Sample records for vertically pointing radars

  1. Retrieving Vertical Air Motion and Raindrop Size Distributions from Vertically Pointing Doppler Radars

    NASA Astrophysics Data System (ADS)

    Williams, C. R.; Chandra, C. V.

    2017-12-01

    The vertical evolution of falling raindrops is a result of evaporation, breakup, and coalescence acting upon those raindrops. Computing these processes using vertically pointing radar observations is a two-step process. First, the raindrop size distribution (DSD) and vertical air motion need to be estimated throughout the rain shaft. Then, the changes in DSD properties need to be quantified as a function of height. The change in liquid water content is a measure of evaporation, and the change in raindrop number concentration and size are indicators of net breakup or coalescence in the vertical column. The DSD and air motion can be retrieved using observations from two vertically pointing radars operating side-by-side and at two different wavelengths. While both radars are observing the same raindrop distribution, they measure different reflectivity and radial velocities due to Rayleigh and Mie scattering properties. As long as raindrops with diameters greater than approximately 2 mm are in the radar pulse volumes, the Rayleigh and Mie scattering signatures are unique enough to estimate DSD parameters using radars operating at 3- and 35-GHz (Williams et al. 2016). Vertical decomposition diagrams (Williams 2016) are used to explore the processes acting on the raindrops. Specifically, changes in liquid water content with height quantify evaporation or accretion. When the raindrops are not evaporating, net raindrop breakup and coalescence are identified by changes in the total number of raindrops and changes in the DSD effective shape as the raindrops. This presentation will focus on describing the DSD and air motion retrieval method using vertical profiling radar observations from the Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) Southern Great Plains (SGP) central facility in Northern Oklahoma.

  2. Normalized vertical ice mass flux profiles from vertically pointing 8-mm-wavelength Doppler radar

    NASA Technical Reports Server (NTRS)

    Orr, Brad W.; Kropfli, Robert A.

    1993-01-01

    During the FIRE 2 (First International Satellite Cloud Climatology Project Regional Experiment) project, NOAA's Wave Propagation Laboratory (WPL) operated its 8-mm wavelength Doppler radar extensively in the vertically pointing mode. This allowed for the calculation of a number of important cirrus cloud parameters, including cloud boundary statistics, cloud particle characteristic sizes and concentrations, and ice mass content (imc). The flux of imc, or, alternatively, ice mass flux (imf), is also an important parameter of a cirrus cloud system. Ice mass flux is important in the vertical redistribution of water substance and thus, in part, determines the cloud evolution. It is important for the development of cloud parameterizations to be able to define the essential physical characteristics of large populations of clouds in the simplest possible way. One method would be to normalize profiles of observed cloud properties, such as those mentioned above, in ways similar to those used in the convective boundary layer. The height then scales from 0.0 at cloud base to 1.0 at cloud top, and the measured cloud parameter scales by its maximum value so that all normalized profiles have 1.0 as their maximum value. The goal is that there will be a 'universal' shape to profiles of the normalized data. This idea was applied to estimates of imf calculated from data obtained by the WPL cloud radar during FIRE II. Other quantities such as median particle diameter, concentration, and ice mass content can also be estimated with this radar, and we expect to also examine normalized profiles of these quantities in time for the 1993 FIRE II meeting.

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

    NASA Technical Reports Server (NTRS)

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

    2014-01-01

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

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

    NASA Astrophysics Data System (ADS)

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

    2014-07-01

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

  5. Utilizing the Vertical Variability of Precipitation to Improve Radar QPE

    NASA Technical Reports Server (NTRS)

    Gatlin, Patrick N.; Petersen, Walter A.

    2016-01-01

    Characteristics of the melting layer and raindrop size distribution can be exploited to further improve radar quantitative precipitation estimation (QPE). Using dual-polarimetric radar and disdrometers, we found that the characteristic size of raindrops reaching the ground in stratiform precipitation often varies linearly with the depth of the melting layer. As a result, a radar rainfall estimator was formulated using D(sub m) that can be employed by polarimetric as well as dual-frequency radars (e.g., space-based radars such as the GPM DPR), to lower the bias and uncertainty of conventional single radar parameter rainfall estimates by as much as 20%. Polarimetric radar also suffers from issues associated with sampling the vertical distribution of precipitation. Hence, we characterized the vertical profile of polarimetric parameters (VP3)-a radar manifestation of the evolving size and shape of hydrometeors as they fall to the ground-on dual-polarimetric rainfall estimation. The VP3 revealed that the profile of ZDR in stratiform rainfall can bias dual-polarimetric rainfall estimators by as much as 50%, even after correction for the vertical profile of reflectivity (VPR). The VP3 correction technique that we developed can improve operational dual-polarimetric rainfall estimates by 13% beyond that offered by a VPR correction alone.

  6. Vertical Variability of Rain Drop Size Distribution from Micro Rain Radar Measurements during IFloodS

    NASA Astrophysics Data System (ADS)

    Adirosi, Elisa; Tokay, Ali; Roberto, Nicoletta; Gorgucci, Eugenio; Montopoli, Mario; Baldini, Luca

    2017-04-01

    Ground based weather radars are highly used to generate rainfall products for meteorological and hydrological applications. However, weather radar quantitative rainfall estimation is obtained at a certain altitude that depends mainly on the radar elevation angle and on the distance from the radar. Therefore, depending on the vertical variability of rainfall, a time-height ambiguity between radar measurement and rainfall at the ground can affect the rainfall products. The vertically pointing radars (such as the Micro Rain Radar, MRR) are great tool to investigate the vertical variability of rainfall and its characteristics and ultimately, to fill the gap between the ground level and the first available radar elevation. Furthermore, the knowledge of rain Drop Size Distribution (DSD) variability is linked to the well-known problem of the non-uniform beam filling that is one of the main uncertainties of Global Precipitation Measurement (GPM) mission Dual frequency Precipitation Radar (DPR). During GPM Ground Validation Iowa Flood Studies (IFloodS) field experiment, data collected with 2D video disdrometers (2DVD), Autonomous OTT Parsivel2 Units (APU), and MRR profilers at different sites were available. In three different sites co-located APU, 2DVD and MRR are available and covered by the S-band Dual Polarimetric Doppler radar (NPOL). The first elevation height of the radar beam varies, among the three sites, between 70 m and 1100 m. The IFloodS set-up has been used to compare disdrometers, MRR and NPOL data and to evaluate the uncertainties of those measurements. First, the performance of disdrometers and MRR in determining different rainfall parameters at ground has been evaluated and then the MRR based parameters have been compared with the ones obtained from NPOL data at the lowest elevations. Furthermore, the vertical variability of DSD and integral rainfall parameters within the MRR bins (from ground to 1085 m each 35 m) has been investigated in order to provide

  7. Doppler radar echoes of lightning and precipitation at vertical incidence

    NASA Technical Reports Server (NTRS)

    Zrnic, D. S.; Rust, W. D.; Taylor, W. L.

    1982-01-01

    Digital time series data at 16 heights within two storms were collected at vertical incidence with a 10-cm Doppler radar. On several occasions during data collection, lightning echoes were observed as increased reflectivity on an oscilloscope display. Simultaneously, lightning signals from nearby electric field change antennas were recorded on an analog recorder together with the radar echoes. Reflectivity, mean velocity, and Doppler spectra were examined by means of time series analysis for times during and after lightning discharges. Spectra from locations where lightning occurred show peaks, due to the motion of the lightning channel at the air speed. These peaks are considerably narrower than the ones due to precipitation. Besides indicating the vertical air velocity that can then be used to estimate hydrometeor-size distribution, the lightning spectra provide a convenient means to estimate the radar cross section of the channel. Subsequent to one discharge, we deduce that a rapid change in the orientation of hydrometeors occurred within the resolution volume.

  8. Estimating vertical velocity and radial flow from Doppler radar observations of tropical cyclones

    NASA Astrophysics Data System (ADS)

    Lee, J. L.; Lee, W. C.; MacDonald, A. E.

    2006-01-01

    The mesoscale vorticity method (MVM) is used in conjunction with the ground-based velocity track display (GBVTD) to derive the inner-core vertical velocity from Doppler radar observations of tropical cyclone (TC) Danny (1997). MVM derives the vertical velocity from vorticity variations in space and in time based on the mesoscale vorticity equation. The use of MVM and GBVTD allows us to derive good correlations among the eye-wall maximum wind, bow-shaped updraught and echo east of the eye-wall in Danny. Furthermore, we demonstrate the dynamically consistent radial flow can be derived from the vertical velocity obtained from MVM using the wind decomposition technique that solves the Poisson equations over a limited-area domain. With the wind decomposition, we combine the rotational wind which is obtained from Doppler radar wind observations and the divergent wind which is inferred dynamically from the rotational wind to form the balanced horizontal wind in TC inner cores, where rotational wind dominates the divergent wind. In this study, we show a realistic horizontal and vertical structure of the vertical velocity and the induced radial flow in Danny's inner core. In the horizontal, the main eye-wall updraught draws in significant surrounding air, converging at the strongest echo where the maximum updraught is located. In the vertical, the main updraught tilts vertically outwards, corresponding very well with the outward-tilting eye-wall. The maximum updraught is located at the inner edge of the eye-wall clouds, while downward motions are found at the outer edge. This study demonstrates that the mesoscale vorticity method can use high-temporal-resolution data observed by Doppler radars to derive realistic vertical velocity and the radial flow of TCs. The vorticity temporal variations crucial to the accuracy of the vorticity method have to be derived from a high-temporal-frequency observing system such as state-of-the-art Doppler radars.

  9. On the vertical resolution for near-nadir looking spaceborne rain radar

    NASA Astrophysics Data System (ADS)

    Kozu, Toshiaki

    A definition of radar resolution for an arbitrary direction is proposed and used to calculate the vertical resolution for a near-nadir looking spaceborne rain radar. Based on the calculation result, a scanning strategy is proposed which efficiently distributes the measurement time to each angle bin and thus increases the number of independent samples compared with a simple linear scanning.

  10. Quantitative precipitation estimation in complex orography using quasi-vertical profiles of dual polarization radar variables

    NASA Astrophysics Data System (ADS)

    Montopoli, Mario; Roberto, Nicoletta; Adirosi, Elisa; Gorgucci, Eugenio; Baldini, Luca

    2017-04-01

    Weather radars are nowadays a unique tool to estimate quantitatively the rain precipitation near the surface. This is an important task for a plenty of applications. For example, to feed hydrological models, mitigate the impact of severe storms at the ground using radar information in modern warning tools as well as aid the validation studies of satellite-based rain products. With respect to the latter application, several ground validation studies of the Global Precipitation Mission (GPM) products have recently highlighted the importance of accurate QPE from ground-based weather radars. To date, a plenty of works analyzed the performance of various QPE algorithms making use of actual and synthetic experiments, possibly trained by measurement of particle size distributions and electromagnetic models. Most of these studies support the use of dual polarization variables not only to ensure a good level of radar data quality but also as a direct input in the rain estimation equations. Among others, one of the most important limiting factors in radar QPE accuracy is the vertical variability of particle size distribution that affects at different levels, all the radar variables acquired as well as rain rates. This is particularly impactful in mountainous areas where the altitudes of the radar sampling is likely several hundred of meters above the surface. In this work, we analyze the impact of the vertical profile variations of rain precipitation on several dual polarization radar QPE algorithms when they are tested a in complex orography scenario. So far, in weather radar studies, more emphasis has been given to the extrapolation strategies that make use of the signature of the vertical profiles in terms of radar co-polar reflectivity. This may limit the use of the radar vertical profiles when dual polarization QPE algorithms are considered because in that case all the radar variables used in the rain estimation process should be consistently extrapolated at the surface

  11. Intercomparison of vertical structure of storms revealed by ground-based (NMQ) and spaceborne radars (CloudSat-CPR and TRMM-PR).

    PubMed

    Fall, Veronica M; Cao, Qing; Hong, Yang

    2013-01-01

    Spaceborne radars provide great opportunities to investigate the vertical structure of clouds and precipitation. Two typical spaceborne radars for such a study are the W-band Cloud Profiling Radar (CPR) and Ku-band Precipitation Radar (PR), which are onboard NASA's CloudSat and TRMM satellites, respectively. Compared to S-band ground-based radars, they have distinct scattering characteristics for different hydrometeors in clouds and precipitation. The combination of spaceborne and ground-based radar observations can help in the identification of hydrometeors and improve the radar-based quantitative precipitation estimation (QPE). This study analyzes the vertical structure of the 18 January, 2009 storm using data from the CloudSat CPR, TRMM PR, and a NEXRAD-based National Mosaic and Multisensor QPE (NMQ) system. Microphysics above, within, and below the melting layer are studied through an intercomparison of multifrequency measurements. Hydrometeors' type and their radar scattering characteristics are analyzed. Additionally, the study of the vertical profile of reflectivity (VPR) reveals the brightband properties in the cold-season precipitation and its effect on the radar-based QPE. In all, the joint analysis of spaceborne and ground-based radar data increases the understanding of the vertical structure of storm systems and provides a good insight into the microphysical modeling for weather forecasts.

  12. Intercomparison of Vertical Structure of Storms Revealed by Ground-Based (NMQ) and Spaceborne Radars (CloudSat-CPR and TRMM-PR)

    PubMed Central

    Fall, Veronica M.; Hong, Yang

    2013-01-01

    Spaceborne radars provide great opportunities to investigate the vertical structure of clouds and precipitation. Two typical spaceborne radars for such a study are the W-band Cloud Profiling Radar (CPR) and Ku-band Precipitation Radar (PR), which are onboard NASA's CloudSat and TRMM satellites, respectively. Compared to S-band ground-based radars, they have distinct scattering characteristics for different hydrometeors in clouds and precipitation. The combination of spaceborne and ground-based radar observations can help in the identification of hydrometeors and improve the radar-based quantitative precipitation estimation (QPE). This study analyzes the vertical structure of the 18 January, 2009 storm using data from the CloudSat CPR, TRMM PR, and a NEXRAD-based National Mosaic and Multisensor QPE (NMQ) system. Microphysics above, within, and below the melting layer are studied through an intercomparison of multifrequency measurements. Hydrometeors' type and their radar scattering characteristics are analyzed. Additionally, the study of the vertical profile of reflectivity (VPR) reveals the brightband properties in the cold-season precipitation and its effect on the radar-based QPE. In all, the joint analysis of spaceborne and ground-based radar data increases the understanding of the vertical structure of storm systems and provides a good insight into the microphysical modeling for weather forecasts. PMID:24459424

  13. Radar Image Simulation: Validation of the Point Scattering Method. Volume 2

    DTIC Science & Technology

    1977-09-01

    the Engineer Topographic Labor - atory (ETL), Fort Belvoir, Virginia. This Radar Simulation Study was performed to validate the point tcattering radar...e.n For radar, the number of Independent samples in a given re.-olution cell is given by 5 ,: N L 2w (16) L Acoso where: 0 Radar incidence angle; w

  14. Classification and Vertical Structure of Radar Precipitation Echoes at Naqu in Central Tibetan Plateau during the TIPEX-III Field Campaign

    NASA Astrophysics Data System (ADS)

    Luo, Y.; Wang, H.; Ma, R.; Zipser, E. J.; Liu, C.

