class="hlt">image</span> cloud and surface features by way of reflected <span class="hlt">airglow</span>, starlight, and zodiacal light illumination. Examples collected during new moon reveal not only meteorological and surface features, but also the direct emission of <span class="hlt">airglow</span> structures in the mesosphere, including expansive regions of diffuse glow and wave patterns forced by tropospheric convection. The ability to leverage diffuse illumination sources for nocturnal environmental sensing applications extends the advantages of visible-light information to moonless nights. PMID:22984179</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApJ...845...30M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApJ...845...30M"><span>Structure and Dynamics of Cool Flare Loops <span class="hlt">Observed</span> by the Interface Region <span class="hlt">Imaging</span> Spectrograph</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mikuła, K.; Heinzel, P.; Liu, W.; Berlicki, A.</p> <p>2017-08-01</p> <p>Flare loops were well <span class="hlt">observed</span> with the Interface Region <span class="hlt">Imaging</span> Spectrograph (IRIS) during the gradual phase of two solar flares on 2014 March 29 and 2015 June 22. Cool flare loops are visible in various spectral lines formed at chromospheric and transition-region temperatures and exhibit large downflows which correspond to the standard scenario. The principal aim of this work is to analyze the structure and dynamics of cool flare loops <span class="hlt">observed</span> in Mg II lines. Synthetic profiles of the Mg II h line are computed using the classical cloud model and assuming a uniform background intensity. In this paper, we study novel IRIS NUV <span class="hlt">observations</span> of such loops in Mg II h and k lines and also show the behavior of hotter lines detected in the FUV channel. We obtained the spatial evolution of the velocities: near the loop top, the flow velocities are small and they are increasing toward the loop legs. Moreover, from slit-jaw <span class="hlt">image</span> (SJI) movies, we <span class="hlt">observe</span> some plasma upflows into the loops, which are also detectable in Mg II spectra. The brightness of the loops systematically decreases with increasing flow velocity, and we ascribe this to the effect of Doppler dimming, which works for Mg II lines. Emission profiles of Mg II were found to be extremely broad, and we explain this through the large unresolved non-thermal motions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22663283-structure-dynamics-cool-flare-loops-observed-interface-region-imaging-spectrograph','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22663283-structure-dynamics-cool-flare-loops-observed-interface-region-imaging-spectrograph"><span>Structure and Dynamics of Cool Flare Loops <span class="hlt">Observed</span> by the Interface Region <span class="hlt">Imaging</span> Spectrograph</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>Mikuła, K.; Berlicki, A.; Heinzel, P.</p> <p></p> <p>Flare loops were well <span class="hlt">observed</span> with the Interface Region <span class="hlt">Imaging</span> Spectrograph ( IRIS ) during the gradual phase of two solar flares on 2014 March 29 and 2015 June 22. Cool flare loops are visible in various spectral lines formed at chromospheric and transition-region temperatures and exhibit large downflows which correspond to the standard scenario. The principal aim of this work is to analyze the structure and dynamics of cool flare loops <span class="hlt">observed</span> in Mg ii lines. Synthetic profiles of the Mg ii h line are computed using the classical cloud model and assuming a uniform background intensity. In thismore » paper, we study novel IRIS NUV <span class="hlt">observations</span> of such loops in Mg ii h and k lines and also show the behavior of hotter lines detected in the FUV channel. We obtained the spatial evolution of the velocities: near the loop top, the flow velocities are small and they are increasing toward the loop legs. Moreover, from slit-jaw <span class="hlt">image</span> (SJI) movies, we <span class="hlt">observe</span> some plasma upflows into the loops, which are also detectable in Mg ii spectra. The brightness of the loops systematically decreases with increasing flow velocity, and we ascribe this to the effect of Doppler dimming, which works for Mg ii lines. Emission profiles of Mg ii were found to be extremely broad, and we explain this through the large unresolved non-thermal motions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003EAEJA.....8673E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003EAEJA.....8673E"><span>Science objectives and <span class="hlt">observing</span> strategy for the OMEGA <span class="hlt">imaging</span> spectrometer on Mars-Express</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Erard, S.; Bibring, J.-P.; Drossart, P.; Forget, F.; Schmitt, B.; OMEGA Team</p> <p>2003-04-01</p> <p>The science objectives of OMEGA, which were first defined at the time of instruments selection for Mars-Express, were recently updated to integrate new results from MGS and Odyssey concerning three main fields: Martian surface and atmosphere, and polar processes. Thematic categories of <span class="hlt">observations</span> are derived from the scientific objectives whenever spectral <span class="hlt">observations</span> from OMEGA are expected to provide insights to Mars present situation and evolution. Targets within these categories are selected on the basis of their expected usefulness, which is related to their intrinsic properties and to the instrument capabilities. The whole surface will be mapped at low resolution (~5 km/pixel) in the course of the nominal mission, and possibly routinely at very coarse resolution to monitor time-varying processes from apocenter. However, only 5% of the surface can be <span class="hlt">observed</span> at high resolution (up to 350 m/pixel) owing to constraints on telemetry rate. HR targets are therefore selected on the basis of telemetry constraints, orbital parameters, <span class="hlt">observing</span> opportunities (visibility under given conditions), and spacecraft functionalities (e.g., depointing capacity), then prioritized within each category according to the probability to perform significant <span class="hlt">observations</span> with OMEGA (in many situations, according to the estimated dust coverage). Target selection is performed interactively between OMEGA co-Is, in close contact with teams from other MEx experiments (mostly HRSC, PFS and Spicam) and other missions (e.g., MER and MRO). Most HR surface targets are selected on the basis of deep examination of Viking, THEMIS, and MOC HR <span class="hlt">images</span>. Other surface targets include areas presenting unusual spectral properties in previous <span class="hlt">observations</span>, or suspected to exhibit signatures of hydrothermal activity. Proposed landing sites and suggested source areas for the SNC meteorites are also included. Atmospheric/polar objectives more often translate as particular <span class="hlt">observing</span> modes, sometimes at HR</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('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 BRIGHT GRAINS <span class="hlt">OBSERVED</span> IN THE TRANSITION REGION <span class="hlt">IMAGING</span> 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 <span class="hlt">Imaging</span> Spectrograph (IRIS) provides spectroscopy and narrow band slit-jaw (SJI) <span class="hlt">imaging</span> of the solar chromosphere and transition region at unprecedented spatial and temporal resolutions. Combined with high-resolution context spectral <span class="hlt">imaging</span> 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 <span class="hlt">images</span> from active regions, which primarily sample the transition region with the Si iv 1394 and 1403 Å lines, reveal ubiquitous bright “grains” which are short-lived (two to five minute)more » bright 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 bright 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://adsabs.harvard.edu/abs/1982STIN...8318976F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1982STIN...8318976F"><span>Application of X-ray television <span class="hlt">image</span> system to <span class="hlt">observation</span> in solid rocket motor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fujiwara, T.; Ito, K.; Tanemura, T.; Shimizu, M.; Godai, T.</p> <p></p> <p>The X-ray television <span class="hlt">image</span> system is used to <span class="hlt">observe</span> the solid propellant burning surface during rocket motor operation as well as to inspect defects in solid rocket motors in a real time manner. This system can test 200 mm diameter dummy propellant rocket motors with under 2 percent discriminative capacity. Viewing of a 50 mm diameter internal-burning rocket motor, propellant burning surface time transition and propellant burning process of the surroundings of artificial defects were satisfactorily <span class="hlt">observed</span>. The system was demonstrated to be effective for nondestructive testing and combustion research of solid rocket motors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21301646-deep-band-imaging-goods-south-observations-data-reduction-first-results','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21301646-deep-band-imaging-goods-south-observations-data-reduction-first-results"><span>DEEP U BAND AND R <span class="hlt">IMAGING</span> OF GOODS-SOUTH: <span class="hlt">OBSERVATIONS</span>, DATA REDUCTION AND FIRST RESULTS ,</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>Nonino, M.; Cristiani, S.; Vanzella, E.</p> <p>2009-08-01</p> <p>We present deep <span class="hlt">imaging</span> in the U band covering an area of 630 arcmin{sup 2} centered on the southern field of the Great Observatories Origins Deep Survey (GOODS). The data were obtained with the VIMOS instrument at the European Southern Observatory (ESO) Very Large Telescope. The final <span class="hlt">images</span> reach a magnitude limit U {sub lim} {approx} 29.8 (AB, 1{sigma}, in a 1'' radius aperture), and have good <span class="hlt">image</span> quality, with full width at half-maximum {approx}0.''8. They are significantly deeper than previous U-band <span class="hlt">images</span> available for the GOODS fields, and better match the sensitivity of other multiwavelength GOODS photometry. The deepermore » U-band data yield significantly improved photometric redshifts, especially in key redshift ranges such as 2 < z < 4, and deeper color-selected galaxy samples, e.g., Lyman break galaxies at z {approx} 3. We also present the co-addition of archival ESO VIMOS R-band data, with R {sub lim} {approx} 29 (AB, 1{sigma}, 1'' radius aperture), and <span class="hlt">image</span> quality {approx}0.''75. We discuss the strategies for the <span class="hlt">observations</span> and data reduction, and present the first results from the analysis of the co-added <span class="hlt">images</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApJ...749L..23D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApJ...749L..23D"><span>First Science <span class="hlt">Observations</span> with SOFIA/FORCAST: 6-37 μm <span class="hlt">Imaging</span> of Orion BN/KL</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>De Buizer, James M.; Morris, Mark R.; Becklin, E. E.; Zinnecker, Hans; Herter, Terry L.; Adams, Joseph D.; Shuping, Ralph Y.; Vacca, William D.