    2017-12-01

    This study examines the vertical structure of precipitation echoes in central Tibetan Plateau using observations collected at Naqu during the Third Tibetan Plateau Atmospheric Scientific Experiment in July-August 2014. Precipitation reaching the surface is classified into stratiform, convective, and other by analyzing the vertical profiles of reflectivity (Ze) at 30-m spacing and 3-s temporal resolution made with the vertical pointing C-band frequency-modulated continuous-wave (C-FMCW) radar. Radar echoes with non-zero surface rainfall rate are observed during 17.96% of the entire observing period. About 52.03% of the precipitation reaching the surface includes a bright band and lacks a thick layer (≥1 km) of large Ze (> 35 dBZ); these are classified as stratiform; non-stratiform echoes with Ze > 35 dBZ are classified as convective (4.99%); the remainder (42.98%) as other. Based on concurrent measurements made with a collocated disdrometer, the classified stratiform, convective, and other precipitation echoes contribute 53.84%, 23.08%, and 23.08%, respectively, to the surface rainfall amount. Distinct internal structural features of each echo type are revealed by collectively analyzing the vertical profiles of Ze, radial velocity (Vr), and spectral width (SW) observed by the C-FMCW radar. The stratiform precipitation contains a melting-layer centered at 0.97 km above ground with an average depth of 415 m. The median Ze at 0°C -15°C levels in convective regions at Naqu is weaker than those in some midlatitude continental convection and stronger than those in some tropical continents, suggesting that convective intensity measured by mixed-phase microphysical processes at Naqu is intermediate.

  15. Using Ground Radar Interferometry for Precise Determining of Deformation and Vertical Deflection of Structures

    NASA Astrophysics Data System (ADS)

    Talich, Milan

    2017-12-01

    The paper describes possibilities of the relatively new technics - ground based radar interferometry for precise determining of deformation of structures. Special focus on the vertical deflection of bridge structures and on the horizontal movements of high-rise buildings and structural objects is presented. The technology of ground based radar interferometry can be used in practice to the contactless determination of deformations of structures with accuracy up to 0.01 mm in real time. It is also possible in real time to capture oscillations of the object with a frequency up to 50 Hz. Deformations can be determined simultaneously in multiple places of the object, for example a bridge structure at points distributed on the bridge deck at intervals of one or more meters. This allows to obtain both overall and detailed information about the properties of the structure during the dynamic load and monitoring the impact of movements either individual vehicles or groups. In the case of high-rise buildings, it is possible to monitor the horizontal vibration of the whole object at its different height levels. It is possible to detect and determine the compound oscillations that occur in some types of buildings. Then prevent any damage or even disasters in these objects. In addition to the necessary theory basic principles of using radar interferometry for determining of deformation of structures are given. Practical examples of determining deformation of bridge structures, water towers reservoirs, factory chimneys and wind power plants are also given. The IBIS-S interferometric radar of the Italian IDS manufacturer was used for the measurements.

  16. Vertical structure of the lower troposphere derived from MU radar, unmanned aerial vehicle, and balloon measurements during ShUREX 2015

    NASA Astrophysics Data System (ADS)

    Luce, Hubert; Kantha, Lakshmi; Hashiguchi, Hiroyuki; Lawrence, Dale; Mixa, Tyler; Yabuki, Masanori; Tsuda, Toshitaka

    2018-12-01

    The ShUREX (Shigaraki UAV Radar Experiment) 2015 campaign carried out at the Shigaraki Middle and Upper atmosphere (MU) observatory (Japan) in June 2015 provided a unique opportunity to compare vertical profiles of atmospheric parameters estimated from unmanned aerial vehicle (UAV), balloon, and radar data in the lower troposphere. The present work is intended primarily as a demonstration of the potential offered by combination of these three instruments for studying the small-scale structure and dynamics in the lower troposphere. Here, we focus on data collected almost simultaneously by two instrumented UAVs and two meteorological balloons, near the MU radar operated continuously during the campaign. The UAVs flew along helical ascending and descending paths at a nearly constant horizontal distance from the radar ( 1.0 km), while the balloons launched from the MU radar site drifted up to 3-5 km in the altitude range of comparisons ( 0.5 to 4.0 km) due to wind advection. Vertical profiles of squared Brünt-Väisälä frequency N 2 and squared vertical gradient of generalized potential refractive index M 2 were estimated at a vertical resolution of 20 m from pressure, temperature, and humidity data collected by UAVs and radiosondes. Profiles of M 2 were also estimated from MU radar echo power at vertical incidence at a vertical sampling of 20 m and various time resolutions (1-4 min). The balloons and the MU radar provided vertical profiles of wind and wind shear S so that two independent estimates of the gradient Richardson number ( Ri = N 2/ S 2) could be obtained at a range resolution of 150 m. The two estimates of Ri profiles also showed remarkable agreement at all altitudes. We show that all three instruments detected the same prominent temperature and humidity gradients, down to decameter scales in stratified conditions. These gradients extended horizontally over a few kilometers at least and persisted for hours without significant changes, indicating that the

  17. A Vertical Census of Precipitation Characteristics using Ground-based Dual-polarimetric Radar Data

    NASA Astrophysics Data System (ADS)

    Wolff, D. B.; Petersen, W. A.; Marks, D. A.; Pippitt, J. L.; Tokay, A.; Gatlin, P. N.

    2017-12-01

    Characterization of the vertical structure/variability of precipitation and resultant microphysics is critical in providing physical validation of space-based precipitation retrievals. In support of NASAs Global Precipitation Measurement (GPM) mission Ground Validation (GV) program, NASA has invested in a state-of-art dual-polarimetric radar known as NPOL. NPOL is routinely deployed on the Delmarva Peninsula in support of NASAs GPM Precipitation Research Facility (PRF). NPOL has also served as the backbone of several GPM field campaigns in Oklahoma, Iowa, South Carolina and most recently in the Olympic Mountains in Washington state. When precipitation is present, NPOL obtains very high-resolution vertical profiles of radar observations (e.g. reflectivity (ZH) and differential reflectivity (ZDR)), from which important particle size distribution parameters are retrieved such as the mass-weight mean diameter (Dm) and the intercept parameter (Nw). These data are then averaged horizontally to match the nadir resolution of the dual-frequency radar (DPR; 5 km) on board the GPM satellite. The GPM DPR, Combined, and radiometer algorithms (such as GPROF) rely on functional relationships built from assumed parametric relationships and/or retrieved parameter profiles and spatial distributions of particle size (PSD), water content, and hydrometeor phase within a given sample volume. Thus, the NPOL-retrieved profiles provide an excellent tool for characterization of the vertical profile structure and variability during GPM overpasses. In this study, we will use many such overpass comparisons to quantify an estimate of the true sub-IFOV variability as a function of hydrometeor and rain type (convective or stratiform). This presentation will discuss the development of a relational database to help provide a census of the vertical structure of precipitation via analysis and correlation of reflectivity, differential reflectivity, mean-weight drop diameter and the normalized

  18. Vertical structure of radar reflectivity in deep intense convective clouds over the tropics

    NASA Astrophysics Data System (ADS)

    Kumar, Shailendra; Bhat, G. S.

    2015-04-01

    This study is based on 10 years of radar reflectivity factor (Z) data derived from the TRMM Precipitation Radar (PR) measurements. We define two types of convective cells, namely, cumulonimbus towers (CbTs) and intense convective clouds (ICCs), essentially following the methodology used in deriving the vertical profiles of radar reflectivity (VPRR). CbT contains Z≥ 20 dBZ at 12 km height with its base height below 3 km. ICCs belong to the top 5% reflectivity population at 3 km and 8 km altitude. Regional differences in the vertical structure of convective cells have been explored for two periods, namely, JJAS (June, July, August and September) and JFM (January, February and March) months. Frequency of occurrences of CbTs and ICCs depend on the region. Africa and Latin America are the most productive regions for the CbTs while the foothills of Western Himalaya contain the most intense profiles. Among the oceanic areas, the Bay of Bengal has the strongest vertical profile, whereas Atlantic Ocean has the weakest profile during JJAS. During JFM months, maritime continent has the strongest vertical profile whereas western equatorial Indian Ocean has the weakest. Monsoon clouds lie between the continental and oceanic cases. The maximum heights of 30 and 40 dBZ reflectivities (denoted by MH30 and MH40, respectively) are also studied. MH40 shows a single mode and peaks around 5.5 km during both JJAS and JFM months. MH30 shows two modes, around 5 km and between 8 km and 10 km, respectively. It is also shown that certain conclusions such as the area/region with the most intense convective cells, depend of the reference height used in defining a convective cell.

  19. Improvement of vertical profiles of raindrop size distribution from micro rain radar using 2D video disdrometer measurements

    NASA Astrophysics Data System (ADS)

    Adirosi, E.; Baldini, L.; Roberto, N.; Gatlin, P.; Tokay, A.

    2016-03-01

    A measurement scheme aimed at investigating precipitation properties based on collocated disdrometer and profiling instruments is used in many experimental campaigns. Raindrop size distribution (RSD) estimated by disdrometer is referred to the ground level; the collocated profiling instrument is supposed to provide complementary estimation at different heights of the precipitation column above the instruments. As part of the Special Observation Period 1 of the HyMeX (Hydrological Cycle in the Mediterranean Experiment) project, conducted between 5 September and 6 November 2012, a K-band vertically pointing micro rain radar (MRR) and a 2D video disdrometer (2DVD) were installed close to each other at a site in the historic center of Rome (Italy). The raindrop size distributions collected by 2D video disdrometer are considered to be fairly accurate within the typical sizes of drops. Vertical profiles of raindrop sizes up to 1085 m are estimated from the Doppler spectra measured by the micro rain radar with a height resolution of 35 m. Several issues related to vertical winds, attenuation correction, Doppler spectra aliasing, and range-Doppler ambiguity limit the performance of MRR in heavy precipitation or in convection, conditions that frequently occur in late summer or in autumn in Mediterranean regions. In this paper, MRR Doppler spectra are reprocessed, exploiting the 2DVD measurements at ground to estimate the effects of vertical winds at 105 m (the most reliable MRR lower height), in order to provide a better estimation of vertical profiles of raindrop size distribution from MRR spectra. Results show that the reprocessing procedure leads to a better agreement between the reflectivity computed at 105 m from the reprocessed MRR spectra and that obtained from the 2DVD data. Finally, vertical profiles of MRR-estimated RSDs and their relevant moments (namely median volume diameter and reflectivity) are presented and discussed in order to investigate the

  20. Estimation of mesospheric vertical winds from a VHF meteor radar at King Sejong Station, Antarctica (62.2S, 58.8W)

    NASA Astrophysics Data System (ADS)

    Kim, Y.; Lee, C.; Kim, J.; Jee, G.

    2013-12-01

    For the first time, vertical winds near the mesopause region were estimated from radial velocities of meteor echoes detected by a VHF meteor radar at King Sejong Station (KSS) in 2011 and 2012. Since the radar usually detects more than a hundred echoes every hour in an altitude bin of 88 - 92 km, much larger than other radars, we were able to fit measured radial velocities of these echoes with a 6 component model that consists of horizontal winds, spatial gradients of horizontal winds and vertical wind. The conventional method of deriving horizontal winds from meteor echoes utilizes a 2 component model, assuming that vertical winds and spatial gradients of horizontal winds are negligible. We analyzed the radar data obtained for 8400 hours in 2012 and 8100 hours in 2011. We found that daily mean values of vertical winds are mostly within +/- 1 m/s, whereas those of zonal winds are a few tens m/s mostly eastward. The daily mean vertical winds sometimes stay positive or negative for more than 20 days, implying that the atmosphere near the mesopause experiences episodically a large scale low and high pressure environments, respectively, like the tropospheric weather system. By conducting Lomb-normalized periodogram analysis, we also found that the vertical winds have diurnal, semidiurnal and terdiurnal tidal components with about equal significance, in contrast to horizontal winds that show a dominant semidiurnal one. We will discuss about uncertainties of the estimated vertical wind and possible reasons of its tidal and daily variations.

  1. Radar Image Simulation: Validation of the Point Scattering Model. Volume 1

    DTIC Science & Technology

    1977-09-01

    I reports the work and results with technical deLails deferred to the appendices. Voluime II Is a collection of appendices containing the individual...separation between successive points on the ground. ’Look-dir- action " Is a very Important concept to imaging radars. It means, given, a particular...point, we have watched as the radar transmitted a pulse of enerqy to the ground. We observed the Inter- action of this pulse with the ground. We followed

  2. Vertical Rise Velocity of Equatorial Plasma Bubbles Estimated from Equatorial Atmosphere Radar Observations and High-Resolution Bubble Model Simulations

    NASA Astrophysics Data System (ADS)

    Yokoyama, T.; Ajith, K. K.; Yamamoto, M.; Niranjan, K.

    2017-12-01

    Equatorial plasma bubble (EPB) is a well-known phenomenon in the equatorial ionospheric F region. As it causes severe scintillation in the amplitude and phase of radio signals, it is important to understand and forecast the occurrence of EPBs from a space weather point of view. The development of EPBs is presently believed as an evolution of the generalized Rayleigh-Taylor instability. We have already developed a 3D high-resolution bubble (HIRB) model with a grid spacing of as small as 1 km and presented nonlinear growth of EPBs which shows very turbulent internal structures such as bifurcation and pinching. As EPBs have field-aligned structures, the latitude range that is affected by EPBs depends on the apex altitude of EPBs over the dip equator. However, it was not easy to observe the apex altitude and vertical rise velocity of EPBs. Equatorial Atmosphere Radar (EAR) in Indonesia is capable of steering radar beams quickly so that the growth phase of EPBs can be captured clearly. The vertical rise velocities of the EPBs observed around the midnight hours are significantly smaller compared to those observed in postsunset hours. Further, the vertical growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The HIRB model with varying background conditions are employed to investigate the possible factors that control the vertical rise velocity and maximum attainable altitudes of EPBs. The estimated rise velocities from EAR observations at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model.

  3. Observations of the structure and vertical transport of the polar upper ionosphere with the EISCAT VHF radar. II - First investigations of the topside O(+) and H(+) vertical ion flows

    NASA Technical Reports Server (NTRS)

    Wu, Jian; Blanc, Michel; Alcayde, Denis; Barakat, Abdullah R.; Fontanari, Jean; Blelly, Pierre-Louis; Kofman, Wlodek

    1992-01-01

    EISCAT VHF radar was used to investigate the vertical flows of H(+) and O(+) ions in the topside high-latitude ionosphere. The radar transmitted a single long pulse to probe the ionosphere from 300 to 1200 km altitude. A calculation scheme is developed to deduce the H(+) drift velocity from the coupled momentum equations of H(+), O(+), and the electrons, using the radar data and a neutral atmosphere model. The H(+) vertical drift velocity was expressed as a linear combination of the different forces acting on the plasma. Two nights, one very quiet, one with moderate magnetic activity, were used to test the technique and to provide a first study of the morphology and orders of magnitudes of ion outflow fluxes over Tromso. O(+) vertical flows were found to be downward or close to zero most of the time in the topside ionosphere; they appeared to be strongly correlated with magnetic activity during the disturbed night. H(+) topside ion fluxes were always directed upward, with velocity reaching 500-1000 m/s. A permanent outflow of H(+) ions is inferred.