</p> <p>2012-04-01</p> <p>The Becklin-Neugebauer/Kleinmann-Low (BN/KL) region of the Orion Nebula is the nearest region of high-mass star formation in our galaxy. As such, it has been the subject of intense investigation at a variety of wavelengths, which have revealed it to be brightest in the infrared to submillimeter wavelength regime. Using the newly commissioned SOFIA airborne telescope and its 5-40 μm camera FORCAST, <span class="hlt">images</span> of the entire BN/KL complex have been acquired. The 31.5 and 37.1 μm <span class="hlt">images</span> represent the highest resolution <span class="hlt">observations</span> (lsim4'') ever obtained of this region at these wavelengths. These <span class="hlt">observations</span> reveal that the BN object is not the dominant brightness source in the complex at wavelengths >= 31.5 μm and that this distinction goes instead to the source IRc4. It was determined from these <span class="hlt">images</span> and derived dust color temperature maps that IRc4 is also likely to be self-luminous. A new source of emission has also been identified at wavelengths >= 31.5 μm that coincides with the northeastern outflow lobe from the protostellar disk associated with radio source I.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100033284','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100033284"><span>MAPIR: An Airborne Polarmetric <span class="hlt">Imaging</span> Radiometer in Support of Hydrologic 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>Laymon, C.; Al-Hamdan, M.; Crosson, W.; Limaye, A.; McCracken, J.; Meyer, P.; Richeson, J.; Sims, W.; Srinivasan, K.; Varnevas, K.</p> <p>2010-01-01</p> <p>In this age of dwindling water resources and increasing demands, accurate estimation of water balance components at every scale is more critical to end users than ever before. Several near-term Earth science satellite missions are aimed at global hydrologic <span class="hlt">observations</span>. The Marshall Airborne Polarimetric <span class="hlt">Imaging</span> Radiometer (MAPIR) is a dual beam, dual angle polarimetric, scanning L band passive microwave radiometer system developed by the <span class="hlt">Observing</span> Microwave Emissions for Geophysical Applications (OMEGA) team at MSFC to support algorithm development and validation efforts in support of these missions. MAPIR <span class="hlt">observes</span> naturally-emitted radiation from the ground primarily for remote sensing of land surface brightness temperature from which we can retrieve soil moisture and possibly surface or water temperature and ocean salinity. MAPIR has achieved Technical Readiness Level 6 with flight heritage on two very different aircraft, the NASA P-3B, and a Piper Navajo.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA31B..07S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA31B..07S"><span>Vertical Coupling and <span class="hlt">Observable</span> Effects of Evanescent Acoustic-Gravity Waves in the Mesosphere and Thermosphere</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Snively, J. B.</p> <p>2017-12-01</p> <p>Our understanding of acoustic-gravity wave (AGW) dynamics at short periods ( minutes to hour) and small scales ( 10s to 100s km) in the mesosphere, thermosphere, and ionosphere (MTI) has benefited considerably from horizontally- and vertically-resolved measurements of layered species. These include, for example, imagery of the mesopause ( 80-100 km) <span class="hlt">airglow</span> layers and vertical profiles of the sodium layer via lidar [e.g., Taylor and Hapgood, PSS, 36(10), 1988; Miller et al., PNAS, 112(49), 2015; Cao et al., JGR, 121, 2016]. In the thermosphere-ionosphere, AGW perturbations are also revealed in electron density profiles [Livneh et al., JGR, 112, 2007] and maps of total electron content (TEC) from global positioning system (GPS) receivers [Nishioka et al., GRL, 40(21), 2013]. To the extent that AGW signatures in layered species can be quantified, and the ambient atmospheric state measured or estimated, numerical models enable investigations of dynamics at intermediate altitudes that cannot readily be measured (e.g., above and below the 80-100 km mesopause region). Here, new 2D and 3D versions of the Model for Acoustic-Gravity Wave Interactions and Coupling (MAGIC) [e.g., Snively and Pasko, JGR, 113(A6), 2008, and references therein] are introduced and applied to investigate spectra of short-period AGW that can pass through the mesopause region to reach and impact the thermosphere. Simulation case studies are constructed to investigate both their signatures through the hydroxyl <span class="hlt">airglow</span> layer [e.g., Snively et al., JGR 115(A11), 2010] and their effects above. These waves, with large vertical wavelengths and fast horizontal phase speeds, also include those that may be subject to evanescence at mesopause or in the middle-thermosphere, with potential for ducting or dissipation between where static stability is higher. Despite complicating interpretations of momentum fluxes, evanescence plays an under-appreciated role in vertical coupling by AGW [Walterscheid and Hecht</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004cosp...35.1787S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004cosp...35.1787S"><span>Star <span class="hlt">Observations</span> by Asteroid Multiband <span class="hlt">Imaging</span> Camera (AMICA) on Hayabusa (MUSES-C) Cruising Phase</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Saito, J.; Hashimoto, T.; Kubota, T.; Hayabusa AMICA Team</p> <p></p> <p>Muses-C is the first Japanese asteroid mission and also a technology demonstration one to the S-type asteroid, 25143 Itokawa (1998SF36). It was launched at May 9, 2003, and renamed Hayabusa after the spacecraft was confirmed to be on the interplanetary orbit. This spacecraft has the event of the Earth-swingby for gravitational assist in the way to Itokawa on 2004 May. The arrival to Itokawa is scheduled on 2005 summer. During the visit to Itokawa, the remote-sensing <span class="hlt">observation</span> with AMICA, NIRS (Near Infrared Spectrometer), XRS (X-ray Fluorescence Spectrometer), and LIDAR are performed, and the spacecraft descends and collects the surface samples at the touch down to the surface. The captured asteroid sample will be returned to the Earth in the middle of 2007. The telescopic optical navigation camera (ONC-T) with seven bandpass filters (and one wide-band filter) and polarizers is called AMICA (Asteroid Multiband <span class="hlt">Imaging</span> CAmera) when ONC-T is used for scientific <span class="hlt">observations</span>. The AMICA's seven bandpass filters are nearly equivalent to the seven filters of the ECAS (Eight Color Asteroid Survey) system. Obtained spectroscopic data will be compared with previously obtained ECAS <span class="hlt">observations</span>. AMICA also has four polarizers, which are located on one edge of the CCD chip (covering 1.1 x 1.1 degrees each). Using the polarizers of AMICA, we can obtain polarimetric information of the target asteroid's surface. Since last November, we planned the test <span class="hlt">observations</span> of some stars and planets by AMICA and could successfully obtain these <span class="hlt">images</span>. Here, we briefly report these <span class="hlt">observations</span> and its calibration by the ground-based <span class="hlt">observational</span> data. In addition, we also present a current status of AMICA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3230945','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3230945"><span>Multitemporal <span class="hlt">Observations</span> of Sugarcane by TerraSAR-X <span class="hlt">Images</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>Baghdadi, Nicolas; Cresson, Rémi; Todoroff, Pierre; Moinet, Soizic</p> <p>2010-01-01</p> <p>The objective of this study is to investigate the potential of TerraSAR-X (X-band) in monitoring sugarcane growth on Reunion Island (located in the Indian Ocean). Multi-temporal TerraSAR data acquired at various incidence angles (17°, 31°, 37°, 47°, 58°) and polarizations (HH, HV, VV) were analyzed in order to study the behaviour of SAR (synthetic aperture radar) signal as a function of sugarcane height and NDVI (Normalized Difference Vegetation Index). The potential of TerraSAR for mapping the sugarcane harvest was also studied. Radar signal increased quickly with crop height until a threshold height, which depended on polarization and incidence angle. Beyond this threshold, the signal increased only slightly, remained constant, or even decreased. The threshold height is slightly higher with cross polarization and higher incidence angles (47° in comparison with 17° and 31°). Results also showed that the co-polarizations channels (HH and VV) were well correlated. High correlation between SAR signal and NDVI calculated from SPOT-4/5 <span class="hlt">images</span> was <span class="hlt">observed</span>. TerraSAR data showed that after strong rains the soil contribution to the backscattering of sugarcane fields can be important for canes with heights of terminal visible dewlap (htvd) less than 50 cm (total cane heights around 155 cm). This increase in radar signal after strong rains could involve an ambiguity between young and mature canes. Indeed, the radar signal on TerraSAR <span class="hlt">images</span> acquired in wet soil conditions could be of the same order for fields recently harvested and mature sugarcane fields, making difficult the detection of cuts. Finally, TerraSAR data at high spatial resolution were shown to be useful for monitoring sugarcane harvest when the fields are of small size or when the cut is spread out in time. The comparison between incidence angles of 17°, 37° and 58° shows that 37° is more suitable to monitor the sugarcane harvest. The cut is easily detectable on TerraSAR <span class="hlt">images</span> for data acquired</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002SPIE.4814..296M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002SPIE.4814..296M"><span>Flight Test Results of the Earth <span class="hlt">Observing</span>-1 Advanced Land <span class="hlt">Imager</span> Advanced Land <span class="hlt">Imager</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mendenhall, Jeffrey A.; Lencioni, Donald E.; Hearn, David R.; Digenis, Constantine J.