  4. Airborne Radar Observations of Severe Hailstorms: Implications for Future Spaceborne Radar

    NASA Technical Reports Server (NTRS)

    Heymsfield, Gerald M.; Tian, Lin; Li, Lihua; McLinden, Matthew; Cervantes, Jaime I.

    2013-01-01

    A new dual-frequency (Ku and Ka band) nadir-pointing Doppler radar on the high-altitude NASA ER-2 aircraft, called the High-Altitude Imaging Wind and Rain Airborne Profiler (HIWRAP), has collected data over severe thunderstorms in Oklahoma and Kansas during the Midlatitude Continental Convective Clouds Experiment (MC3E). The overarching motivation for this study is to understand the behavior of the dualwavelength airborne radar measurements in a global variety of thunderstorms and how these may relate to future spaceborne-radar measurements. HIWRAP is operated at frequencies that are similar to those of the precipitation radar on the Tropical Rainfall Measuring Mission (Ku band) and the upcoming Global Precipitation Measurement mission satellite's dual-frequency (Ku and Ka bands) precipitation radar. The aircraft measurements of strong hailstorms have been combined with ground-based polarimetric measurements to obtain a better understanding of the response of the Ku- and Ka-band radar to the vertical distribution of the hydrometeors, including hail. Data from two flight lines on 24 May 2011 are presented. Doppler velocities were approx. 39m/s2at 10.7-km altitude from the first flight line early on 24 May, and the lower value of approx. 25m/s on a second flight line later in the day. Vertical motions estimated using a fall speed estimate for large graupel and hail suggested that the first storm had an updraft that possibly exceeded 60m/s for the more intense part of the storm. This large updraft speed along with reports of 5-cm hail at the surface, reflectivities reaching 70 dBZ at S band in the storm cores, and hail signals from polarimetric data provide a highly challenging situation for spaceborne-radar measurements in intense convective systems. The Ku- and Ka-band reflectivities rarely exceed approx. 47 and approx. 37 dBZ, respectively, in these storms.

  5. Antenna induced range smearing in MST radars

    NASA Technical Reports Server (NTRS)

    Watkins, B. J.; Johnston, P. E.

    1984-01-01

    There is considerable interest in developing stratosphere troposphere (ST) and mesosphere stratosphere troposphere (MST) radars for higher resolution to study small-scale turbulent structures and waves. At present most ST and MST radars have resolutions of 150 meters or larger, and are not able to distinguish the thin (40 - 100 m) turbulent layers that are known to occur in the troposphere and stratosphere, and possibly in the mesosphere. However the antenna beam width and sidelobe level become important considerations for radars with superior height resolution. The objective of this paper is to point out that for radars with range resolutions of about 150 meters or less, there may be significant range smearing of the signals from mesospheric altitudes due to the finite beam width of the radar antenna. At both stratospheric and mesospheric heights the antenna sidelobe level for lear equally spaced phased arrays may also produce range aliased signals. To illustrate this effect the range smearing functions for two vertically directed antennas have been calculated, (1) an array of 32 coaxial-collinear strings each with 48 elements that simulates the vertical beam of the Poker Flat, Glaska, MST radar; and (2) a similar, but smaller, array of 16 coaxial-collinear strings each with 24 elements.

  6. Radar - ARL Wind Profilerwith RASS, Boardman - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  7. Radar antenna pointing for optimized signal to noise ratio.

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

    Doerry, Armin Walter; Marquette, Brandeis

    2013-01-01

    The Signal-to-Noise Ratio (SNR) of a radar echo signal will vary across a range swath, due to spherical wavefront spreading, atmospheric attenuation, and antenna beam illumination. The antenna beam illumination will depend on antenna pointing. Calculations of geometry are complicated by the curved earth, and atmospheric refraction. This report investigates optimizing antenna pointing to maximize the minimum SNR across the range swath.

  8. Kinematic and Hydrometer Data Products from Scanning Radars during MC3E

    DOE Data Explorer

    matthews, Alyssa; Dolan, Brenda; Rutledge, Steven

    2016-02-29

    Recently the Radar Meteorology Group at Colorado State University has completed major case studies of some top cases from MC3E including 25 April, 20 May and 23 May 2011. A discussion on the analysis methods as well as radar quality control methods is included. For each case, a brief overview is first provided. Then, multiple Doppler (using available X-SAPR and C-SAPR data) analyses are presented including statistics on vertical air motions, sub-divided by convective and stratiform precipitation. Mean profiles and CFAD's of vertical motion are included to facilitate comparison with ASR model simulations. Retrieved vertical motion has also been verified with vertically pointing profiler data. Finally for each case, hydrometeor types are included derived from polarimetric radar observations. The latter can be used to provide comparisons to model-generated hydrometeor fields. Instructions for accessing all the data fields are also included. The web page can be found at: http://radarmet.atmos.colostate.edu/mc3e/research/

  9. Radar - ANL Wind Profiler with RASS, Yakima - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  10. Radar - ESRL Wind Profiler with RASS, Condon - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  11. Radar - ESRL Wind Profiler with RASS, Prineville - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  12. Radar - ESRL Wind Profiler with RASS, Troutdale - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  13. Radar - ANL Wind Profiler with RASS, Goldendale - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  14. Radar Image of Dublin, Ireland

    NASA Image and Video Library

    2017-12-08

    Visualization Date 1994-04-11 This radar image of Dublin, Ireland, shows how the radar distingishes between densely populated urban areas and nearby areas that are relatively unsettled. In the center of the image is the city's natural harbor along the Irish Sea. The pinkish areas in the center are the densely populated parts of the city and the blue/green areas are the suburbs. The two ends of the Dublin Bay are Howth Point, the circular peninsula near the upper right side of the image, and Dun Laoghaire, the point to the south. The small island just north of Howth is called "Ireland's Eye," and the larger island, near the upper right corner of the image is Lambay Island. The yellow/green mountains in the lower left of the image (south) are the Wicklow Mountains. The large lake in the lower left, nestled within these mountains, is the Poulaphouca Reservoir along River Liffey. The River Liffey, the River Dodder and the Tolka River are the three rivers that flow into Dublin. The straight features west of the city are the Grand Canal and the three rivers are the faint lines above and below these structures. The dark X-shaped feature just to the north of the city is the Dublin International Airport. The image was acquired by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture (SIR-C/X-SAR) when it flew aboard the space shuttle Endeavour on April 11, 1994. This area is centered at 53.3 degrees north latitude, 6.2 degrees west longitude. The area shown is approximately 55 kilometers by 42 kilometers (34 miles by 26 miles). The colors are assigned to different frequencies and polarizations of the radar as follows: Red is L-band horizontally transmitted, horizontally received; green is L-band vertically transmitted, vertically received; and blue is C-band vertically transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian, and the United States space agencies, is part of NASA's Mission to Planet Earth. Credit: NASA/GSFC For more

  15. Ultrahigh vertical resolution radar measurements in the lower stratosphere at Arecibo

    NASA Technical Reports Server (NTRS)

    Ierkic, H. M.; Perillat, P.; Woodman, R. F.

    1990-01-01

    The paper reports on heretofore unprecedented observations of the turbulent layers in the lower stratosphere using the Arecibo 2380-MHz radar. Spectral profiles with about 20 m height and 15 s time resolutions at altitudes in the range 16-19 km are used to parametrize relevant characteristics of the turbulence, namely, vertical widths, distributions, lifetimes, and cutoffs height. These measurements validate previous deconvolved estimates and are free from contaminating factors like shear or beam broadening and partial reflections. Some theoretical predictions are verified, in particular those relating to the height of cutoff and the outer scale of the turbulence.

  16. Vertical air motion retrievals in deep convective clouds using the ARM scanning radar network in Oklahoma during MC3E

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

    North, Kirk W.; Oue, Mariko; Kollias, Pavlos

    The US Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) program's Southern Great Plains (SGP) site includes a heterogeneous distributed scanning Doppler radar network suitable for collecting coordinated Doppler velocity measurements in deep convective clouds. The surrounding National Weather Service (NWS) Next Generation Weather Surveillance Radar 1988 Doppler (NEXRAD WSR-88D) further supplements this network. Radar velocity measurements are assimilated in a three-dimensional variational (3DVAR) algorithm that retrieves horizontal and vertical air motions over a large analysis domain (100 km × 100 km) at storm-scale resolutions (250 m). For the first time, direct evaluation of retrieved vertical air velocities with thosemore » from collocated 915 MHz radar wind profilers is performed. Mean absolute and root-mean-square differences between the two sources are of the order of 1 and 2 m s -1, respectively, and time–height correlations are of the order of 0.5. An empirical sensitivity analysis is done to determine a range of 3DVAR constraint weights that adequately satisfy the velocity observations and anelastic mass continuity. It is shown that the vertical velocity spread over this range is of the order of 1 m s -1. The 3DVAR retrievals are also compared to those obtained from an iterative upwards integration technique. Lastly, the results suggest that the 3DVAR technique provides a robust, stable solution for cases in which integration techniques have difficulty satisfying velocity observations and mass continuity simultaneously.« less

  17. Vertical air motion retrievals in deep convective clouds using the ARM scanning radar network in Oklahoma during MC3E

    DOE PAGES

    North, Kirk W.; Oue, Mariko; Kollias, Pavlos; ...

    2017-08-04

    The US Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) program's Southern Great Plains (SGP) site includes a heterogeneous distributed scanning Doppler radar network suitable for collecting coordinated Doppler velocity measurements in deep convective clouds. The surrounding National Weather Service (NWS) Next Generation Weather Surveillance Radar 1988 Doppler (NEXRAD WSR-88D) further supplements this network. Radar velocity measurements are assimilated in a three-dimensional variational (3DVAR) algorithm that retrieves horizontal and vertical air motions over a large analysis domain (100 km × 100 km) at storm-scale resolutions (250 m). For the first time, direct evaluation of retrieved vertical air velocities with thosemore » from collocated 915 MHz radar wind profilers is performed. Mean absolute and root-mean-square differences between the two sources are of the order of 1 and 2 m s -1, respectively, and time–height correlations are of the order of 0.5. An empirical sensitivity analysis is done to determine a range of 3DVAR constraint weights that adequately satisfy the velocity observations and anelastic mass continuity. It is shown that the vertical velocity spread over this range is of the order of 1 m s -1. The 3DVAR retrievals are also compared to those obtained from an iterative upwards integration technique. Lastly, the results suggest that the 3DVAR technique provides a robust, stable solution for cases in which integration techniques have difficulty satisfying velocity observations and mass continuity simultaneously.« less

  18. Radar - ANL Wind Profiler with RASS, Walla Walla - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  19. Radar - ESRL Wind Profiler with RASS, Wasco Airport - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2017-10-23

    **Winds** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature** To measure atmospheric temperature, a radio acoustic sound system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60-m up to 3.5k m for the 449 MHz.

  20. Radar QPE for hydrological design: Intensity-Duration-Frequency curves

    NASA Astrophysics Data System (ADS)

    Marra, Francesco; Morin, Efrat

    2015-04-01

    Intensity-duration-frequency (IDF) curves are widely used in flood risk management since they provide an easy link between the characteristics of a rainfall event and the probability of its occurrence. They are estimated analyzing the extreme values of rainfall records, usually basing on raingauge data. This point-based approach raises two issues: first, hydrological design applications generally need IDF information for the entire catchment rather than a point, second, the representativeness of point measurements decreases with the distance from measure location, especially in regions characterized by steep climatological gradients. Weather radar, providing high resolution distributed rainfall estimates over wide areas, has the potential to overcome these issues. Two objections usually restrain this approach: (i) the short length of data records and (ii) the reliability of quantitative precipitation estimation (QPE) of the extremes. This work explores the potential use of weather radar estimates for the identification of IDF curves by means of a long length radar archive and a combined physical- and quantitative- adjustment of radar estimates. Shacham weather radar, located in the eastern Mediterranean area (Tel Aviv, Israel), archives data since 1990 providing rainfall estimates for 23 years over a region characterized by strong climatological gradients. Radar QPE is obtained correcting the effects of pointing errors, ground echoes, beam blockage, attenuation and vertical variations of reflectivity. Quantitative accuracy is then ensured with a range-dependent bias adjustment technique and reliability of radar QPE is assessed by comparison with gauge measurements. IDF curves are derived from the radar data using the annual extremes method and compared with gauge-based curves. Results from 14 study cases will be presented focusing on the effects of record length and QPE accuracy, exploring the potential application of radar IDF curves for ungauged locations and

  1. Vertical radar profiles for the calibration of unsaturated flow models under dynamic water table conditions

    NASA Astrophysics Data System (ADS)

    Cassiani, G.; Gallotti, L.; Ventura, V.; Andreotti, G.

    2003-04-01

    The identification of flow and transport characteristics in the vadose zone is a fundamental step towards understanding the dynamics of contaminated sites and the resulting risk of groundwater pollution. Borehole radar has gained popularity for the monitoring of moisture content changes, thanks to its apparent simplicity and its high resolution characteristics. However, cross-hole radar requires closely spaced (a few meters), plastic-cased boreholes, that are rarely available as a standard feature in sites of practical interest. Unlike cross-hole applications, Vertical Radar Profiles (VRP) require only one borehole, with practical and financial benefits. High-resolution, time-lapse VRPs have been acquired at a crude oil contaminated site in Trecate, Northern Italy, on a few existing boreholes originally developed for remediation via bioventing. The dynamic water table conditions, with yearly oscillations of roughly 5 m from 6 to 11 m bgl, offers a good opportunity to observe via VRP a field scale drainage-imbibition process. Arrival time inversion has been carried out using a regularized tomographic algorithm, in order to overcome the noise introduced by first arrival picking. Interpretation of the vertical profiles in terms of moisture content has been based on standard models (Topp et al., 1980; Roth et al., 1990). The sedimentary sequence manifests itself as a cyclic pattern in moisture content over most of the profiles. We performed preliminary Richards' equation simulations with time varying later table boundary conditions, in order to estimate the unsaturated flow parameters, and the results have been compared with laboratory evidence from cores.