</p> <p>2002-09-01</p> <p>The Advanced Land <span class="hlt">Imager</span> (ALI) is the primary instrument on the Earth <span class="hlt">Observing</span>-1 spacecraft (EO-1) and was developed under NASA's New Millennium Program (NMP). The NMP mission objective is to flight-validate advanced technologies that will enable dramatic improvements in performance, cost, mass, and schedule for future, Landsat-like, Earth Science Enterprise instruments. ALI contains a number of innovative features designed to achieve this objective. These include the basic instrument architecture, which employs a push-broom data collection mode, a wide field-of-view optical design, compact multi-spectral detector arrays, non-cryogenic HgCdTe for the short wave infrared bands, silicon carbide optics, and a multi-level solar calibration technique. The sensor includes detector arrays that operate in ten bands, one panchromatic, six VNIR and three SWIR, spanning the range from 0.433 to 2.35 μm. Launched on November 21, 2000, ALI instrument performance was monitored during its first year on orbit using data collected during solar, lunar, stellar, and earth <span class="hlt">observations</span>. This paper will provide an overview of EO-1 mission activities during this period. Additionally, the on-orbit spatial and radiometric performance of the instrument will be compared to pre-flight measurements and the temporal stability of ALI will be presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013DPS....4521104G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013DPS....4521104G"><span>The Ultraviolet Spectrograph (UVS) on ESA’s JUICE Mission</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gladstone, Randy; Retherford, K.; Steffl, A.; Eterno, J.; Davis, M.; Versteeg, M.; Greathouse, T.; Araujo, M.; Walther, B.; Persson, K.; Persyn, S.; Dirks, G.; McGrath, M.; Feldman, P.; Bagenal, F.; Spencer, J.; Schindhelm, E.; Fletcher, L.</p> <p>2013-10-01</p> <p>The Jupiter Icy Moons Explorer (JUICE) was selected in May 2012 as the first L-class mission of ESA’s Cosmic Vision Program. JUICE will launch in 2022 on a 7.6-year journey to the Jovian system, including a Venus and multiple Earth gravity assists, before entering Jupiter orbit in January 2030. JUICE will study the entire Jovian system for 3.5 years, concentrating on Europa, Ganymede, and Callisto, with the last 10 months spent in Ganymede orbit. The Ultraviolet Spectrograph (UVS) on JUICE was jointly selected by NASA and ESA as part of its ~130 kg payload of 11 scientific instruments. UVS is the fifth in a series of successful ultraviolet <span class="hlt">imaging</span> spectrographs (Rosetta-Alice, New Horizons Pluto-Alice, LRO-LAMP) and is largely based on the most recent of these, Juno-UVS. It <span class="hlt">observes</span> photons in the 55-210 nm wavelength range, at moderate spectral and spatial resolution along a 7.5-degree slit. A main entrance “<span class="hlt">airglow</span> port” (AP) is used for most <span class="hlt">observations</span> (e.g., <span class="hlt">airglow</span>, aurora, surface mapping, and stellar occultations), while a separate “solar port” (SP) allows for solar occultations. Another aperture door, with a small hole through the centre, is used as a “high-spatial-resolution port” (HP) for detailed <span class="hlt">observations</span> of bright targets. Time-tagging (pixel list mode) and programmable spectral <span class="hlt">imaging</span> (histogram mode) allow for <span class="hlt">observational</span> flexibility and optimal data management. As on Juno-UVS, the effects of penetrating electron radiation on electronic parts and data quality are substantially mitigated through contiguous shielding, filtering of pulse height amplitudes, management of high voltage settings, and careful use of radiation-hard, flight-tested parts. The science goals of UVS are to: 1) explore the atmospheres, plasma interactions, and surfaces of the Galilean satellites; 2) determine the dynamics, chemistry, and vertical structure of Jupiter’s upper atmosphere from equator to pole; and 3) investigate the Jupiter-Io connection by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFMSA41B..02W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFMSA41B..02W"><span>Using Heliospheric <span class="hlt">Imaging</span> for Storm Forecasting - SMEI CME <span class="hlt">Observations</span> as a Tool for Operational Forecasting at AFWA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Webb, D. F.; Johnston, J. C.; Fry, C. D.; Kuchar, T. A.</p> <p>2008-12-01</p> <p><span class="hlt">Observations</span> of coronal mass ejections (CMEs) from heliospheric <span class="hlt">imagers</span> such as the Solar Mass Ejection <span class="hlt">Imager</span> (SMEI) can lead to significant improvements in operational space weather forecasting. We are working with the Air Force Weather Agency (AFWA) to ingest SMEI all-sky imagery with appropriate tools to help forecasters improve their operational space weather forecasts. We describe two approaches: 1) Near- real time analysis of propagating CMEs from SMEI <span class="hlt">images</span> alone combined with near-Sun <span class="hlt">observations</span> of CME onsets and, 2) Using these calculations of speed as a mid-course correction to the HAFv2 solar wind model forecasts. HAFv2 became operational at AFWA in late 2006. The objective is to determine a set of practical procedures that the duty forecaster can use to update or correct a solar wind forecast using heliospheric <span class="hlt">imager</span> data. SMEI <span class="hlt">observations</span> can be used inclusively to make storm forecasts, as recently discussed in Webb et al. (Space Weather, in press, 2008). We have developed a point-and-click analysis tool for use with SMEI <span class="hlt">images</span> and are working with AFWA to ensure that timely SMEI <span class="hlt">images</span> are available for analyses. When a frontside solar eruption occurs, especially if within about 45 deg. of Sun center, a forecaster checks for an associated CME <span class="hlt">observed</span> by a coronagraph within an appropriate time window. If found, especially if the CME is a halo type, the forecaster checks SMEI <span class="hlt">observations</span> about a day later, depending on the apparent initial CME speed, for possibly associated CMEs. If one is found, then the leading edge is measured over several successive frames and an elongation-time plot constructed. A minimum of three data points, i.e., over 3-4 orbits or about 6 hours, are necessary for such a plot. Using the solar source location and onset time of the CME from, e.g., SOHO <span class="hlt">observations</span>, and assuming radial propagation, a distance-time relation is calculated and extrapolated to the 1 AU distance. As shown by Webb et al., the storm onset time</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFMSM51C2196P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFMSM51C2196P"><span>The Origin of Low Altitude ENA Emissions from Storms in 2000-2005 as <span class="hlt">Observed</span> by <span class="hlt">IMAGE</span>/MENA</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Perez, J. D.; Sheehan, M. M.; Jahn, J.; Mackler, D.; Pollock, C. J.</p> <p>2013-12-01</p> <p>Low Altitude Emissions (LAEs) are prevalent features of Energetic Neutral Atom (ENA) <span class="hlt">images</span> of the inner magnetosphere. It is believed that they are created by precipitating ions that reach altitudes near 500 km and then charge exchange with oxygen atoms, subsequently escaping to be <span class="hlt">observed</span> by satellite borne ENA <span class="hlt">imagers</span>. In this study, LAEs from the MENA instrument onboard the <span class="hlt">IMAGE</span> satellite are studied in order to learn about the origin of the precipitating ions. Using the Tsyganenko 05 magnetic field model, the bright pixels capturing the LAEs are mapped to the equator. The LAEs are believed to originate from ions near their mirroring point, i.e., with pitch angles near 90o. Therefore the angle between the line-of-sight and the magnetic field at the point of origin is used to further constrain possible magnetospheric regions that are the origin of the ENAs. By <span class="hlt">observing</span> the time dependence of the strength and location of the LAEs during geomagnetic storms in the years 2000-2005, the dynamics of the emptying and filling of the loss cone by injected particles is <span class="hlt">observed</span>. Thus, information regarding the coupling between the inner magnetosphere and the ionosphere is obtained.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=S81-40245&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=S81-40245&hterms=photo+image&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphoto%2Bimage"><span>Photos of earth <span class="hlt">observations</span> taken by JPL with the Shuttle <span class="hlt">Imaging</span> Radar-A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1981-01-01</p> <p>Photos of earth <span class="hlt">observations</span> taken by the Jet Propulsion Laboratory (JPL) with the Shuttle <span class="hlt">Imaging</span> Radar-A (SIR-A). The first <span class="hlt">image</span> show northern Peloponnesia and part of southern Greece. The Corinthian Canal is visible as a bright line cutting across the narrow Corinth Isthmus (upper center). Black area to the right is the Aegian Sea; on the left, the Gulf of Corinth. Islands to the right, starting at the top, are Salamis, Aegene and Angistrion, and the Peninsula of Methana. Southwest of the canal on the gulf coast is the city of Corinth, appearing as bright, white spots (40244); This <span class="hlt">image</span> shows the Hamersley mountain range in Western Australia. A circular pattern of eroded folds surround a prominent granite dome, remnants of a volcanic past, is seen in the center of the photograph. The Hardey River is seen running vertically to the right of the center circular dome, and the small town of Paraburdoo appears as a patch of tiny bright rectangles in the lower right corner (40245).