  2. Space Radar Image of Saline Valley, California

    NASA Technical Reports Server (NTRS)

    1999-01-01

    This is a three-dimensional perspective view of Saline Valley, about 30 km (19 miles) east of the town of Independence, California created by combining two spaceborne radar images using a technique known as interferometry. Visualizations like this one are helpful to scientists because they clarify the relationships of the different types of surfaces detected by the radar and the shapes of the topographic features such as mountains and valleys. The view is looking southwest across Saline Valley. The high peaks in the background are the Inyo Mountains, which rise more than 3,000 meters (10,000 feet) above the valley floor. The dark blue patch near the center of the image is an area of sand dunes. The brighter patches to the left of the dunes are the dry, salty lake beds of Saline Valley. The brown and orange areas are deposits of boulders, gravel and sand known as alluvial fans. The image was constructed by overlaying a color composite radar image on top of a digital elevation map. The radar image was taken by the Spaceborne Imaging Radar-C/X-bandSynthetic Aperture Radar (SIR-C/X-SAR) on board the space shuttleEndeavour in October 1994. The digital elevation map was producedusing radar interferometry, a process in which radar data are acquired on different passes of the space shuttle. The two data passes are compared to obtain elevation information. The elevation data were derived from a 1,500-km-long (930-mile) digital topographic map processed at JPL. Radar image data are draped over the topography to provide the color with the following assignments: red is L-band vertically transmitted, vertically received; green is C-band vertically transmitted, vetically received; and blue is the ratio of C-band vertically transmitted, vertically received to L-band vertically transmitted, vertically received. This image is centered near 36.8 degrees north latitude and 117.7 degrees west longitude. No vertical exaggeration factor has been applied to the data. SIR-C/X-SAR, a joint

  3. Space Radar Image of Karakax Valley, China 3-D

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This three-dimensional perspective of the remote Karakax Valley in the northern Tibetan Plateau of western China was created by combining two spaceborne radar images using a technique known as interferometry. Visualizations like this are helpful to scientists because they reveal where the slopes of the valley are cut by erosion, as well as the accumulations of gravel deposits at the base of the mountains. These gravel deposits, called alluvial fans, are a common landform in desert regions that scientists are mapping in order to learn more about Earth's past climate changes. Higher up the valley side is a clear break in the slope, running straight, just below the ridge line. This is the trace of the Altyn Tagh fault, which is much longer than California's San Andreas fault. Geophysicists are studying this fault for clues it may be able to give them about large faults. Elevations range from 4000 m (13,100 ft) in the valley to over 6000 m (19,700 ft) at the peaks of the glaciated Kun Lun mountains running from the front right towards the back. Scale varies in this perspective view, but the area is about 20 km (12 miles) wide in the middle of the image, and there is no vertical exaggeration. The two radar images were acquired on separate days during the second flight of the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) aboard the space shuttle Endeavour in October 1994. The interferometry technique provides elevation measurements of all points in the scene. The resulting digital topographic map was used to create this view, looking northwest from high over the valley. Variations in the colors can be related to gravel, sand and rock outcrops. This image is centered at 36.1 degrees north latitude, 79.2 degrees east longitude. Radar image data are draped over the topography to provide the color with the following assignments: Red is L-band vertically transmitted, vertically received; green is the average of L-band vertically transmitted

  4. Radar - 449MHz - Forks, WA (FKS) - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  5. Radar - 449MHz - Forks, WA (FKS) - Reviewed Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  6. Radar - 449MHz - Astoria, OR (AST) - Reviewed Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  7. Radar - 449MHz - Astoria, OR (AST) - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  8. The vertical profile of radar reflectivity of convective cells: A strong indicator of storm intensity and lightning probability?

    NASA Technical Reports Server (NTRS)

    Zipser, Edward J.; Lutz, Kurt R.

    1994-01-01

    Reflectivity data from Doppler radars are used to construct vertical profiles of radar reflectivity (VPRR) of convective cells in mesoscale convective systems (MCSs) in three different environmental regimes. The National Center for Atmospheric Research CP-3 and CP-4 radars are used to calculate median VPRR for MCSs in the Oklahoma-Kansas Preliminary Regional Experiment for STORM-Central in 1985. The National Oceanic and Atmospheric Administration-Tropical Ocean Global Atmosphere radar in Darwin, Australia, is used to calculate VPRR for MCSs observed both in oceanic, monsoon regimes and in continental, break period regimes during the wet seasons of 1987/88 and 1988/89. The midlatitude and tropical continental VPRRs both exhibit maximum reflectivity somewhat above the surface and have a gradual decrease in reflectivity with height above the freezing level. In sharp contrast, the tropical oceanic profile has a maximum reflectivity at the lowest level and a very rapid decrease in reflectivity with height beginning just above the freezing level. The tropical oceanic profile in the Darwin area is almost the same shape as that for two other tropical oceanic regimes, leading to the conclustion that it is characteristic. The absolute values of reflectivity in the 0 to 20 C range are compared with values in the literature thought to represent a threshold for rapid storm electrification leading to lightning, about 40 dBZ at -10 C. The large negative vertical gradient of reflectivity in this temperature range for oceanic storms is hypothesized to be a direct result of the characteristically weaker vertical velocities observed in MCSs over tropical oceans. It is proposed, as a necessary condition for rapid electrification, that a convective cell must have its updraft speed exceed some threshold value. Based upon field program data, a tentative estimate for the magnitude of this threshold is 6-7 m/s for mean speed and 10-12 m/s for peak speed.

  9. Space Radar Image of Owens Valley, California

    NASA Technical Reports Server (NTRS)

    1999-01-01

    This is a three-dimensional perspective view of Owens Valley, near the town of Bishop, California that was created by combining two spaceborne radar images using a technique known as interferometry. Visualizations like this one are helpful to scientists because they clarify the relationships of the different types of surfaces detected by the radar and the shapes of the topographic features such as mountains and valleys. The view is looking southeast along the eastern edge of Owens Valley. The White Mountains are in the center of the image, and the Inyo Mountains loom in the background. The high peaks of the White Mountains rise more than 3,000 meters (10,000 feet) above the valley floor. The runways of the Bishop airport are visible at the right edge of the image. The meandering course of the Owens River and its tributaries appear light blue on the valley floor. Blue areas in the image are smooth, yellow areas are rock outcrops, and brown areas near the mountains are deposits of boulders, gravel and sand known as alluvial fans. The image was constructed by overlaying a color composite radar image on top of a digital elevation map. The radar data were taken by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) on board the space shuttle Endeavour in October 1994. The digital elevation map was produced using radar interferometry, a process in which radar data are acquired on different passes of the space shuttle. The two data passes are compared to obtain elevation information. The elevation data were derived from a 1,500-km-long (930-mile) digital topographic map processed at JPL. Radar image data are draped over the topography to provide the color with the following assignments: red is L-band vertically transmitted, vertically received; green is C-band vertically transmitted, vertically received; and blue is the ratio of C-band vertically transmitted, vertically received to L-band vertically transmitted, vertically received. This image is

  10. 51. View of upper radar scanner switch in radar scanner ...

    Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey

    51. View of upper radar scanner switch in radar scanner building 105 from upper catwalk level showing emanating waveguides from upper switch (upper one-fourth of photograph) and emanating waveguides from lower radar scanner switch in vertical runs. - Clear Air Force Station, Ballistic Missile Early Warning System Site II, One mile west of mile marker 293.5 on Parks Highway, 5 miles southwest of Anderson, Anderson, Denali Borough, AK

  11. The proposed flatland radar

    NASA Technical Reports Server (NTRS)

    Green, J. L.; Gage, K. S.; Vanzandt, T. E.; Nastrom, G. D.

    1986-01-01

    A flexible very high frequency (VHF) stratosphere-troposphere (ST) radar configured for meteorological research is to be constructed near Urbana, Illinois. Measurement of small vertical velocities associated with synoptic-scale meteorology can be performed. A large Doppler microwave radar (CHILL) is located a few km from the site of the proposed ST radar. Since the microwave radar can measure the location and velocity of hydrometeors and the VHF ST radar can measure clear (or cloudy) air velocities, simultaneous observations by these two radars of stratiform or convective weather systems would provide valuable meteorological information.

  12. Radar - ESRL Wind Profiler with RASS, Wasco Airport - Derived Data

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

    McCaffrey, Katherine

    Profiles of turbulence dissipation rate for 15-minute intervals, time-stamped at the beginning of the 15-minute period, during the final 30 minutes of each hour. During that time, the 915-MHz wind profiling radar was in an optimized configuration with a vertically pointing beam only for measuring accurate spectral widths of vertical velocity. A bias-corrected dissipation rate also was profiled (described in McCaffrey et al. 2017). Hourly files contain two 15-minute profiles.

  13. Radar - 449MHz - North Bend, OR (OTH) - Raw Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  14. Radar - 449MHz - North Bend, OR (OTH) - Reviewed Data

    DOE Data Explorer

    Gottas, Daniel

    2018-06-25

    **Winds.** A radar wind profiler measures the Doppler shift of electromagnetic energy scattered back from atmospheric turbulence and hydrometeors along 3-5 vertical and off-vertical point beam directions. Back-scattered signal strength and radial-component velocities are remotely sensed along all beam directions and are combined to derive the horizontal wind field over the radar. These data typically are sampled and averaged hourly and usually have 6-m and/or 100-m vertical resolutions up to 4 km for the 915 MHz and 8 km for the 449 MHz systems. **Temperature.** To measure atmospheric temperature, a radio acoustic sounding system (RASS) is used in conjunction with the wind profile. These data typically are sampled and averaged for five minutes each hour and have a 60-m vertical resolution up to 1.5 km for the 915 MHz and 60 m up to 3.5 km for the 449 MHz. **Moments and Spectra.** The raw spectra and moments data are available for all dwells along each beam and are stored in daily files. For each day, there are files labeled "header" and "data." These files are generated by the radar data acquisition system (LAP-XM) and are encoded in a proprietary binary format. Values of spectral density at each Doppler velocity (FFT point), as well as the radial velocity, signal-to-noise ratio, and spectra width for the selected signal peak are included in these files. Attached zip files, *449mhz-spectra-data-extraction.zip* and *449mhz-moment-data-extraction.zip*, include executables to unpack the spectra, (GetSpectra32.exe) and moments (GetMomSp32.exe), respectively. Documentation on usage and output file formats also are included in the zip files.

  15. Space Radar Image of Oil Slicks

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This is a radar image of an offshore drilling field about 150 km (93 miles) west of Bombay, India, in the Arabian Sea. The dark streaks are extensive oil slicks surrounding many of the drilling platforms, which appear as bright white spots. Radar images are useful for detecting and measuring the extent of oil seepages on the ocean surface, from both natural and industrial sources. The long, thin streaks extending from many of the platforms are spreading across the sea surface, pushed by local winds. The larger dark patches are dispersed slicks that were likely discharged earlier than the longer streaks, when the winds were probably from a different direction. The dispersed oil will eventually spread out over the more dense water and become a layer which is a single molecule thick. Many forms of oil, both from biological and from petroleum sources, smooth out the ocean surface, causing the area to appear dark in radar images. There are also two forms of ocean waves shown in this image. The dominant group of large waves (upper center) are called internal waves. These waves are formed below the ocean surface at the boundary between layers of warm and cold water and they appear in the radar image because of the way they change the ocean surface. Ocean swells, which are waves generated by winds, are shown throughout the image but are most distinct in the blue area adjacent to the internal waves. Identification of waves provide oceanographers with information about the smaller scale dynamic processes of the ocean. This image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) aboard the space shuttle Endeavour on October 9, 1994. The colors are assigned to different frequencies and polarizations of the radar as follows: Red is L-band vertically transmitted, vertically received; green is the average of L-band vertically transmitted, vertically received and C-band vertically transmitted, vertically received; blue is C

  16. Ice Cloud Optical Thickness and Extinction Estimates from Radar Measurements.

    NASA Astrophysics Data System (ADS)

    Matrosov, Sergey Y.; Shupe, Matthew D.; Heymsfield, Andrew J.; Zuidema, Paquita

    2003-11-01

    A remote sensing method is proposed to derive vertical profiles of the visible extinction coefficients in ice clouds from measurements of the radar reflectivity and Doppler velocity taken by a vertically pointing 35-GHz cloud radar. The extinction coefficient and its vertical integral, optical thickness τ, are among the fundamental cloud optical parameters that, to a large extent, determine the radiative impact of clouds. The results obtained with this method could be used as input for different climate and radiation models and for comparisons with parameterizations that relate cloud microphysical parameters and optical properties. An important advantage of the proposed method is its potential applicability to multicloud situations and mixed-phase conditions. In the latter case, it might be able to provide the information on the ice component of mixed-phase clouds if the radar moments are dominated by this component. The uncertainties of radar-based retrievals of cloud visible optical thickness are estimated by comparing retrieval results with optical thicknesses obtained independently from radiometric measurements during the yearlong Surface Heat Budget of the Arctic Ocean (SHEBA) field experiment. The radiometric measurements provide a robust way to estimate τ but are applicable only to optically thin ice clouds without intervening liquid layers. The comparisons of cloud optical thicknesses retrieved from radar and from radiometer measurements indicate an uncertainty of about 77% and a bias of about -14% in the radar estimates of τ relative to radiometric retrievals. One possible explanation of the negative bias is an inherently low sensitivity of radar measurements to smaller cloud particles that still contribute noticeably to the cloud extinction. This estimate of the uncertainty is in line with simple theoretical considerations, and the associated retrieval accuracy should be considered good for a nonoptical instrument, such as radar. This paper also

  17. Evaluation of meteorological airborne Doppler radar

    NASA Technical Reports Server (NTRS)

    Hildebrand, P. H.; Mueller, C. K.

    1984-01-01

    This paper will discuss the capabilities of airborne Doppler radar for atmospheric sciences research. The evaluation is based on airborne and ground based Doppler radar observations of convective storms. The capability of airborne Doppler radar to measure horizontal and vertical air motions is evaluated. Airborne Doppler radar is shown to be a viable tool for atmospheric sciences research.

  18. Combination of radar and daily precipitation data to estimate meaningful sub-daily point precipitation extremes

    NASA Astrophysics Data System (ADS)

    Bárdossy, András; Pegram, Geoffrey

    2017-01-01

    The use of radar measurements for the space time estimation of precipitation has for many decades been a central topic in hydro-meteorology. In this paper we are interested specifically in daily and sub-daily extreme values of precipitation at gauged or ungauged locations which are important for design. The purpose of the paper is to develop a methodology to combine daily precipitation observations and radar measurements to estimate sub-daily extremes at point locations. Radar data corrected using precipitation-reflectivity relationships lead to biased estimations of extremes. Different possibilities of correcting systematic errors using the daily observations are investigated. Observed gauged daily amounts are interpolated to unsampled points and subsequently disaggregated using the sub-daily values obtained by the radar. Different corrections based on the spatial variability and the subdaily entropy of scaled rainfall distributions are used to provide unbiased corrections of short duration extremes. Additionally a statistical procedure not based on a matching day by day correction is tested. In this last procedure as we are only interested in rare extremes, low to medium values of rainfall depth were neglected leaving a small number of L days of ranked daily maxima in each set per year, whose sum typically comprises about 50% of each annual rainfall total. The sum of these L day maxima is first iterpolated using a Kriging procedure. Subsequently this sum is disaggregated to daily values using a nearest neighbour procedure. The daily sums are then disaggregated by using the relative values of the biggest L radar based days. Of course, the timings of radar and gauge maxima can be different, so the method presented here uses radar for disaggregating daily gauge totals down to 15 min intervals in order to extract the maxima of sub-hourly through to daily rainfall. The methodologies were tested in South Africa, where an S-band radar operated relatively continuously at

  19. External calibration of polarimetric radars using point and distributed targets

    NASA Technical Reports Server (NTRS)

    Yueh, S. H.; Kong, J. A.; Shin, R. T.

    1991-01-01

    Polarimetric calibration algorithms using combinations of point targets and reciprocal distributed targets are developed. From the reciprocity relations of distributed targets, and equivalent point target response is derived. Then the problem of polarimetric calibration using two point targets and one distributed target reduces to that using three point targets, which has been previously solved. For calibration using one point target and one reciprocal distributed target, two cases are analyzed with the point target being a trihedral reflector or a polarimetric active radar calibrator (PARC). For both cases, the general solutions of the system distortion matrices are written as a product of a particular solution and a matrix with one free parameter. For the trihedral-reflector case, this free parameter is determined by assuming azimuthal symmetry for the distributed target. For the PARC case, knowledge of one ratio of two covariance matrix elements of the distributed target is required to solve for the free parameter. Numerical results are simulated to demonstrate the usefulness of the developed algorithms.

  20. External calibration of polarimetric radars using point and distributed targets

    NASA Astrophysics Data System (ADS)

    Yueh, S. H.; Kong, J. A.; Shin, R. T.

    1991-08-01

    Polarimetric calibration algorithms using combinations of point targets and reciprocal distributed targets are developed. From the reciprocity relations of distributed targets, and equivalent point target response is derived. Then the problem of polarimetric calibration using two point targets and one distributed target reduces to that using three point targets, which has been previously solved. For calibration using one point target and one reciprocal distributed target, two cases are analyzed with the point target being a trihedral reflector or a polarimetric active radar calibrator (PARC). For both cases, the general solutions of the system distortion matrices are written as a product of a particular solution and a matrix with one free parameter. For the trihedral-reflector case, this free parameter is determined by assuming azimuthal symmetry for the distributed target. For the PARC case, knowledge of one ratio of two covariance matrix elements of the distributed target is required to solve for the free parameter. Numerical results are simulated to demonstrate the usefulness of the developed algorithms.

  1. Cloud-Scale Vertical Velocity and Turbulent Dissipation Rate Retrievals

    DOE Data Explorer

    Shupe, Matthew

    2013-05-22

    Time-height fields of retrieved in-cloud vertical wind velocity and turbulent dissipation rate, both retrieved primarily from vertically-pointing, Ka-band cloud radar measurements. Files are available for manually-selected, stratiform, mixed-phase cloud cases observed at the North Slope of Alaska (NSA) site during periods covering the Mixed-Phase Arctic Cloud Experiment (MPACE, late September through early November 2004) and the Indirect and Semi-Direct Aerosol Campaign (ISDAC, April-early May 2008). These time periods will be expanded in a future submission.

  2. Observations of vertical velocities in the tropical upper troposphere and lower stratosphere using the Arecibo 430-MHz radar

    NASA Technical Reports Server (NTRS)

    Cornish, C. R.

    1988-01-01

    The first clear-air observations of vertical velocities in the tropical upper troposphere and lower stratosphere (8-22 km) using the Arecibo 430-MHz radar are presented. Oscillations in the vertical velocity near the Brunt-Vaisala period are observed in the lower stratosphere during the 12-hour observation period. Frequency power spectra from the vertical velocity time series show a slope between -0.5 and -1.0. Vertical wave number spectra computed from the height profiles of vertical velocities have slopes between -1.0 and -1.5. These observed slopes do not agree well with the slopes of +1/3 and -2.5 for frequency and vertical wave number spectra, respectively, predicted by a universal gravity-wave spectrum model. The spectral power of wave number spectra of a radial beam directed 15 deg off-zenith is enhanced by an order of magnitude over the spectral power levels of the vertical beam. This enhancement suggests that other geophysical processes besides gravity waves are present in the horizontal flow. The steepening of the wave number spectrum of the off-vertical beam in the lower stratosphere to near -2.0 is attributed to a quasi-inertial period wave, which was present in the horizontal flow during the observation period.

  3. Space Radar Image of Munich, Germany

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This spaceborne radar image of Munich, Germany illustrates the capability of a multi-frequency radar system to highlight different land use patterns in the area surrounding Bavaria's largest city. Central Munich is the white area at the middle of the image, on the banks of the Isar River. Pink areas are forested, while green areas indicate clear-cut and agricultural terrain. The Munich region served as a primary 'supersite' for studies in ecology, hydrology and radar calibration during the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) missions. Scientists were able to use these data to map patterns of forest damage from storms and areas affected by bark beetle infestation. The image was acquired by SIR-C/X-SAR onboard the space shuttle Endeavour on April 18, 1994. The image is 37 kilometers by 32 kilometers (23 miles by 20 miles) and is centered at 48.2 degrees North latitude, 11.5 degrees East longitude. North is toward the upper right. The colors are assigned to different radar frequencies and polarizations of the radar as follows: red is L-band, vertically transmitted and horizontally received; green is C-band, vertically transmitted and horizontally received; and blue is C-band vertically transmitted and received. SIR-C/X-SAR, a joint mission of the German, Italian, and United States space agencies, is part of NASA's Mission to Planet Earth.

  4. Three-dimensional mosaicking of the South Korean radar network

    NASA Astrophysics Data System (ADS)

    Berenguer, Marc; Sempere-Torres, Daniel; Lee, GyuWon

    2016-04-01

    Dense radar networks offer the possibility of improved Quantitative Precipitation Estimation thanks to the additional information collected in the overlapping areas, which allows mitigating errors associated with the Vertical Profile of Reflectivity or path attenuation by intense rain. With this aim, Roca-Sancho et al. (2014) proposed a technique to generate 3-D reflectivity mosaics from the multiple radars of a network. The technique is based on an inverse method that simulates the radar sampling of the atmosphere considering the characteristics (location, frequency and scanning protocol) of each individual radar. This technique has been applied to mosaic the observations of the radar network of South Korea (composed of 14 S-band radars), and integrate the observations of the small X-band network which to be installed near Seoul in the framework of a project funded by the Korea Agency for Infrastructure Technology Advancement (KAIA). The evaluation of the generated 3-D mosaics has been done by comparison with point measurements (i.e. rain gauges and disdrometers) and with the observations of independent radars. Reference: Roca-Sancho, J., M. Berenguer, and D. Sempere-Torres (2014), An inverse method to retrieve 3D radar reflectivity composites, Journal of Hydrology, 519, 947-965, doi: 10.1016/j.jhydrol.2014.07.039.

  5. Combination of radar and daily precipitation data to estimate meaningful sub-daily point precipitation extremes

    NASA Astrophysics Data System (ADS)

    Pegram, Geoff; Bardossy, Andras; Sinclair, Scott

    2017-04-01

    The use of radar measurements for the space time estimation of precipitation has for many decades been a central topic in hydro-meteorology. In this presentation we are interested specifically in daily and sub-daily extreme values of precipitation at gauged or ungauged locations which are important for design. The purpose of the presentation is to develop a methodology to combine daily precipitation observations and radar measurements to estimate sub-daily extremes at point locations. Radar data corrected using precipitation-reflectivity relationships lead to biased estimations of extremes. Different possibilities of correcting systematic errors using the daily observations are investigated. Observed gauged daily amounts are interpolated to un-sampled points and subsequently disaggregated using the sub-daily values obtained by the radar. Different corrections based on the spatial variability and the sub-daily entropy of scaled rainfall distributions are used to provide unbiased corrections of short duration extremes. In addition, a statistical procedure not based on a matching day by day correction is tested. In this last procedure, as we are only interested in rare extremes, low to medium values of rainfall depth were neglected leaving 12 days of ranked daily maxima in each set per year, whose sum typically comprises about 50% of each annual rainfall total. The sum of these 12 day maxima is first interpolated using a Kriging procedure. Subsequently this sum is disaggregated to daily values using a nearest neighbour procedure. The daily sums are then disaggregated by using the relative values of the biggest 12 radar based days in each year. Of course, the timings of radar and gauge maxima can be different, so the new method presented here uses radar for disaggregating daily gauge totals down to 15 min intervals in order to extract the maxima of sub-hourly through to daily rainfall. The methodologies were tested in South Africa, where an S-band radar operated

  6. Exploiting Cloud Radar Doppler Spectra of Mixed-Phase Clouds during ACCEPT Field Experiment to Identify Microphysical Processes

    NASA Astrophysics Data System (ADS)

    Kalesse, H.; Myagkov, A.; Seifert, P.; Buehl, J.

    2015-12-01

    Cloud radar Doppler spectra offer much information about cloud processes. By analyzing millimeter radar Doppler spectra from cloud-top to -base in mixed-phase clouds in which super-cooled liquid-layers are present we try to tell the microphysical evolution story of particles that are present by disentangling the contributions of the solid and liquid particles to the total radar returns. Instead of considering vertical profiles, dynamical effects are taken into account by following the particle population evolution along slanted paths which are caused by horizontal advection of the cloud. The goal is to identify regions in which different microphysical processes such as new particle formation (nucleation), water vapor deposition, aggregation, riming, or sublimation occurr. Cloud radar measurements are supplemented by Doppler lidar and Raman lidar observations as well as observations with MWR, wind profiler, and radio sondes. The presence of super-cooled liquid layers is identified by positive liquid water paths in MWR measurements, the vertical location of liquid layers (in non-raining systems and below lidar extinction) is derived from regions of high-backscatter and low depolarization in Raman lidar observations. In collocated cloud radar measurements, we try to identify cloud phase in the cloud radar Doppler spectrum via location of the Doppler peak(s), the existence of multi-modalities or the spectral skewness. Additionally, within the super-cooled liquid layers, the radar-identified liquid droplets are used as air motion tracer to correct the radar Doppler spectrum for vertical air motion w. These radar-derived estimates of w are validated by independent estimates of w from collocated Doppler lidar measurements. A 35 GHz vertically pointing cloud Doppler radar (METEK MIRA-35) in linear depolarization (LDR) mode is used. Data is from the deployment of the Leipzig Aerosol and Cloud Remote Observations System (LACROS) during the Analysis of the Composition of

  7. Fly eye radar or micro-radar sensor technology

    NASA Astrophysics Data System (ADS)

    Molchanov, Pavlo; Asmolova, Olga

    2014-05-01

    To compensate for its eye's inability to point its eye at a target, the fly's eye consists of multiple angularly spaced sensors giving the fly the wide-area visual coverage it needs to detect and avoid the threats around him. Based on a similar concept a revolutionary new micro-radar sensor technology is proposed for detecting and tracking ground and/or airborne low profile low altitude targets in harsh urban environments. Distributed along a border or around a protected object (military facility and buildings, camp, stadium) small size, low power unattended radar sensors can be used for target detection and tracking, threat warning, pre-shot sniper protection and provides effective support for homeland security. In addition it can provide 3D recognition and targets classification due to its use of five orders more pulses than any scanning radar to each space point, by using few points of view, diversity signals and intelligent processing. The application of an array of directional antennas eliminates the need for a mechanical scanning antenna or phase processor. It radically decreases radar size and increases bearing accuracy several folds. The proposed micro-radar sensors can be easy connected to one or several operators by point-to-point invisible protected communication. The directional antennas have higher gain, can be multi-frequency and connected to a multi-functional network. Fly eye micro-radars are inexpensive, can be expendable and will reduce cost of defense.

  8. The Coplane Analysis Technique for Three-Dimensional Wind Retrieval Using the HIWRAP Airborne Doppler Radar

    NASA Technical Reports Server (NTRS)

    Didlake, Anthony C., Jr.; Heymsfield, Gerald M.; Tian, Lin; Guimond, Stephen R.

    2015-01-01

    The coplane analysis technique for mapping the three-dimensional wind field of precipitating systems is applied to the NASA High Altitude Wind and Rain Airborne Profiler (HIWRAP). HIWRAP is a dual-frequency Doppler radar system with two downward pointing and conically scanning beams. The coplane technique interpolates radar measurements to a natural coordinate frame, directly solves for two wind components, and integrates the mass continuity equation to retrieve the unobserved third wind component. This technique is tested using a model simulation of a hurricane and compared to a global optimization retrieval. The coplane method produced lower errors for the cross-track and vertical wind components, while the global optimization method produced lower errors for the along-track wind component. Cross-track and vertical wind errors were dependent upon the accuracy of the estimated boundary condition winds near the surface and at nadir, which were derived by making certain assumptions about the vertical velocity field. The coplane technique was then applied successfully to HIWRAP observations of Hurricane Ingrid (2013). Unlike the global optimization method, the coplane analysis allows for a transparent connection between the radar observations and specific analysis results. With this ability, small-scale features can be analyzed more adequately and erroneous radar measurements can be identified more easily.

  9. a Point Cloud Classification Approach Based on Vertical Structures of Ground Objects

    NASA Astrophysics Data System (ADS)

    Zhao, Y.; Hu, Q.; Hu, W.

    2018-04-01

    This paper proposes a novel method for point cloud classification using vertical structural characteristics of ground objects. Since urbanization develops rapidly nowadays, urban ground objects also change frequently. Conventional photogrammetric methods cannot satisfy the requirements of updating the ground objects' information efficiently, so LiDAR (Light Detection and Ranging) technology is employed to accomplish this task. LiDAR data, namely point cloud data, can obtain detailed three-dimensional coordinates of ground objects, but this kind of data is discrete and unorganized. To accomplish ground objects classification with point cloud, we first construct horizontal grids and vertical layers to organize point cloud data, and then calculate vertical characteristics, including density and measures of dispersion, and form characteristic curves for each grids. With the help of PCA processing and K-means algorithm, we analyze the similarities and differences of characteristic curves. Curves that have similar features will be classified into the same class and point cloud correspond to these curves will be classified as well. The whole process is simple but effective, and this approach does not need assistance of other data sources. In this study, point cloud data are classified into three classes, which are vegetation, buildings, and roads. When horizontal grid spacing and vertical layer spacing are 3 m and 1 m respectively, vertical characteristic is set as density, and the number of dimensions after PCA processing is 11, the overall precision of classification result is about 86.31 %. The result can help us quickly understand the distribution of various ground objects.

  10. Spatial judgments in the horizontal and vertical planes from different vantage points.

    PubMed

    Prytz, Erik; Scerbo, Mark W

    2012-01-01

    Todorović (2008 Perception 37 106-125) reported that there are systematic errors in the perception of 3-D space when viewing 2-D linear perspective drawings depending on the observer's vantage point. Because these findings were restricted to the horizontal plane, the current study was designed to determine the nature of these errors in the vertical plane. Participants viewed an image containing multiple colonnades aligned on parallel converging lines receding to a vanishing point. They were asked to judge where, in the physical room, the next column should be placed. The results support Todorović in that systematic deviations in the spatial judgments depended on vantage point for both the horizontal and vertical planes. However, there are also marked differences between the two planes. While judgments in both planes failed to compensate adequately for the vantage-point shift, the vertical plane induced greater distortions of the stimulus image itself within each vantage point.

  11. The Multiple Doppler Radar Workshop, November 1979.

    NASA Astrophysics Data System (ADS)

    Carbone, R. E.; Harris, F. I.; Hildebrand, P. H.; Kropfli, R. A.; Miller, L. J.; Moninger, W.; Strauch, R. G.; Doviak, R. J.; Johnson, K. W.; Nelson, S. P.; Ray, P. S.; Gilet, M.