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080021640','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080021640"><span>Atmospheric Correction Prototype Algorithm for High Spatial Resolution Multispectral Earth <span class="hlt">Observing</span> <span class="hlt">Imaging</span> Systems</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pagnutti, Mary</p> <p>2006-01-01</p> <p>This viewgraph presentation reviews the creation of a prototype algorithm for atmospheric correction using high spatial resolution earth <span class="hlt">observing</span> <span class="hlt">imaging</span> systems. The objective of the work was to evaluate accuracy of a prototype algorithm that uses satellite-derived atmospheric products to generate scene reflectance maps for high spatial resolution (HSR) systems. This presentation focused on preliminary results of only the satellite-based atmospheric correction algorithm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930047931&hterms=environment+attitudes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Denvironment%2Battitudes','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930047931&hterms=environment+attitudes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Denvironment%2Battitudes"><span>A study of the impact of the Space Shuttle environment on faint far-UV geophysical and astronomical phenomena</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lampton, Michael; Sasseen, Timothy P.; Wu, Xiaoyi; Bowyer, Stuart</p> <p>1993-01-01</p> <p>FAUST is a far ultraviolet (1400-1800 A) photon-counting <span class="hlt">imaging</span> telescope featuring a wide field of view (7.6 deg) and a high sensitivity to extended emission features. During its flight as part of the ATLAS-1 payload aboard the STS-45 mission in March 1992, 19 deep-space nighttime viewing opportunities were utilized by FAUST. Here we report the <span class="hlt">observed</span> fluxes and their time and space variations, and identify the signatures of postsunset <span class="hlt">airglow</span> phenomena and Orbiter Vernier attitude control thruster firing events. We find that the Space Shuttle nighttime environment at 296 km altitude is often sufficiently dark to permit geophysical and astronomical UV <span class="hlt">observations</span> down to levels on the order of 1000 photons/sq cm sr A sec, or 0.01 Rayleighs/A. We also find evidence for occasional geophysical fluxes of some tens or hundreds of Rayleighs in the upward-looking direction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010004238','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010004238"><span>The 1999 Leonid Multi-Instrument Aircraft Campaign - An Early Review</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jenniskens, Peter; Butow, Steven J.; Fonda, Mark; DeVincenzi, Donald L. (Technical Monitor)</p> <p>2000-01-01</p> <p>The Leonid meteor storm of 1999 was <span class="hlt">observed</span> from two B707-type research aircraft by a team of 35 scientists of seven nationalities over the Mediterranean Sea on Nov. 18, 1999. The mission was sponsored by various science programs of NASA, and offered the best possible <span class="hlt">observing</span> conditions, free of clouds and at a prime location for viewing the storm. The 1999 mission followed a similar effort in 1998, improving upon mission strategy and scope. As before, spectroscopic and <span class="hlt">imaging</span> experiments targeted meteors and persistent trains, but also <span class="hlt">airglow</span>, aurora, elves and sprites. The research aimed to address outstanding questions in Planetary Science, Astronomy, Astrobiology and upper atmospheric research, including Aeronornie. In addition, near real-time flux measurements contributed to a USAF sponsored program for space weather awareness. An overview of the first results is given, which are discussed in preparation for future missions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ChPhC..40h6203G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ChPhC..40h6203G"><span>Application of a multiscale maximum entropy <span class="hlt">image</span> restoration algorithm to HXMT <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>Guan, Ju; Song, Li-Ming; Huo, Zhuo-Xi</p> <p>2016-08-01</p> <p>This paper introduces a multiscale maximum entropy (MSME) algorithm for <span class="hlt">image</span> restoration of the Hard X-ray Modulation Telescope (HXMT), which is a collimated scan X-ray satellite mainly devoted to a sensitive all-sky survey and pointed <span class="hlt">observations</span> in the 1-250 keV range. The novelty of the MSME method is to use wavelet decomposition and multiresolution support to control noise amplification at different scales. Our work is focused on the application and modification of this method to restore diffuse sources detected by HXMT scanning <span class="hlt">observations</span>. An improved method, the ensemble multiscale maximum entropy (EMSME) algorithm, is proposed to alleviate the problem of mode mixing exiting in MSME. Simulations have been performed on the detection of the diffuse source Cen A by HXMT in all-sky survey mode. The results show that the MSME method is adapted to the deconvolution task of HXMT for diffuse source detection and the improved method could suppress noise and improve the correlation and signal-to-noise ratio, thus proving itself a better algorithm for <span class="hlt">image</span> restoration. Through one all-sky survey, HXMT could reach a capacity of detecting a diffuse source with maximum differential flux of 0.5 mCrab. Supported by Strategic Priority Research Program on Space Science, Chinese Academy of Sciences (XDA04010300) and National Natural Science Foundation of China (11403014)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8446E..1XD','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8446E..1XD"><span>A demonstration test of the dual-beam polarimetry differential <span class="hlt">imaging</span> system for the high-contrast <span class="hlt">observation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dou, Jiangpei; Ren, Deqing; Zhu, Yongtian; Wang, Xue; Zhang, Xi; Li, Rong</p> <p>2012-09-01</p> <p>We propose a dual-beam polarimetry differential <span class="hlt">imaging</span> test system that can be used for the direct <span class="hlt">imaging</span> of the exoplanets. The system is composed of a liquid crystal variable retarder (LCVR) in the pupil to switch between two orthogonal polarized states, and a Wollaston prism (WP) that will be inserted before the final focal focus of the system to create two polarized <span class="hlt">images</span> for the differential subtraction. Such a system can work separately or be integrated in the coronagraph system to enhance the high-contrast <span class="hlt">imaging</span>. To demonstrate the feasibility of the proposed system, here we show the initial test result both with and without integrating our developed coronagraph. A unique feature for this system is that each channel can subtract with itself by using the retarder to rotate the planet's polarization orientation, which has the best performance according to our lab test results. Finally, it is shown that the polarimetry differential <span class="hlt">imaging</span> system is a promising technique and can be used for the direct <span class="hlt">imaging</span> <span class="hlt">observation</span> of reflected lights from the exoplanets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009SPIE.7263E..0IP','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009SPIE.7263E..0IP"><span>Spatial frequency characteristics at <span class="hlt">image</span> decision-point locations for <span class="hlt">observers</span> with different radiological backgrounds in lung nodule detection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pietrzyk, Mariusz W.; Manning, David J.; Dix, Alan; Donovan, Tim</p> <p>2009-02-01</p> <p>Aim: The goal of the study is to determine the spatial frequency characteristics at locations in the <span class="hlt">image</span> of overt and covert <span class="hlt">observers</span>' decisions and find out if there are any similarities in different <span class="hlt">observers</span>' groups: the same radiological experience group or the same accuracy scored level. Background: The radiological task is described as a visual searching decision making procedure involving visual perception and cognitive processing. Humans perceive the world through a number of spatial frequency channels, each sensitive to visual information carried by different spatial frequency ranges and orientations. Recent studies have shown that particular physical properties of local and global <span class="hlt">image</span>-based elements are correlated with the performance and the level of experience of human <span class="hlt">observers</span> in breast cancer and lung nodule detections. Neurological findings in visual perception were an inspiration for wavelet applications in vision research because the methodology tries to mimic the brain processing algorithms. Methods: The wavelet approach to the set of postero-anterior chest radiographs analysis has been used to characterize perceptual preferences <span class="hlt">observers</span> with different levels of experience in the radiological task. Psychophysical methodology has been applied to track eye movements over the <span class="hlt">image</span>, where particular ROIs related to the <span class="hlt">observers</span>' fixation clusters has been analysed in the spaces frame by Daubechies functions. Results: Significance differences have been found between the spatial frequency characteristics at the location of different decisions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28267641','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28267641"><span>Graphene-supporting films and low-voltage STEM in SEM toward <span class="hlt">imaging</span> nanobio materials without staining: <span class="hlt">Observation</span> of insulin amyloid fibrils.