    1980-10-01

    the dual Doppler and multiple Doppler cases. Various filters and techniques, including statistical and variational approaches, are mentioned. Emphasis is placed on the importance of experiment design and procedures, technological improvements, incorporation of all information from supporting sensors, and analysis priority for physically simple cases. Integrated reliability is proposed as an objective tool for radar siting.Verification of multiple Doppler-derived vertical velocity is discussed in Part V. Three categories of verification are defined as direct, deductive, and theoretical/numerical. Direct verification consists of zenith-pointing radar measurements (from either airborne or ground-based systems), air motion sensing aircraft, instrumented towers, and tracking of radar chaff. Deductive sources include mesonetworks, aircraft (thermodynamic and microphysical) measurements, satellite observations, radar reflectivity, multiple Doppler consistency, and atmospheric soundings. Theoretical/numerical sources of verification include proxy data simulation, momentum checking, and numerical cloud models. New technology, principally in the form of wide bandwidth radars, is seen as a development that may reduce the need for extensive verification of multiple Doppler-derived vertical air motions. Airborne Doppler radar is perceived as the single most important source of verification within the bounds of existing technology.Nine stages of data processing and display are identified in Part VI. The stages are identified as field checks, archival, selection, editing, coordinate transformation, synthesis of Cartesian fields, filtering, display, and physical analysis. Display of data is considered to be a problem critical to assimilation of data at all stages. Interactive computing systems and software are concluded to be very important, particularly for the editing stage. Three- and 4-dimensional displays are considered essential for data assimilation, particularly at the

  12. The EDOP radar system on the high-altitude NASA ER-2 aircraft

    USGS Publications Warehouse

    Heymsfield, G.M.; Bidwell, S.W.; Caylor, I.J.; Ameen, S.; Nicholson, S.; Boncyk, W.; Miller, L.; Vandemark, D.; Racette, P.E.; Dod, L.R.

    1996-01-01

    The NASA ER-2 high-altitude (20 km) aircraft that emulates a satellite view of precipitation systems carries a variety of passive and active (lidar) remote sensing instruments. A new Doppler weather radar system at X band (9.6 GHz) called the ER-2 Doppler radar (EDOP) has been developed and flown on the ER-2 aircraft. EDOP is a fully coherent Doppler weather radar with fixed nadir and forward pointing (33?? off nadir) beams that map out Doppler winds and reflectivities in the vertical plane along the aircraft motion vector. Doppler winds from the two beams can be used to derive vertical and along-track air motions. In addition, the forward beam provides linear depolarization measurements that are useful in discriminating microphysical characteristics of the precipitation. This paper deals with a general description of the EDOP instrument including the measurement concept, the system configuration and hardware, and recently obtained data examples from the instrument. The combined remote sensing package on the ER-2, along with EDOP, provides a unique platform for simulating spaceborne remote sensing of precipitation.

  13. W-band spaceborne radar observations of atmospheric river events

    NASA Astrophysics Data System (ADS)

    Matrosov, S. Y.

    2010-12-01

    While the main objective of the world first W-band radar aboard the CloudSat satellite is to provide vertically resolved information on clouds, it proved to be a valuable tool for observing precipitation. The CloudSat radar is generally able to resolve precipitating cloud systems in their vertical entirety. Although measurements from the liquid hydrometer layer containing rainfall are strongly attenuated, special retrieval approaches can be used to estimate rainfall parameters. These approaches are based on vertical gradients of observed radar reflectivity factor rather than on absolute estimates of reflectivity. Concurrent independent estimations of ice cloud parameters in the same vertical column allow characterization of precipitating systems and provide information on coupling between clouds and rainfall they produce. The potential of CloudSat for observations atmospheric river events affecting the West Coast of North America is evaluated. It is shown that spaceborne radar measurements can provide high resolution information on the height of the freezing level thus separating areas of rainfall and snowfall. CloudSat precipitation rate estimates complement information from the surface-based radars. Observations of atmospheric rivers at different locations above the ocean and during landfall help to understand evolutions of atmospheric rivers and their structures.

  14. Radar monitoring of oil pollution

    NASA Technical Reports Server (NTRS)

    Guinard, N. W.

    1970-01-01

    Radar is currently used for detecting and monitoring oil slicks on the sea surface. The four-frequency radar system is used to acquire synthetic aperature imagery of the sea surface on which the oil slicks appear as a nonreflecting area on the surface surrounded by the usual sea return. The value of this technique was demonstrated, when the four-frequency radar system was used to image the oil spill of tanker which has wrecked. Imagery was acquired on both linear polarization (horizontal, vertical) for frequencies of 428, 1228, and 8910 megahertz. Vertical returns strongly indicated the presence of oil while horizontal returns failed to detect the slicks. Such a result is characteristic of the return from the sea and cannot presently be interpreted as characteristics of oil spills. Because an airborne imaging radar is capable of providing a wide-swath coverage under almost all weather conditions, it offers promise in the development of a pollution-monitoring system that can provide a coastal watch for oil slicks.

  15. The NASA radar entomology program at Wallops Flight Center

    NASA Technical Reports Server (NTRS)

    Vaughn, C. R.

    1979-01-01

    NASA contribution to radar entomology is presented. Wallops Flight Center is described in terms of its radar systems. Radar tracking of birds and insects was recorded from helicopters for airspeed and vertical speed.

  16. Space Radar Image of Mississippi Delta

    NASA Image and Video Library

    1999-04-15

    This is a radar image of the Mississippi River Delta where the river enters into the Gulf of Mexico along the coast of Louisiana. This multi-frequency image demonstrates the capability of the radar to distinguish different types of wetlands surfaces in river deltas. This image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) aboard the space shuttle Endeavour on October 2, 1995. The image is centered on latitude 29.3 degrees North latitude and 89.28 degrees West longitude. The area shown is approximately 63 kilometers by 43 kilometers (39 miles by 26 miles). North is towards the upper right of the image. As the river enters the Gulf of Mexico, it loses energy and dumps its load of sediment that it has carried on its journey through the mid-continent. This pile of sediment, or mud, accumulates over the years building up the delta front. As one part of the delta becomes clogged with sediment, the delta front will migrate in search of new areas to grow. The area shown on this image is the currently active delta front of the Mississippi. The migratory nature of the delta forms natural traps for oil and the numerous bright spots along the outside of the delta are drilling platforms. Most of the land in the image consists of mud flats and marsh lands. There is little human settlement in this area due to the instability of the sediments. The main shipping channel of the Mississippi River is the broad red stripe running northwest to southeast down the left side of the image. The bright spots within the channel are ships. The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band vertically transmitted, vertically received; green is C-band vertically transmitted, vertically received; blue is X-band vertically transmitted, vertically received. http://photojournal.jpl.nasa.gov/catalog/PIA01784

  17. Shuttle orbiter radar cross-sectional analysis

    NASA Technical Reports Server (NTRS)

    Cooper, D. W.; James, R.

    1979-01-01

    Theoretical and model simulation studies on signal to noise levels and shuttle radar cross section are described. Pre-mission system calibrations, system configuration, and postmission system calibration of the tracking radars are described. Conversion of target range, azimuth, and elevation into radar centered east north vertical position coordinates are evaluated. The location of the impinging rf energy with respect to the target vehicles body axis triad is calculated. Cross section correlation between the two radars is presented.

  18. Convective cloud vertical velocity and mass-flux characteristics from radar wind profiler observations during GoAmazon2014/5: VERTICAL VELOCITY GOAMAZON2014/5

    DOE PAGES

    Giangrande, Scott E.; Toto, Tami; Jensen, Michael P.; ...

    2016-11-15

    A radar wind profiler data set collected during the 2 year Department of Energy Atmospheric Radiation Measurement Observations and Modeling of the Green Ocean Amazon (GoAmazon2014/5) campaign is used to estimate convective cloud vertical velocity, area fraction, and mass flux profiles. Vertical velocity observations are presented using cumulative frequency histograms and weighted mean profiles to provide insights in a manner suitable for global climate model scale comparisons (spatial domains from 20 km to 60 km). Convective profile sensitivity to changes in environmental conditions and seasonal regime controls is also considered. Aggregate and ensemble average vertical velocity, convective area fraction, andmore » mass flux profiles, as well as magnitudes and relative profile behaviors, are found consistent with previous studies. Updrafts and downdrafts increase in magnitude with height to midlevels (6 to 10 km), with updraft area also increasing with height. Updraft mass flux profiles similarly increase with height, showing a peak in magnitude near 8 km. Downdrafts are observed to be most frequent below the freezing level, with downdraft area monotonically decreasing with height. Updraft and downdraft profile behaviors are further stratified according to environmental controls. These results indicate stronger vertical velocity profile behaviors under higher convective available potential energy and lower low-level moisture conditions. Sharp contrasts in convective area fraction and mass flux profiles are most pronounced when retrievals are segregated according to Amazonian wet and dry season conditions. During this deployment, wet season regimes favored higher domain mass flux profiles, attributed to more frequent convection that offsets weaker average convective cell vertical velocities.« less

  19. Exploring Stratocumulus Cloud-Top Entrainment Processes and Parameterizations by Using Doppler Cloud Radar Observations

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

    Albrecht, Bruce; Fang, Ming; Ghate, Virendra

    2016-02-01

    Observations from an upward-pointing Doppler cloud radar are used to examine cloud-top entrainment processes and parameterizations in a non-precipitating continental stratocumulus cloud deck maintained by time varying surface buoyancy fluxes and cloud-top radiative cooling. Radar and ancillary observations were made at the Atmospheric Radiation Measurement (ARM)’s Southern Great Plains (SGP) site located near Lamont, Oklahoma of unbroken, non-precipitating stratocumulus clouds observed for a 14-hour period starting 0900 Central Standard Time on 25 March 2005. The vertical velocity variance and energy dissipation rate (EDR) terms in a parameterized turbulence kinetic energy (TKE) budget of the entrainment zone are estimated using themore » radar vertical velocity and the radar spectrum width observations from the upward-pointing millimeter cloud radar (MMCR) operating at the SGP site. Hourly averages of the vertical velocity variance term in the TKE entrainment formulation correlates strongly (r=0.72) to the dissipation rate term in the entrainment zone. However, the ratio of the variance term to the dissipation decreases at night due to decoupling of the boundary layer. When the night -time decoupling is accounted for, the correlation between the variance and the EDR term increases (r=0.92). To obtain bulk coefficients for the entrainment parameterizations derived from the TKE budget, independent estimate of entrainment were obtained from an inversion height budget using ARM SGP observations of the local time derivative and the horizontal advection of the cloud-top height. The large-scale vertical velocity at the inversion needed for this budget from EMWF reanalysis. This budget gives a mean entrainment rate for the observing period of 0.76±0.15 cm/s. This mean value is applied to the TKE budget parameterizations to obtain the bulk coefficients needed in these parameterizations. These bulk coefficients are compared with those from previous and are used to in

  20. Analysis of 35 GHz Cloud Radar polarimetric variables to identify stratiform and convective precipitation.

    NASA Astrophysics Data System (ADS)

    Fontaine, Emmanuel; Illingworth, Anthony, J.; Stein, Thorwald

    2017-04-01

    This study is performed using vertical profiles of radar measurements at 35GHz, for the period going from 29th of February to 1rst October 2016, at the Chilbolton observatory in United Kingdom. During this period, more than 40 days with precipitation events are investigated. The investigation uses the synergy of radar reflectivity factors, vertical velocity, Doppler spectrum width, and linear depolarization ratio (LDR) to differentiate between stratiform and convective rain events. The depth of the layer with Doppler spectrum width values greater than 0.5 m s-1 is shown to be a suitable proxy to distinguish between convective and stratiform events. Using LDR to detect the radar bright band, bright band characteristics such as depth of the layer and maximum LDR are shown to vary with the amount of turbulence aloft. Profiles of radar measurements are also compared to rain gauge measurements to study the contribution of convective and stratiform rainfall to total rain duration and amount. To conclude, this study points out differences between convective and stratiform rains and quantifies their contributions over a precipitation event, highlighting that convective and stratiform rainfall should be considered as a continuum rather than a dichotomy.

  1. GPM and TRMM Radar Vertical Profiles and Impact on Large-scale Variations of Surface Rain

    NASA Astrophysics Data System (ADS)

    Wang, J. J.; Adler, R. F.

    2017-12-01

    Previous studies by the authors using Tropical Rainfall Measuring Mission (TRMM) and Global Precipitation Measurement (GPM) data have shown that TRMM Precipitation Radar (PR) and GPM Dual-Frequency Precipitation Radar (DPR) surface rain estimates do not have corresponding amplitudes of inter-annual variations over the tropical oceans as do passive microwave observations by TRMM Microwave Imager (TMI) and GPM Microwave Imager (GMI). This includes differences in surface temperature-rainfall variations. We re-investigate these relations with the new GPM Version 5 data with an emphasis on understanding these differences with respect to the DPR vertical profiles of reflectivity and rainfall and the associated convective and stratiform proportions. For the inter-annual variation of ocean rainfall from both passive microwave (TMI and GMI) and active microwave (PR and DPR) estimates, it is found that for stratiform rainfall both TMI-PR and GMI-DPR show very good correlation. However, the correlation of GMI-DPR is much higher than TMI-PR in convective rainfall. The analysis of vertical profile of PR and DPR rainfall during the TRMM and GPM overlap period (March-August, 2014) reveals that PR and DPR have about the same rainrate at 4km and above, but PR rainrate is more than 10% lower that of DPR at the surface. In other words, it seems that convective rainfall is better defined with DPR near surface. However, even though the DPR results agree better with the passive microwave results, there still is a significant difference, which may be a result of DPR retrieval error, or inherent passive/active retrieval differences. Monthly and instantaneous GMI and DPR data need to be analyzed in details to better understand the differences.

  2. Development of High Altitude UAV Weather Radars for Hurricane Research

    NASA Technical Reports Server (NTRS)

    Heymsfield, Gerald; Li, Li-Hua

    2005-01-01

    A proposed effort within NASA called (ASHE) over the past few years was aimed at studying the genesis of tropical disturbances off the east coast of Africa. This effort was focused on using an instrumented Global Hawk UAV with high altitude (%Ok ft) and long duration (30 h) capability. While the Global Hawk availability remains uncertain, development of two relevant instruments, a Doppler radar (URAD - UAV Radar) and a backscatter lidar (CPL-UAV - Cloud Physics Lidar), are in progress. The radar to be discussed here is based on two previous high-altitude, autonomously operating radars on the NASA ER-2 aircraft, the ER-2 Doppler Radar (EDOP) at X-band (9.6 GHz), and the Cloud Radar System (CRS) at W- band (94 GHz). The nadir-pointing EDOP and CRS radars profile vertical reflectivity structure and vertical Doppler winds in precipitation and clouds, respectively. EDOP has flown in all of the CAMEX flight series to study hurricanes over storms such as Hurricanes Bonnie, Humberto, Georges, Erin, and TS Chantal. These radars were developed at Goddard over the last decade and have been used for satellite algorithm development and validation (TRMM and Cloudsat), and for hurricane and convective storm research. We describe here the development of URAD that will measure wind and reflectivity in hurricanes and other weather systems from a top down, high-altitude view. URAD for the Global Hawk consists of two subsystems both of which are at X-band (9.3-9.6 GHz) and Doppler: a nadir fixed-beam Doppler radar for vertical motion and precipitation measurement, and a Conical scanning radar for horizontal winds in cloud and at the surface, and precipitation structure. These radars are being designed with size, weight, and power consumption suitable for the Global Hawk and other UAV's. The nadir radar uses a magnetron transmitter and the scanning radar uses a TWT transmitter. With conical scanning of the radar at a 35" incidence angle over an ocean surface in the absence of

  3. Radar Sounder

    DTIC Science & Technology

    1988-09-01

    S’ardard Form 298 Rev 2-89) • " Del " 1 , -iNS, 19 , q f .If - ACKNOWLEDGMENTS The authors would like to acknowledge the support of numerous...plates, etc.); estimation of rain rate and the observation of the horizontal and vertical structure of rain. The data from the radar sounder will be...crytal habit. The microphysical properties and vertical structure of the clouds are needed for applications of interest to the Air Force such as

  4. Cross-hole radar scanning of two vertical, permeable, reactive-iron walls at the Massachusetts Military Reservation, Cape Cod, Massachusetts

    USGS Publications Warehouse

    Lane, J.W.; Joesten, P.K.; Savoie, J.G.

    2001-01-01

    A pilot-scale study was conducted by the U.S. Army National Guard (USANG) at the Massachusetts Military Reservation (MMR) on Cape Cod, Massachusetts, to assess the use of a hydraulic-fracturing method to create vertical, permeable walls of zero-valent iron to passively remediate ground water contaminated with chlorinated solvents. The study was conducted near the source area of the Chemical Spill-10 (CS-10) plume, a plume containing chlorinated solvents that underlies the MMR. Ground-water contamination near the source area extends from about 24 m (meters) to 35 m below land surface. The USANG designed two reactive-iron walls to be 12 m long and positioned 24 to 37 m below land surface to intersect and remediate part of the CS-10 plume.Because iron, as an electrical conductor, absorbs electromagnetic energy, the US Geological Survey used a cross-hole common-depth, radar scanning method to assess the continuity and to estimate the lateral and vertical extent of the two reactive-iron walls. The cross-hole radar surveys were conducted in boreholes on opposite sides of the iron injection zones using electric-dipole antennas with dominant center frequencies of 100 and 250 MHz. Significant decreases in the radar-pulse amplitudes observed in scans that traversed the injection zones were interpreted by comparing field data to results of two-dimensional finite-difference time-domain numerical models and laboratory-scale physical models.The numerical and physical models simulate a wall of perfectly conducting material embedded in saturated sand. Results from the numerical and physical models show that the amplitude of the radar pulse transmitted across the edge of a conductive wall is about 43 percent of the amplitude of a radar pulse transmitted across background material. The amplitude of a radar pulse transmitted through a hole in a conductive wall increases as the aperture of the hole increases. The modeling results indicate that holes with an aperture of less than 40

  5. The calibration of an HF radar used for ionospheric research

    NASA Astrophysics Data System (ADS)

    From, W. R.; Whitehead, J. D.

    1984-02-01

    The HF radar on Bribie Island, Australia, uses crossed-fan beams produced by crossed linear transmitter and receiver arrays of 10 elements each to simulate a pencil beam. The beam points vertically when all the array elements are in phase, and is steerable by up to 20 deg off vertical at the central one of the three operating frequencies. Phase and gain changes within the transmitters and receivers are compensated for by an automatic system of adjustment. The 10 transmitting antennas are, as nearly as possible, physically identical as are the 10 receiving antennas. Antenna calibration using high flying aircraft or satellites is not possible. A method is described for using the ionospheric reflections to measure the polar diagram and also to correct for errors in the direction of pointing.

  6. Frequency-Tracking CW Doppler Radar Solving Small-Angle Approximation and Null Point Issues in Non-Contact Vital Signs Monitoring.

    PubMed

    Mercuri, Marco; Liu, Yao-Hong; Lorato, Ilde; Torfs, Tom; Bourdoux, Andre; Van Hoof, Chris

    2017-06-01

    A Doppler radar operating as a Phase-Locked-Loop (PLL) in frequency demodulator configuration is presented and discussed. The proposed radar presents a unique architecture, using a single channel mixer, and allows to detect contactless vital signs parameters while solving the null point issue and without requiring the small angle approximation condition. Spectral analysis, simulations, and experimental results are presented and detailed to demonstrate the feasibility and the operational principle of the proposed radar architecture.

  7. Space Radar Image of Sakura-Jima Volcano, Japan

    NASA Technical Reports Server (NTRS)

    1994-01-01

    The active volcano Sakura-Jima on the island of Kyushu, Japan is shown in the center of this radar image. The volcano occupies the peninsula in the center of Kagoshima Bay, which was formed by the explosion and collapse of an ancient predecessor of today's volcano. The volcano has been in near continuous eruption since 1955. Its explosions of ash and gas are closely monitored by local authorities due to the proximity of the city of Kagoshima across a narrow strait from the volcano's center, shown below and to the left of the central peninsula in this image. City residents have grown accustomed to clearing ash deposits from sidewalks, cars and buildings following Sakura-jima's eruptions. The volcano is one of 15 identified by scientists as potentially hazardous to local populations, as part of the international 'Decade Volcano' program. The image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 9, 1994. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered at 31.6 degrees North latitude and 130.6 degrees East longitude. North is toward the upper left. The area shown measures 37.5 kilometers by 46.5 kilometers (23.3 miles by 28.8 miles). The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band vertically transmitted, vertically received; green is the average of L-band vertically transmitted, vertically received and C-band vertically transmitted, vertically received; blue is C-band vertically transmitted, vertically received.

  8. Space Radar Image of Sakura-Jima Volcano, Japan

    NASA Image and Video Library

    1999-04-15

    The active volcano Sakura-Jima on the island of Kyushu, Japan is shown in the center of this radar image. The volcano occupies the peninsula in the center of Kagoshima Bay, which was formed by the explosion and collapse of an ancient predecessor of today's volcano. The volcano has been in near continuous eruption since 1955. Its explosions of ash and gas are closely monitored by local authorities due to the proximity of the city of Kagoshima across a narrow strait from the volcano's center, shown below and to the left of the central peninsula in this image. City residents have grown accustomed to clearing ash deposits from sidewalks, cars and buildings following Sakura-jima's eruptions. The volcano is one of 15 identified by scientists as potentially hazardous to local populations, as part of the international "Decade Volcano" program. The image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 9, 1994. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered at 31.6 degrees North latitude and 130.6 degrees East longitude. North is toward the upper left. The area shown measures 37.5 kilometers by 46.5 kilometers (23.3 miles by 28.8 miles). The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band vertically transmitted, vertically received; green is the average of L-band vertically transmitted, vertically received and C-band vertically transmitted, vertically received; blue is C-band vertically transmitted, vertically received. http://photojournal.jpl.nasa.gov/catalog/PIA01777

  9. Identification and uncertainty estimation of vertical reflectivity profiles using a Lagrangian approach to support quantitative precipitation measurements by weather radar

    NASA Astrophysics Data System (ADS)

    Hazenberg, P.; Torfs, P. J. J. F.; Leijnse, H.; Delrieu, G.; Uijlenhoet, R.

    2013-09-01

    This paper presents a novel approach to estimate the vertical profile of reflectivity (VPR) from volumetric weather radar data using both a traditional Eulerian as well as a newly proposed Lagrangian implementation. For this latter implementation, the recently developed Rotational Carpenter Square Cluster Algorithm (RoCaSCA) is used to delineate precipitation regions at different reflectivity levels. A piecewise linear VPR is estimated for either stratiform or neither stratiform/convective precipitation. As a second aspect of this paper, a novel approach is presented which is able to account for the impact of VPR uncertainty on the estimated radar rainfall variability. Results show that implementation of the VPR identification and correction procedure has a positive impact on quantitative precipitation estimates from radar. Unfortunately, visibility problems severely limit the impact of the Lagrangian implementation beyond distances of 100 km. However, by combining this procedure with the global Eulerian VPR estimation procedure for a given rainfall type (stratiform and neither stratiform/convective), the quality of the quantitative precipitation estimates increases up to a distance of 150 km. Analyses of the impact of VPR uncertainty shows that this aspect accounts for a large fraction of the differences between weather radar rainfall estimates and rain gauge measurements.

  10. Space Radar Image of Mississippi Delta

    NASA Technical Reports Server (NTRS)

    1999-01-01

    This is a radar image of the Mississippi River Delta where the river enters into the Gulf of Mexico along the coast of Louisiana. This multi-frequency image demonstrates the capability of the radar to distinguish different types of wetlands surfaces in river deltas. This image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) aboard the space shuttle Endeavour on October 2, 1995. The image is centered on latitude 29.3 degrees North latitude and 89.28 degrees West longitude. The area shown is approximately 63 kilometers by 43 kilometers (39 miles by 26 miles). North is towards the upper right of the image. As the river enters the Gulf of Mexico, it loses energy and dumps its load of sediment that it has carried on its journey through the mid-continent. This pile of sediment, or mud, accumulates over the years building up the delta front. As one part of the delta becomes clogged with sediment, the delta front will migrate in search of new areas to grow. The area shown on this image is the currently active delta front of the Mississippi. The migratory nature of the delta forms natural traps for oil and the numerous bright spots along the outside of the delta are drilling platforms. Most of the land in the image consists of mud flats and marsh lands. There is little human settlement in this area due to the instability of the sediments. The main shipping channel of the Mississippi River is the broad red stripe running northwest to southeast down the left side of the image. The bright spots within the channel are ships. The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band vertically transmitted, vertically received; green is C-band vertically transmitted, vertically received; blue is X-band vertically transmitted, vertically received. Spaceborne Imaging Radar-C and X-band Synthetic Aperture Radar (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The radars

  11. Vertical transport of Kelut volcanic stratospheric aerosols observed by the equatorial lidar and the Equatorial Atmosphere Radar

    NASA Astrophysics Data System (ADS)

    Nagasawa, C.; Abo, M.; Shibata, Y.

    2017-12-01

    The transport of substance between stratosphere and troposphere in the equatorial region makes an impact to the global climate change, but it has a lot of unknown behaviors. We have performed the lidar observations for survey of atmospheric structure of troposphere, stratosphere, and mesosphere over Kototabang (0.2S, 100.3E), Indonesia in the equatorial region since 2004. Kelut volcano (7.9S, 112.3E) in the Java island of Indonesia erupted on 13 February 2014. The CALIOP observed that the eruption cloud reached 26km above sea level in the tropical stratosphere, but most of the plume remained at 19-20 km over the tropopause. By CALIOP data analysis, aerosol clouds spread in the longitude direction with the lapse of time and arrived at equator in 5 days. After aerosol clouds reached equator, they moved towards the east along the equator by strong eastward equatorial wind of QBO. In June 2014 (4 months after the eruption), aerosol transport from the stratosphere to the troposphere were observed by the polarization lidar at Kototabang. At the same time, we can clearly see down phase structure of vertical wind velocity observed by EAR (Equatorial Atmosphere Radar) generated by the equatorial Kelvin wave. We investigate the transport of substance between stratosphere and troposphere in the equatorial region by data which have been collected by the polarization lidar at Kototabang and the EAR after Kelut volcano eruption. Using combination of ground based lidar, satellite based lidar, and atmosphere radar, we can get valuable evidence of equatorial transport of substance between the troposphere and the lower stratosphere. This work was supported by Collaborative Research based on MU Radar and Equatorial Atmosphere Radar.

  12. Two loop QCD vertices at the symmetric point

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

    Gracey, J. A.

    2011-10-15

    We compute the triple gluon, quark-gluon and ghost-gluon vertices of QCD at the symmetric subtraction point at two loops in the MS scheme. In addition we renormalize each of the three vertices in their respective momentum subtraction schemes, MOMggg, MOMq and MOMh. The conversion functions of all the wave functions, coupling constant and gauge parameter renormalization constants of each of the schemes relative to MS are determined analytically. These are then used to derive the three loop anomalous dimensions of the gluon, quark, Faddeev-Popov ghost and gauge parameter as well as the {beta} function in an arbitrary linear covariant gaugemore » for each MOM scheme. There is good agreement of the latter with earlier Landau gauge numerical estimates of Chetyrkin and Seidensticker.« less

  13. Comparison between S. T. radar and in situ balloon measurements

    NASA Technical Reports Server (NTRS)

    Dalaudier, F.; Barat, J.; Bertin, F.; Brun, E.; Crochet, M.; Cuq, F.

    1986-01-01

    A campaign for simultaneous in situ and remote observation of both troposphere and stratosphere took place near Aire-sur-l'Adour (in southeastern France) on May 4, 1984. The aim of this campaign was a better understanding of the physics of radar echoes. The backscattered signal obtained with a stratosphere-troposphere radar both at the vertical and 15 deg. off vertical is compared with the velocity and temperature measurements made in the same region (about 10 km north of the radar site) by balloon-borne ionic anenometers and temperature sensors. In situ measurements clearly indicate that the temperature fluctuations are not always consistent with the standard turbulent theory. Nevertheless, the assumptions generally made (isotropy and turbulent field in k) and the classical formulation so derived for radar reflectivity are able to reproduce the shape of the radar return power profiles in oblique directions. Another significant result is the confirmation of the role played by the atmospheric stratification in the vertical echo power. It is important to develop these simultaneous in situ and remote experiments for a better description of the dynamical and thermal structure of the atmosphere and for a better understanding of the mechanisms governing clear-air radar reflectivity.

  14. Space Radar Image of Long Island Optical/Radar

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This pair of images of the Long Island, New York region is a comparison of an optical photograph (top) and a radar image (bottom), both taken in darkness in April 1994. The photograph at the top was taken by the Endeavour astronauts at about 3 a.m. Eastern time on April 20, 1994. The image at the bottom was acquired at about the same time four days earlier on April 16,1994 by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) system aboard the space shuttle Endeavour. Both images show an area approximately 100 kilometers by 40 kilometers (62 miles by 25 miles) that is centered at 40.7 degrees North latitude and 73.5 degrees West longitude. North is toward the upper right. The optical image is dominated by city lights, which are particularly bright in the densely developed urban areas of New York City located on the left half of the photo. The brightest white zones appear on the island of Manhattan in the left center, and Central Park can be seen as a darker area in the middle of Manhattan. To the northeast (right) of the city, suburban Long Island appears as a less densely illuminated area, with the brightest zones occurring along major transportation and development corridors. Since radar is an active sensing system that provides its own illumination, the radar image shows a great amount of surface detail, despite the night-time acquisition. The colors in the radar image were obtained using the following radar channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and vertically received); blue represents the C-band (horizontally transmitted and vertically received). In this image, the water surface - the Atlantic Ocean along the bottom edge and Long Island Sound shown at the top edge - appears red because small waves at the surface strongly reflect the horizontally transmitted and received L-band radar signal. Networks of highways and railroad lines are clearly

  15. Space Radar Image of Long Island Optical/Radar

    NASA Image and Video Library

    1999-05-01

    This pair of images of the Long Island, New York region is a comparison of an optical photograph (top) and a radar image (bottom), both taken in darkness in April 1994. The photograph at the top was taken by the Endeavour astronauts at about 3 a.m. Eastern time on April 20, 1994. The image at the bottom was acquired at about the same time four days earlier on April 16,1994 by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) system aboard the space shuttle Endeavour. Both images show an area approximately 100 kilometers by 40 kilometers (62 miles by 25 miles) that is centered at 40.7 degrees North latitude and 73.5 degrees West longitude. North is toward the upper right. The optical image is dominated by city lights, which are particularly bright in the densely developed urban areas of New York City located on the left half of the photo. The brightest white zones appear on the island of Manhattan in the left center, and Central Park can be seen as a darker area in the middle of Manhattan. To the northeast (right) of the city, suburban Long Island appears as a less densely illuminated area, with the brightest zones occurring along major transportation and development corridors. Since radar is an active sensing system that provides its own illumination, the radar image shows a great amount of surface detail, despite the night-time acquisition. The colors in the radar image were obtained using the following radar channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and vertically received); blue represents the C-band (horizontally transmitted and vertically received). In this image, the water surface - the Atlantic Ocean along the bottom edge and Long Island Sound shown at the top edge - appears red because small waves at the surface strongly reflect the horizontally transmitted and received L-band radar signal. Networks of highways and railroad lines are clearly

  16. A 34-meter VAWT (Vertical Axis Wind Turbine) point design

    NASA Astrophysics Data System (ADS)

    Ashwill, T. D.; Berg, D. E.; Dodd, H. M.; Rumsey, M. A.; Sutherland, H. J.; Veers, P. S.