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ogawa, Takashi; Gang, Geun Won; Thieu, Minh Thu; Kwon, Hyuksang; Ahn, Sang Jung; Ha, Tai Hwan; Cho, Boklae</p> <p>2017-05-01</p> <p>Utilization of graphene-supporting films and low-voltage scanning transmission electron microscopy (LV-STEM) in scanning electron microscopy (SEM) is shown to be an effective means of <span class="hlt">observing</span> unstained nanobio materials. Insulin amyloid fibrils, which are implicated as a cause of type II diabetes, are formed in vitro and <span class="hlt">observed</span> without staining at room temperature. An in-lens cold field-emission SEM, equipped with an additional homemade STEM detector, provides dark field (DF)-STEM <span class="hlt">images</span> in the low energy range of 5-30keV, together with secondary electron (SE) <span class="hlt">images</span>. Analysis based on Lenz's theory is used to interpret the experimental results. Graphene films, where the fibrils are deposited, reduce the background level of the STEM <span class="hlt">images</span> compared with instances when conventional amorphous carbon films are used. Using 30keV, which is lower than that for conventional TEM (100-300keV), together with low detection angles (15-55mrad) enhances the signals from the fibrils. These factors improve <span class="hlt">image</span> quality, which enables <span class="hlt">observation</span> of thin fibrils with widths of 7-8nm. STEM <span class="hlt">imaging</span> clearly reveals a twisted-ribbon structure of a fibril, and SE <span class="hlt">imaging</span> shows an emphasized striped pattern of the fibril. The LV-STEM in SEM enables acquisition of two types of <span class="hlt">images</span> of an identical fibril in a single instrument, which is useful for understanding the structure. This study expands the application of SEM to other systems of interest, which is beneficial to a large number of users. The method in this study can be applied to the <span class="hlt">observation</span> of various nanobio materials and analysis of their native structures, thus contributing to research in materials and life sciences. Copyright © 2017 Elsevier Ltd. All rights reserved.</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://adsabs.harvard.edu/abs/2015AGUFMIN41A1682G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMIN41A1682G"><span>Registering parameters and granules of wave <span class="hlt">observations</span>: <span class="hlt">IMAGE</span> RPI success story</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Galkin, I. A.; Charisi, A.; Fung, S. F.; Benson, R. F.; Reinisch, B. W.</p> <p>2015-12-01</p> <p>Modern metadata systems strive to help scientists locate data relevant to their research and then retrieve them quickly. Success of this mission depends on the organization and completeness of metadata. Each relevant data resource has to be registered; each content has to be described; each data file has to be accessible. Ultimately, data discoverability is about the practical ability to describe data content and location. Correspondingly, data registration has a "Parameter" level, at which content is specified by listing available <span class="hlt">observed</span> properties (parameters), and a "Granule" level, at which download links are given to data records (granules). Until recently, both parameter- and granule-level data registrations were accomplished at NASA Virtual System Observatory easily by listing provided parameters and building Granule documents with URLs to the datafile locations, usually those at NASA CDAWeb data warehouse. With the introduction of the Virtual Wave Observatory (VWO), however, the parameter/granule concept faced a scalability challenge. The wave phenomenon content is rich with descriptors of the wave generation, propagation, interaction with propagation media, and <span class="hlt">observation</span> processes. Additionally, the wave phenomenon content varies from record to record, reflecting changes in the constituent processes, making it necessary to generate granule documents at sub-minute resolution. We will present the first success story of registering 234,178 records of <span class="hlt">IMAGE</span> Radio Plasma <span class="hlt">Imager</span> (RPI) plasmagram data and Level 2 derived data products in ESPAS (near-Earth Space Data Infrastructure for e-Science), using the VWO-inspired wave ontology. The granules are arranged in overlapping display and numerical data collections. Display data include (a) auto-prospected plasmagrams of potential interest, (b) interesting plasmagrams annotated by human analysts or software, and (c) spectacular plasmagrams annotated by analysts as publication-quality examples of the RPI science</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930028542&hterms=SCI&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DSCI','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930028542&hterms=SCI&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DSCI"><span>Ultraviolet <span class="hlt">Imaging</span> Telescope <span class="hlt">observations</span> of the ScI galaxy NGC 628 (M74)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chen, Peter C.; Cornett, Robert H.; Roberts, Morton S.; Bohlin, Ralph C.; Neff, Susan G.; O'Connell, Robert W.; Parise, Ronald A.; Smith, Andrew M.; Stecher, Theodore P.</p> <p>1992-01-01</p> <p>Ultraviolet <span class="hlt">images</span> of NGC 628 at 1520 and 2490 A show that the nucleus has an oblong appearance and that the arms and disk exhibit features not seen in blue or H-alpha <span class="hlt">images</span>. Aperture photometry of the nucleus gives results that are compatible with <span class="hlt">observations</span> in other bandpasses and with models. The spiral arms appear more symmetrical in the UV than in other colors; in particular, two gaps are seen on either side of the nucleus. Combined UV and radio data appear to support a large-scale collective phenomenon, perhaps a quasi-static spiral structure mechanism, as being the dominant mode of spiral formation in this galaxy. We report the detection of a low surface brightness object at a distance of 7.6 arcmin southwest of the nucleus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AJ....153...71F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AJ....153...71F"><span>The Kepler Follow-up <span class="hlt">Observation</span> Program. I. A Catalog of Companions to Kepler Stars from High-Resolution <span class="hlt">Imaging</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Furlan, E.; Ciardi, D. R.; Everett, M. E.; Saylors, M.; Teske, J. K.; Horch, E. P.; Howell, S. B.; van Belle, G. T.; Hirsch, L. A.; Gautier, T. N., III; Adams, E. R.; Barrado, D.; Cartier, K. M. S.; Dressing, C. D.; Dupree, A. K.; Gilliland, R. L.; Lillo-Box, J.; Lucas, P. W.; Wang, J.</p> <p>2017-02-01</p> <p>We present results from high-resolution, optical to near-IR <span class="hlt">imaging</span> of host stars of Kepler Objects of Interest (KOIs), identified in the original Kepler field. Part of the data were obtained under the Kepler <span class="hlt">imaging</span> follow-up <span class="hlt">observation</span> program over six years (2009-2015). Almost 90% of stars that are hosts to planet candidates or confirmed planets were <span class="hlt">observed</span>. We combine measurements of companions to KOI host stars from different bands to create a comprehensive catalog of projected separations, position angles, and magnitude differences for all detected companion stars (some of which may not be bound). Our compilation includes 2297 companions around 1903 primary stars. From high-resolution <span class="hlt">imaging</span>, we find that ˜10% (˜30%) of the <span class="hlt">observed</span> stars have at least one companion detected within 1″ (4″). The true fraction of systems with close (≲4″) companions is larger than the <span class="hlt">observed</span> one due to the limited sensitivities of the <span class="hlt">imaging</span> data. We derive correction factors for planet radii caused by the dilution of the transit depth: assuming that planets orbit the primary stars or the brightest companion stars, the average correction factors are 1.06 and 3.09, respectively. The true effect of transit dilution lies in between these two cases and varies with each system. Applying these factors to planet radii decreases the number of KOI planets with radii smaller than 2 {R}\\oplus by ˜2%-23% and thus affects planet occurrence rates. This effect will also be important for the yield of small planets from future transit missions such as TESS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4316545','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4316545"><span>Mirror <span class="hlt">Observation</span> of Finger Action Enhances Activity in Anterior Intraparietal Sulcus: A Functional Magnetic Resonance <span class="hlt">Imaging</span> Study</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Murayama, Takashi; Takasugi, Jun; Monma, Masahiko; Oga, Masaru</p> <p>2013-01-01</p> <p>Mirror therapy can be used to promote recovery from paralysis in patients with post-stroke hemiplegia, There are a lot of reports that mirror-<span class="hlt">image</span> <span class="hlt">observation</span> of the unilateral moving hand enhanced the excitability of the primary motor area (M1) ipsilateral to the moving hand in healthy subjects. but the neural mechanisms underlying its therapeutic effects are currently unclear. To investigate this issue, we used functional magnetic resonance <span class="hlt">imaging</span> to measure activity in brain regions related to visual information processing during mirror <span class="hlt">image</span> movement <span class="hlt">observation</span>. Thirteen healthy subjects performed a finger-thumb opposition task with the left and right hands separately, with or without access to mirror <span class="hlt">observation</span>. In the mirror condition, one hand was reflected in a mirror placed above the abdomen in the MRI scanner. In the masked mirror condition, subjects performed the same task but with the mirror obscured. In both conditions, the other hand was held at rest behind the mirror. A between-task comparison (mirror versus masked mirror) revealed significant activation in the ipsilateral hemisphere in the anterior intraparietal sulcus (aIP) while performing all tasks, regardless of which hand was used. The right aIP was significantly activated while moving the right hand. In contrast, in the left aIP, a small number of voxels showed a tendency toward activation during both left and right hand movement. The enhancement of ipsilateral aIP activity by the mirror <span class="hlt">image</span> <span class="hlt">observation</span> of finger action suggests that bimodal aIP neurons can be activated by visual information. We propose that activation in the M1 ipsilateral to the moving hand can be induced by information passing through the ventral premotor area from the aIP. PMID:25792898</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRD..123..276H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRD..123..276H"><span><span class="hlt">Observations</span> of the Breakdown of Mountain Waves Over the Andes Lidar Observatory at Cerro Pachon on 8/9 July 2012</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hecht, J. H.; Fritts, D. C.; Wang, L.; Gelinas, L. J.; Rudy, R. J.; Walterscheid, R. L.; Taylor, M. J.; Pautet, P. D.; Smith, S.; Franke, S. J.</p> <p>2018-01-01</p> <p>Although mountain waves (MWs) are thought to be a ubiquitous feature of the wintertime southern Andes stratosphere, it was not known whether these waves propagated up to the mesopause region until Smith et al. (2009) confirmed their presence via <span class="hlt">airglow</span> <span class="hlt">observations</span>. The new Andes Lidar Observatory at Cerro Pachon in Chile provided the opportunity for a further study of these waves. Since MWs have near-zero phase speed, and zero wind lines often occur in the winter upper mesosphere (80 to 100 km altitude) region due to the reversal of the zonal mean and tidal wind, MW breakdown may routinely occur at these altitudes. Here we report on very high spatial/temporal resolution <span class="hlt">observations</span> of the initiation of MW breakdown in the mesopause region. Because the waves are nearly stationary, the breakdown process was <span class="hlt">observed</span> over several hours; a much longer interval than has previously been <span class="hlt">observed</span> for any gravity wave breakdown. During the breakdown process <span class="hlt">observations</span> were made of initial horseshoe-shaped vortices, leading to successive vortex rings, as is also commonly seen in Direct Numerical Simulations (DNS) of idealized and multiscale gravity wave breaking. Kelvin-Helmholtz instability (KHI) structures were also <span class="hlt">observed</span> to form. Comparing the structure of <span class="hlt">observed</span> KHI with the results of existing DNS allowed an estimate of the turbulent kinematic viscosity. This viscosity was found to be around 25 m2/s, a value larger than the nominal viscosity that is used in models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27279782','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27279782"><span>Low-radio-frequency eclipses of the redback pulsar J2215+5135 <span class="hlt">observed</span> in the <span class="hlt">image</span> plane with LOFAR.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Broderick, J W; Fender, R P; Breton, R P; Stewart, A J; Rowlinson, A; Swinbank, J D; Hessels, J W T; Staley, T D; van der Horst, A J; Bell, M E; Carbone, D; Cendes, Y; Corbel, S; Eislöffel, J; Falcke, H; Grießmeier, J-M; Hassall, T E; Jonker, P; Kramer, M; Kuniyoshi, M; Law, C J; Markoff, S; Molenaar, G J; Pietka, M; Scheers, L H A; Serylak, M; Stappers, B W; Ter Veen, S; van Leeuwen, J; Wijers, R A M J; Wijnands, R; Wise, M W; Zarka, P</p> <p>2016-07-01</p> <p>The eclipses of certain types of binary millisecond pulsars (i.e. 'black widows' and 'redbacks') are often studied using high-time-resolution, 'beamformed' radio <span class="hlt">observations</span>. However, they may also be detected in <span class="hlt">images</span> generated from interferometric data. As part of a larger <span class="hlt">imaging</span> project to characterize the variable and transient sky at radio frequencies <200 MHz, we have blindly detected the redback system PSR J2215+5135 as a variable source of interest with the Low-Frequency Array (LOFAR). Using <span class="hlt">observations</span> with cadences of two weeks - six months, we find preliminary evidence that the eclipse duration is frequency dependent (∝ν -0.4 ), such that the pulsar is eclipsed for longer at lower frequencies, in broad agreement with beamformed studies of other similar sources. Furthermore, the detection of the eclipses in <span class="hlt">imaging</span> data suggests an eclipsing medium that absorbs the pulsed emission, rather than scattering it. Our study is also a demonstration of the prospects of finding pulsars in wide-field <span class="hlt">imaging</span> surveys with the current generation of low-frequency radio telescopes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150003281','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150003281"><span>Stroboscopic <span class="hlt">Image</span> Modulation to Reduce the Visual Blur of an Object Being Viewed by an <span class="hlt">Observer</span> Experiencing Vibration</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kaiser, Mary K. (Inventor); Adelstein, Bernard D. (Inventor); Anderson, Mark R. (Inventor); Beutter, Brent R. (Inventor); Ahumada, Albert J., Jr. (Inventor); McCann, Robert S. (Inventor)</p> <p>2014-01-01</p> <p>A method and apparatus for reducing the visual blur of an object being viewed by an <span class="hlt">observer</span> experiencing vibration. In various embodiments of the present invention, the visual blur is reduced through stroboscopic <span class="hlt">image</span> modulation (SIM). A SIM device is operated in an alternating "on/off" temporal pattern according to a SIM drive signal (SDS) derived from the vibration being experienced by the <span class="hlt">observer</span>. A SIM device (controlled by a SIM control system) operates according to the SDS serves to reduce visual blur by "freezing" (or reducing an <span class="hlt">image</span>'s motion to a slow drift) the visual <span class="hlt">image</span> of the viewed object. In various embodiments, the SIM device is selected from the group consisting of illuminator(s), shutter(s), display control system(s), and combinations of the foregoing (including the use of multiple illuminators, shutters, and display control systems).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001JASS...18...21W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001JASS...18...21W"><span>Studies of Gravity Waves Using Michelson Interferometer Measurements of OH (3-1) Bands</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Won, Young-In; Cho, Young-Min; Lee, Bang Yong; Kim, J.</p> <p>2001-06-01</p> <p>As part of a long-term program for polar upper atmospheric studies, temperatures and intensities of the OH (3-1) bands were derived from spectrometric <span class="hlt">observations</span> of <span class="hlt">airglow</span> emissions over King Sejong station (62.22o S, 301.25o E). These measurements were made with a Michelson interferometer to cover wavelength regions between 1000 nm and 2000 nm. A spectral analysis was performed to individual nights of data to acquire information on the waves in the upper mesosphere/lower thermosphere. It is assumed that the measured fluctuations in the intensity and temperature of the OH (3-1) <span class="hlt">airglow</span> were caused by gravity waves propagating through the emission layer. Correlation of intensity and temperature variation revealed oscillations with periods ranging from 2 to 9 hours. We also calculated Krassovsky's parameter and compared with published values.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AAS...23144203J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AAS...23144203J"><span>Wide-Field <span class="hlt">Imaging</span> Telescope-0 (WIT0) with automatic <span class="hlt">observing</span> system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ji, Tae-Geun; Byeon, Seoyeon; Lee, Hye-In; Park, Woojin; Lee, Sang-Yun; Hwang, Sungyong; Choi, Changsu; Gibson, Coyne Andrew; Kuehne, John W.; Prochaska, Travis; Marshall, Jennifer L.; Im, Myungshin; Pak, Soojong</p> <p>2018-01-01</p> <p>We introduce Wide-Field <span class="hlt">Imaging</span> Telescope-0 (WIT0), with an automatic <span class="hlt">observing</span> system. It is developed for monitoring the variabilities of many sources at a time, e.g. young stellar objects and active galactic nuclei. It can also find the locations of transient sources such as a supernova or gamma-ray bursts. In 2017 February, we installed the wide-field 10-inch telescope (Takahashi CCA-250) as a piggyback system on the 30-inch telescope at the McDonald Observatory in Texas, US. The 10-inch telescope has a 2.35 × 2.35 deg field-of-view with a 4k × 4k CCD Camera (FLI ML16803). To improve the <span class="hlt">observational</span> efficiency of the system, we developed a new automatic <span class="hlt">observing</span> software, KAOS30 (KHU Automatic <span class="hlt">Observing</span> Software for McDonald 30-inch telescope), which was developed by Visual C++ on the basis of a windows operating system. The software consists of four control packages: the Telescope Control Package (TCP), the Data Acquisition Package (DAP), the Auto Focus Package (AFP), and the Script Mode Package (SMP). Since it also supports the instruments that are using the ASCOM driver, the additional hardware installations become quite simplified. We commissioned KAOS30 in 2017 August and are in the process of testing. Based on the WIT0 experiences, we will extend KAOS30 to control multiple telescopes in future projects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950035282&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dtwilight','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950035282&hterms=twilight&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dtwilight"><span>Recent <span class="hlt">observations</span> of the OI 8446 A emission over Millstone Hill</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lancaster, R. S.; Kerr, R. B.; Ng, K.; Noto, J.; Franco, M.; Solomon, Stanley C.</p> <p>1994-01-01</p> <p>Evening twilight spectra of the OI 8446 A emission were obtained during May and June of 1993 using a single-etalon, pressure scanning, Fabry-Perot interferometer located in the Millstone Hill Optical Facility. The goals of this work are to positively identify the 8446 A emission in the twilight <span class="hlt">airglow</span> and to determine the intensity decay as a function of solar depression angle. Also, a study of the relative triplet line strengths is performed in hopes of establishing the importance of the primary excitation mechanisms (photoelectron impact or Bowen fluorescence) during the twilight period. Although absent in most of the data, a distinct auroral influence is also found to contribute considerably, on occasion, to the emission over Millstone Hill. The ratio of the combined 8446.26 A and 8446.38 A intensities to the 8446.76 A intensity varies as 0.13 +/- 0.03 per degree of solar depression angle, indicating that secondary excitation mechanisms are becoming increasingly important as evening twilight progresses. Bowen fluorescence is not found to be the primary excitation mechanism at any time during twilight, contributing just a few Rayleighs at most. These <span class="hlt">observations</span> are an important first step toward a better characterization of highly variable thermospheric oxygen concentrations through ground-based measurements of the OI 8446 A emission.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10136E..0NP','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10136E..0NP"><span>Real space channelization for generic DBT system <span class="hlt">image</span> quality evaluation with channelized Hotelling <span class="hlt">observer</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Petrov, Dimitar; Cockmartin, Lesley; Marshall, Nicholas; Vancoillie, Liesbeth; Young, Kenneth; Bosmans, Hilde</p> <p>2017-03-01</p> <p>Digital breast tomosynthesis (DBT) is a relatively new 3D mammography technique that promises better detection of low contrast masses than conventional 2D mammography. The parameter space for DBT is large however and finding an optimal balance between dose and <span class="hlt">image</span> quality remains challenging. Given the large number of conditions and <span class="hlt">images</span> required in optimization studies, the use of human <span class="hlt">observers</span> (HO) is time consuming and certainly not feasible for the tuning of all degrees of freedom. Our goal was to develop a model <span class="hlt">observer</span> (MO) that could predict human detectability for clinically relevant details embedded within a newly developed structured phantom for DBT applications. DBT series were acquired on GE SenoClaire 3D, Giotto Class, Fujifilm AMULET Innovality and Philips MicroDose systems at different dose levels, Siemens Inspiration DBT acquisitions were reconstructed with different algorithms, while a larger set of DBT series was acquired on Hologic Dimensions system for first reproducibility testing. A channelized Hotelling <span class="hlt">observer</span> (CHO) with Gabor channels was developed The parameters of the Gabor channels were tuned on all systems at standard scanning conditions and the candidate that produced the best fit for all systems was chosen. After tuning, the MO was applied to all systems and conditions. Linear regression lines between MO and HO scores were calculated, giving correlation coefficients between 0.87 and 0.99 for all tested conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990019629&hterms=doolittle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3Ddoolittle','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990019629&hterms=doolittle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D10%26Ntt%3Ddoolittle"><span>Conjugate <span class="hlt">Observations</span> of Optical Aurora with POLAR Satellite and Ground Based <span class="hlt">Imagers</span> in Antarctica</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mende, S. H.; Frey, H.; Vo, H.; Geller, S. P.; Doolittle, J. H.; Spann, J. F., Jr.</p> <p>1998-01-01</p> <p>Operation of the ultraviolet <span class="hlt">imager</span> on the POLAR satellite permits the <span class="hlt">observation</span> of Aurora Borealis in daylight during northern summer. With optical <span class="hlt">imagers</span> in the Automatic Geophysical Observatories (AGO-s) large regions of the oval of Aurora Australis can be <span class="hlt">observed</span> simultaneously during the southern winter polar night. This opportunity permits conducting a systematic study of the properties of auroras on opposite ends of the same field line. It is expected that simultaneously <span class="hlt">observed</span> conjugate auroras occurring on closed field lines should be similar to each other in appearance because of the close connection between the two hemispheres through particle scattering and mirroring processes. On open or greatly distorted field lines there is no a priori expectation of similarity between conjugate auroras. To investigate the influence of different IMF conditions on auroral behavior we have examined conjugate data for periods of southward IMF. Sudden brightening and subsequent poleward expansions are <span class="hlt">observed</span> to occur simultaneously in both hemispheres. The POLAR data show that sudden brightening are initiated at various local time regions. When the local time of this region is in the field of view of the AGO station network then corresponding brightening is also found to occur in the southern hemisphere. Large features such as substorm induced westward propagation and resulting auroral brightening seem to occur simultaneously on conjugate hemispheres. The widely different view scales make it difficult to make unique identification of individual auroral forms in the POLAR and in the ground based data but in a general sense the data is consistent with conjugate behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030054367&hterms=coulomb+law&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcoulomb%2Blaw','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030054367&hterms=coulomb+law&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dcoulomb%2Blaw"><span>Theoretical Model <span class="hlt">Images</span> and Spectra for Comparison with HESSI and Microwave <span class="hlt">Observations</span> of Solar Flares</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fisher, Richard R. (Technical Monitor); Holman, G. D.; Sui, L.; McTiernan, J. M.; Petrosian, V.</p> <p>2003-01-01</p> <p>We have computed bremsstrahlung and gyrosynchrotron <span class="hlt">images</span> and spectra from a model flare loop. Electrons with a power-law energy distribution are continuously injected at the top of a semi-circular magnetic loop. The Fokker-Planck equation is integrated to obtain the steady-state electron distribution throughout the loop. Coulomb scattering and energy losses and magnetic mirroring are included in the model. The resulting electron distributions are used to compute the radiative emissions. Sample <span class="hlt">images</span> and spectra are presented. We are developing these models for the interpretation of the High Energy Solar Spectroscopic <span class="hlt">Imager</span> (HESSI) x-ray/gamma ray data and coordinated microwave <span class="hlt">observations</span>. The Fokker-Planck and radiation codes are available on the Web at http://hesperia.gsfc.nasa.gov/hessi/modelware.htm This work is supported in part by the NASA Sun-Earth Connection Program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4352060','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4352060"><span>Expanding <span class="hlt">Imaging</span> Capabilities for Microfluidics: Applicability of Darkfield Internal Reflection Illumination (DIRI) to <span class="hlt">Observations</span> in Microfluidics</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Kawano, Yoshihiro; Otsuka, Chino; Sanzo, James; Higgins, Christopher; Nirei, Tatsuo; Schilling, Tobias; Ishikawa, Takuji</p> <p>2015-01-01</p> <p>Microfluidics is used increasingly for engineering and biomedical applications due to recent advances in microfabrication technologies. Visualization of bubbles, tracer particles, and cells in a microfluidic device is important for designing a device and analyzing results. However, with conventional methods, it is difficult to <span class="hlt">observe</span> the channel geometry and such particles simultaneously. To overcome this limitation, we developed a Darkfield Internal Reflection Illumination (DIRI) system that improved the drawbacks of a conventional darkfield illuminator. This study was performed to investigate its utility in the field of microfluidics. The results showed that the developed system could clearly visualize both microbubbles and the channel wall by utilizing brightfield and DIRI illumination simultaneously. The methodology is useful not only for static phenomena, such as clogging, but also for dynamic phenomena, such as the detection of bubbles flowing in a channel. The system was also applied to simultaneous fluorescence and DIRI <span class="hlt">imaging</span>. Fluorescent tracer beads and channel walls were <span class="hlt">observed</span> clearly, which may be an advantage for future microparticle <span class="hlt">image</span> velocimetry (μPIV) analysis, especially near a wall. Two types of cell stained with different colors, and the channel wall, can be recognized using the combined confocal and DIRI system. Whole-slide <span class="hlt">imaging</span> was also conducted successfully using this system. The tiling function significantly expands the <span class="hlt">observing</span> area of microfluidics. The developed system will be useful for a wide variety of engineering and biomedical applications for the growing field of microfluidics. PMID:25748425</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25748425','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25748425"><span>Expanding <span class="hlt">imaging</span> capabilities for microfluidics: applicability of darkfield internal reflection illumination (DIRI) to <span class="hlt">observations</span> in microfluidics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kawano, Yoshihiro; Otsuka, Chino; Sanzo, James; Higgins, Christopher; Nirei, Tatsuo; Schilling, Tobias; Ishikawa, Takuji</p> <p>2015-01-01</p> <p>Microfluidics is used increasingly for engineering and biomedical applications due to recent advances in microfabrication technologies. Visualization of bubbles, tracer particles, and cells in a microfluidic device is important for designing a device and analyzing results. However, with conventional methods, it is difficult to <span class="hlt">observe</span> the channel geometry and such particles simultaneously. To overcome this limitation, we developed a Darkfield Internal Reflection Illumination (DIRI) system that improved the drawbacks of a conventional darkfield illuminator. This study was performed to investigate its utility in the field of microfluidics. The results showed that the developed system could clearly visualize both microbubbles and the channel wall by utilizing brightfield and DIRI illumination simultaneously. The methodology is useful not only for static phenomena, such as clogging, but also for dynamic phenomena, such as the detection of bubbles flowing in a channel. The system was also applied to simultaneous fluorescence and DIRI <span class="hlt">imaging</span>. Fluorescent tracer beads and channel walls were <span class="hlt">observed</span> clearly, which may be an advantage for future microparticle <span class="hlt">image</span> velocimetry (μPIV) analysis, especially near a wall. Two types of cell stained with different colors, and the channel wall, can be recognized using the combined confocal and DIRI system. Whole-slide <span class="hlt">imaging</span> was also conducted successfully using this system. The tiling function significantly expands the <span class="hlt">observing</span> area of microfluidics. The developed system will be useful for a wide variety of engineering and biomedical applications for the growing field of microfluidics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22649059-su-clinically-observed-discrepancy-between-image-based-log-based-mlc-position','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22649059-su-clinically-observed-discrepancy-between-image-based-log-based-mlc-position"><span>SU-F-T-469: A Clinically <span class="hlt">Observed</span> Discrepancy Between <span class="hlt">Image</span>-Based and Log- Based MLC Position</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>Neal, B; Ahmed, M; Siebers, J</p> <p>2016-06-15</p> <p>Purpose: To present a clinical case which challenges the base assumption of log-file based QA, by showing that the actual position of a MLC leaf can suddenly deviate from its programmed and logged position by >1 mm as <span class="hlt">observed</span> with real-time <span class="hlt">imaging</span>. Methods: An EPID-based exit-fluence dosimetry system designed to prevent gross delivery errors was used in cine mode to capture portal <span class="hlt">images</span> during treatment. Visual monitoring identified an anomalous MLC leaf pair gap not otherwise detected by the automatic position verification. The position of the erred leaf was measured on EPID <span class="hlt">images</span> and log files were analyzed for themore » treatment in question, the prior day’s treatment, and for daily MLC test patterns acquired on those treatment days. Additional standard test patterns were used to quantify the leaf position. Results: Whereas the log file reported no difference between planned and recorded positions, <span class="hlt">image</span>-based measurements showed the leaf to be 1.3±0.1 mm medial from the planned position. This offset was confirmed with the test pattern irradiations. Conclusion: It has been clinically <span class="hlt">observed</span> that log-file derived leaf positions can differ from their actual positions by >1 mm, and therefore cannot be considered to be the actual leaf positions. This cautions the use of log-based methods for MLC or patient quality assurance without independent confirmation of log integrity. Frequent verification of MLC positions through independent means is a necessary precondition to trusting log file records. Intra-treatment EPID <span class="hlt">imaging</span> provides a method to capture departures from MLC planned positions. Work was supported in part by Varian Medical Systems.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950048242&hterms=Petit&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DPetit','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950048242&hterms=Petit&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DPetit"><span>Ultraviolet <span class="hlt">imaging</span> telescope and optical emission-line <span class="hlt">observations</span> of H II regions in M81</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hill, Jesse K.; Cheng, K.-P.; Bohlin, Ralph C.; Cornett, Robert H.; Hintzen, P. M. N.; O'Connell, Robert W.; Roberts, Morton S.; Smith, Andrew M.; Smith, Eric P.; Stecher, Theodore P.</p> <p>1995-01-01</p> <p><span class="hlt">Images</span> of the type Sab spiral galaxy M81 were obtained in far-UV and near-UV bands by the Ultraviolet <span class="hlt">Imaging</span> Telescope (UIT) during the Astro-1 Spacelab mission of 1990 December. Magnitudes in the two UV bands are determined for 52 H II regions from the catalog of Petit, Sivan, & Karachentsev (1988). Fluxes of the H-alpha and H-beta emission lines are determined from CCD <span class="hlt">images</span>. Extinctions for the brightest H II regions are determined from <span class="hlt">observed</span> Balmer decrements. Fainter H II regions are assigned the average of published radio-H-alpha extinctions for several bright H II regions. The radiative transfer models of Witt, Thronson, & Capuano (1992) are shown to predict a relationship between Balmer Decrement and H-alpha extinction consistent with <span class="hlt">observed</span> line and radio fluxes for the brightest 7 H II regions and are used to estimate the UV extinction. Ratios of Lyman continuum with ratios predicted by model spectra computed for initial mass function (IMF) slope equal to -1.0 and stellar masses ranging from 5 to 120 solar mass. Ages and masses are estimated by comparing the H-alpha and far-UV fluxes and their ratio with the models. The total of the estimated stellar masses for the 52 H II regions is 1.4 x 10(exp 5) solar mass. The star-formation rate inferred for M81 from the <span class="hlt">observed</span> UV and H-alpha fluxes is low for a spiral galaxy at approximately 0.13 solar mass/yr, but consistent with the low star-formation rates obtained by Kennicutt (1983) and Caldwell et al. (1991) for early-type spirals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010100545&hterms=fourier&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dfourier','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010100545&hterms=fourier&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dfourier"><span>Demonstration of <span class="hlt">Imaging</span> Fourier Transform Spectrometer (FTS) Performance for Planetary and Geostationary Earth <span class="hlt">Observing</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Revercomb, Henry E.; Sromovsky, Lawrence A.; Fry, Patrick M.; Best, Fred A.; LaPorte, Daniel D.</p> <p>2001-01-01</p> <p>The combination of massively parallel spatial sampling and accurate spectral radiometry offered by <span class="hlt">imaging</span> FTS makes it extremely attractive for earth and planetary remote sensing. We constructed a breadboard instrument to help assess the potential for planetary applications of small <span class="hlt">imaging</span> FTS instruments in the 1 - 5 micrometer range. The results also support definition of the NASA Geostationary <span class="hlt">Imaging</span> FTS (GIFTS) instrument that will make key meteorological and climate <span class="hlt">observations</span> from geostationary earth orbit. The Planetary <span class="hlt">Imaging</span> FTS (PIFTS) breadboard is based on a custom miniaturized Bomen interferometer that uses corner cube reflectors, a wishbone pivoting voice-coil delay scan mechanism, and a laser diode metrology system. The interferometer optical output is measured by a commercial infrared camera procured from Santa Barbara Focalplane. It uses an InSb 128x128 detector array that covers the entire FOV of the instrument when coupled with a 25 mm focal length commercial camera lens. With appropriate lenses and cold filters the instrument can be used from the visible to 5 micrometers. The delay scan is continuous, but slow, covering the maximum range of +/- 0.4 cm in 37.56 sec at a rate of 500 <span class="hlt">image</span> frames per second. <span class="hlt">Image</span> exposures are timed to be centered around predicted zero crossings. The design allows for prediction algorithms that account for the most recent fringe rate so that timing jitter produced by scan speed variations can be minimized. Response to a fixed source is linear with exposure time nearly to the point of saturation. Linearity with respect to input variations was demonstrated to within 0.16% using a 3-point blackbody calibration. <span class="hlt">Imaging</span> of external complex scenes was carried out at low and high spectral resolution. These require full complex calibration to remove background contributions that vary dramatically over the instrument FOV. Testing is continuing to demonstrate the precise radiometric accuracy and noise characteristics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015DPS....4741915S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015DPS....4741915S"><span>New <span class="hlt">Observations</span> of Molecular Nitrogen by the <span class="hlt">Imaging</span> Ultraviolet Spectrograph on MAVEN</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stevens, Michael H.; Evans, J. S.; Schneider, Nicholas M.; Stewart, A. I. F.; Deighan, Justin; Jain, Sonal K.; Crismani, Matteo M. J.; Stiepen, Arnaud; Chaffin, Michael S.; McClintock, William E.; Holsclaw, Greg M.; Lefevre, Franck; Montmessin, Franck; Lo, Daniel Y.; Clarke, John T.; Bougher, Stephen W.; Jakosky, Bruce M.</p> <p>2015-11-01</p> <p>The Martian ultraviolet dayglow provides information on the basic state of the Martian upper atmosphere. The <span class="hlt">Imaging</span> Ultraviolet Spectrograph (IUVS) on NASA’s Mars Atmosphere and Volatile Evolution (MAVEN) mission has <span class="hlt">observed</span> Mars at mid and far-UV wavelengths since its arrival in September 2014. In this work, we describe a linear regression method used to extract components of UV spectra from IUVS limb <span class="hlt">observations</span> and focus in particular on molecular nitrogen (N2) photoelectron excited emissions. We identify N2 Lyman-Birge-Hopfield (LBH) emissions for the first time at Mars and we also confirm the tentative identification of N2 Vegard-Kaplan (VK) emissions. We compare <span class="hlt">observed</span> VK and LBH limb radiance profiles to model results between 90 and 210 km. Finally, we compare retrieved N2 density profiles to general circulation (GCM) model results. Contrary to earlier analyses using other satellite data that indicated N2 densities were a factor of three less than predictions, we find that N2 abundances exceed GCM results by about a factor of two at 130 km but are in agreement at 150 km.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150002681','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150002681"><span>Terrestrial Myriametric Radio Burst <span class="hlt">Observed</span> by <span class="hlt">IMAGE</span> and Geotail Satellites</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fung, Shing F.; Hashimoto, KoZo; Kojima, Hirotsugu; Boardson, Scott A.; Garcia, Leonard N.; Matsumoto, Hiroshi; Green, James L.; Reinisch, Bodo W.</p> <p>2013-01-01</p> <p>We report the simultaneous detection of a terrestrial myriametric radio burst (TMRB) by <span class="hlt">IMAGE</span> and Geotail on 19 August 2001. The TMRB was confined in time (0830-1006 UT) and frequency (12-50kHz). Comparisons with all known nonthermal myriametric radiation components reveal that the TMRB might be a distinct radiation with a source that is unrelated to the previously known radiation. Considerations of beaming from spin-modulation analysis and <span class="hlt">observing</span> satellite and source locations suggest that the TMRB may have a fan beamlike radiation pattern emitted by a discrete, dayside source located along the poleward edge of magnetospheric cusp field lines. TMRB responsiveness to IMF Bz and By orientations suggests that a possible source of the TMRB could be due to dayside magnetic reconnection instigated by northward interplanetary field condition.</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>