    The Wind Energy Division at Sandia National Laboratories recently completed a point design based on the 34-m Vertical Axis Wind Turbine (VAWT) Test Bed. The 34-m Test Bed research machine incorporates several innovations that improve Darrieus technology, including increased energy production, over previous machines. The point design differs minimally from the Test Bed; but by removing research-related items, its estimated cost is substantially reduced. The point design is a first step towards a Test-Bed-based commercial machine that would be competitive with conventional sources of power in the mid-1990s.

  17. Preliminary radar systems analysis for Venus orbiter missions

    NASA Technical Reports Server (NTRS)

    Brandenburg, R. K.; Spadoni, D. J.

    1971-01-01

    A short, preliminary analysis is presented of the problems involved in mapping the surface of Venus with radar from an orbiting spacecraft. Two types of radar, the noncoherent sidelooking and the focused synthetic aperture systems, are sized to fulfill two assumed levels of Venus exploration. The two exploration levels, regional and local, assumed for this study are based on previous Astro Sciences work (Klopp 1969). The regional level is defined as 1 to 3 kilometer spatial and 0.5 to 1 km vertical resolution of 100 percent 0 of the planet's surface. The local level is defined as 100 to 200 meter spatial and 50-10 m vertical resolution of about 100 percent of the surfAce (based on the regional survey). A 10cm operating frequency was chosen for both radar systems in order to minimize the antenna size and maximize the apparent radar cross section of the surface.

  18. Experimental and theoretical determination of sea-state bias in radar altimetry

    NASA Technical Reports Server (NTRS)

    Stewart, Robert H.

    1991-01-01

    The major unknown error in radar altimetry is due to waves on the sea surface which cause the mean radar-reflecting surface to be displaced from mean sea level. This is the electromagnetic bias. The primary motivation for the project was to understand the causes of the bias so that the error it produces in radar altimetry could be calculated and removed from altimeter measurements made from space by the Topex/Poseidon altimetric satellite. The goals of the project were: (1) observe radar scatter at vertical incidence using a simple radar on a platform for a wide variety of environmental conditions at the same time wind and wave conditions were measured; (2) calculate electromagnetic bias from the radar observations; (3) investigate the limitations of the present theory describing radar scatter at vertical incidence; (4) compare measured electromagnetic bias with bias calculated from theory using measurements of wind and waves made at the time of the radar measurements; and (5) if possible, extend the theory so bias can be calculated for a wider range of environmental conditions.

  19. 3D And 4D Cloud Lifecycle Investigations Using Innovative Scanning Radar Analysis Methods. Final report

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

    Kollias, Pavlos

    2017-04-23

    With the vast upgrades to the ARM program radar measurement capabilities in 2010 and beyond, our ability to probe the 3D structure of clouds and associated precipitation has increased dramatically. This project build on the PI's and co-I's expertisein the analysis of radar observations. The first research thrust aims to document the 3D morphological (as depicted by the radar reflectivity structure) and 3D dynamical (cloud$-$scale eddies) structure of boundary layer clouds. Unraveling the 3D dynamical structure of stratocumulus and shallow cumulus clouds requires decomposition of the environmental wind contribution and particle sedimentation velocity from the observed radial Doppler velocity. Themore » second thrust proposes to unravel the mechanism of cumulus entrainment (location, scales) and its impact on microphysics utilizing radar measurements from the vertically pointing and new scanning radars at the ARM sites. The third research thrust requires the development of a cloud$-$tracking algorithm that monitors the properties of cloud.« less

  20. Use of the X-Band Radar to Support the Detection of In-Flight Icing Hazards by the NASA Icing Remote Sensing System

    NASA Technical Reports Server (NTRS)

    Serke, David J.; Politovich, Marcia K.; Reehorst, Andrew L.; Gaydos, Andrew

    2009-01-01

    The Alliance Icing Research Study-II (AIRS-II) field program was conducted near Montreal, Canada during the winter of 2003. The NASA Icing Remote Detection System (NIRSS) was deployed to detect in-flight icing hazards and consisted of a vertically pointing multichannel radiometer, a ceilometer and an x-band cloud radar. The radiometer was used to derive atmospheric temperature soundings and integrated liquid water, while the ceilometer and radar were used only to define cloud boundaries. The purpose of this study is to show that the radar reflectivity profiles from AIRS-II case studies could be used to provide a qualitative icing hazard.

  1. SWCNT-MoS2 -SWCNT Vertical Point Heterostructures.

    PubMed

    Zhang, Jin; Wei, Yang; Yao, Fengrui; Li, Dongqi; Ma, He; Lei, Peng; Fang, Hehai; Xiao, Xiaoyang; Lu, Zhixing; Yang, Juehan; Li, Jingbo; Jiao, Liying; Hu, Weida; Liu, Kaihui; Liu, Kai; Liu, Peng; Li, Qunqing; Lu, Wei; Fan, Shoushan; Jiang, Kaili

    2017-02-01

    A vertical point heterostructure (VPH) is constructed by sandwiching a two-dimensional (2D) MoS 2 flake with two cross-stacked metallic single-walled carbon nanotubes. It can be used as a field-effect transistor with high on/off ratio and a light detector with high spatial resolution. Moreover, the hybrid 1D-2D-1D VPHs open up new possibilities for nanoelectronics and nano-optoelectronics. © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

  2. Radar research on thunderstorms and lightning

    NASA Technical Reports Server (NTRS)

    Rust, W. D.; Doviak, R. J.

    1982-01-01

    Applications of Doppler radar to detection of storm hazards are reviewed. Normal radar sweeps reveal data on reflectivity fields of rain drops, ionized lightning paths, and irregularities in humidity and temperature. Doppler radar permits identification of the targets' speed toward or away from the transmitter through interpretation of the shifts in the microwave frequency. Wind velocity fields can be characterized in three dimensions by the use of two radar units, with a Nyquist limit on the highest wind speeds that may be recorded. Comparisons with models numerically derived from Doppler radar data show substantial agreement in storm formation predictions based on information gathered before the storm. Examples are provided of tornado observations with expanded Nyquist limits, gust fronts, turbulence, lightning and storm structures. Obtaining vertical velocities from reflectivity spectra is discussed.

  3. The ARM Cloud Radar Simulator for Global Climate Models: Bridging Field Data and Climate Models

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

    Zhang, Yuying; Xie, Shaocheng; Klein, Stephen A.

    Clouds play an important role in Earth’s radiation budget and hydrological cycle. However, current global climate models (GCMs) have had difficulties in accurately simulating clouds and precipitation. To improve the representation of clouds in climate models, it is crucial to identify where simulated clouds differ from real world observations of them. This can be difficult, since significant differences exist between how a climate model represents clouds and what instruments observe, both in terms of spatial scale and the properties of the hydrometeors which are either modeled or observed. To address these issues and minimize impacts of instrument limitations, the conceptmore » of instrument “simulators”, which convert model variables into pseudo-instrument observations, has evolved with the goal to improve and to facilitate the comparison of modeled clouds with observations. Many simulators have (and continue to be developed) for a variety of instruments and purposes. A community satellite simulator package, the Cloud Feedback Model Intercomparison Project (CFMIP) Observation Simulator Package (COSP; Bodas-Salcedo et al. 2011), contains several independent satellite simulators and is being widely used in the global climate modeling community to exploit satellite observations for model cloud evaluation (e.g., Klein et al. 2013; Zhang et al. 2010). This article introduces a ground-based cloud radar simulator developed by the U.S. Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) program for comparing climate model clouds with ARM observations from its vertically pointing 35-GHz radars. As compared to CloudSat radar observations, ARM radar measurements occur with higher temporal resolution and finer vertical resolution. This enables users to investigate more fully the detailed vertical structures within clouds, resolve thin clouds, and quantify the diurnal variability of clouds. Particularly, ARM radars are sensitive to low-level clouds

  4. Radar target classification studies: Software development and documentation

    NASA Astrophysics Data System (ADS)

    Kamis, A.; Garber, F.; Walton, E.

    1985-09-01

    Three computer programs were developed to process and analyze calibrated radar returns. The first program, called DATABASE, was developed to create and manage a random accessed data base. The second program, called FTRAN DB, was developed to process horizontal and vertical polarizations radar returns into different formats (i.e., time domain, circular polarizations and polarization parameters). The third program, called RSSE, was developed to simulate a variety of radar systems and to evaluate their ability to identify radar returns. Complete computer listings are included in the appendix volumes.

  5. Radar Observations of Convective Systems from a High-Altitude Aircraft

    NASA Technical Reports Server (NTRS)

    Heymsfield, G.; Geerts, B.; Tian, L.

    1999-01-01

    Reflectivity data collected by the precipitation radar on board the tropical Rainfall Measuring Mission (TRMM) satellite, orbiting at 350 km altitude, are compared to reflectivity data collected nearly simultaneously by a doppler radar aboard the NASA ER-2 flying at 19-20 km altitude, i.e. above even the deepest convection. The TRMM precipitation radar is a scanning device with a ground swath width of 215 km, and has a resolution of about a4.4 km in the horizontal and 250 m in the vertical (125 m in the core swath 48 km wide). The TRMM radar has a wavelength of 217 cm (13.8 GHz) and the Nadir mirror echo below the surface is used to correct reflectivity for loss by attenuation. The ER-2 Doppler radar (EDOP) has two antennas, one pointing to the nadir, 34 degrees forward. The forward pointing beam receives both the normal and the cross-polarized echos, so the linear polarization ratio field can be monitored. EDOP has a wavelength of 3.12 cm (9.6 GHz), a vertical resolution of 37.5 m and a horizontal along-track resolution of about 100 m. The 2-D along track airflow field can be synthesized from the radial velocities of both beams, if a reflectivity-based hydrometer fall speed relation can be assumed. It is primarily the superb vertical resolution that distinguishes EDOP from other ground-based or airborne radars. Two experiments were conducted during 1998 into validate TRMM reflectivity data over convection and convectively-generated stratiform precipitation regions. The Teflun-A (TEXAS-Florida Underflight) experiment, was conducted in April and May and focused on mesoscale convective systems mainly in southeast Texas. TEFLUN-B was conducted in August-September in central Florida, in coordination with CAMEX-3 (Convection and Moisture Experiment). The latter was focused on hurricanes, especially during landfall, whereas TEFLUN-B concentrated on central; Florida convection, which is largely driven and organized by surface heating and ensuing sea breeze circulations

  6. Radiosonde pressure sensor performance - Evaluation using tracking radars

    NASA Technical Reports Server (NTRS)

    Parsons, C. L.; Norcross, G. A.; Brooks, R. L.

    1984-01-01

    The standard balloon-borne radiosonde employed for synoptic meteorology provides vertical profiles of temperature, pressure, and humidity as a function of elapsed time. These parameters are used in the hypsometric equation to calculate the geopotential altitude at each sampling point during the balloon's flight. It is important that the vertical location information be accurate. The present investigation was conducted with the objective to evaluate the altitude determination accuracy of the standard radiosonde throughout the entire balloon profile. The tests included two other commercially available pressure sensors to see if they could provide improved accuracy in the stratosphere. The pressure-measuring performance of standard baroswitches, premium baroswitches, and hypsometers in balloon-borne sondes was correlated with tracking radars. It was found that the standard and premium baroswitches perform well up to about 25 km altitude, while hypsometers provide more reliable data above 25 km.

  7. Vertical velocity structure and geometry of clear air convective elements

    NASA Technical Reports Server (NTRS)

    Rowland, J. R.; Arnold, A.

    1975-01-01

    The paper discusses observations of individual convective elements with a high-power narrow-beam scanning radar, an FM-CW radar, and an acoustic sounder, including the determination of the vertical air velocity patterns of convective structures with the FM-CW radar and acoustic sounder. Data are presented which link the observed velocity structure and geometrical patterns to previously proposed models of boundary layer convection. It is shown that the high-power radar provides a clear three-dimensional picture of convective cells and fields over a large area with a resolution of 150 m, where the convective cells are roughly spherical. Analysis of time-height records of the FM-CW radar and acoustic sounder confirms the downdraft-entrainment mechanism of the convective cell. The Doppler return of the acoustic sounder and the insect-trail slopes on FM-CW radar records are independent but redundant methods for obtaining the vertical velocity patterns of convective structures.

  8. Simulation of a weather radar display for over-water airborne radar approaches

    NASA Technical Reports Server (NTRS)

    Clary, G. R.

    1983-01-01

    Airborne radar approach (ARA) concepts are being investigated as a part of NASA's Rotorcraft All-Weather Operations Research Program on advanced guidance and navigation methods. This research is being conducted using both piloted simulations and flight test evaluations. For the piloted simulations, a mathematical model of the airborne radar was developed for over-water ARAs to offshore platforms. This simulated flight scenario requires radar simulation of point targets, such as oil rigs and ships, distributed sea clutter, and transponder beacon replies. Radar theory, weather radar characteristics, and empirical data derived from in-flight radar photographs are combined to model a civil weather/mapping radar typical of those used in offshore rotorcraft operations. The resulting radar simulation is realistic and provides the needed simulation capability for ongoing ARA research.

  9. Differential absorption radar techniques: water vapor retrievals

    NASA Astrophysics Data System (ADS)

    Millán, Luis; Lebsock, Matthew; Livesey, Nathaniel; Tanelli, Simone

    2016-06-01

    Two radar pulses sent at different frequencies near the 183 GHz water vapor line can be used to determine total column water vapor and water vapor profiles (within clouds or precipitation) exploiting the differential absorption on and off the line. We assess these water vapor measurements by applying a radar instrument simulator to CloudSat pixels and then running end-to-end retrieval simulations. These end-to-end retrievals enable us to fully characterize not only the expected precision but also their potential biases, allowing us to select radar tones that maximize the water vapor signal minimizing potential errors due to spectral variations in the target extinction properties. A hypothetical CloudSat-like instrument with 500 m by ˜ 1 km vertical and horizontal resolution and a minimum detectable signal and radar precision of -30 and 0.16 dBZ, respectively, can estimate total column water vapor with an expected precision of around 0.03 cm, with potential biases smaller than 0.26 cm most of the time, even under rainy conditions. The expected precision for water vapor profiles was found to be around 89 % on average, with potential biases smaller than 77 % most of the time when the profile is being retrieved close to surface but smaller than 38 % above 3 km. By using either horizontal or vertical averaging, the precision will improve vastly, with the measurements still retaining a considerably high vertical and/or horizontal resolution.

  10. Parametric dependence of ocean wave-radar modulation transfer functions

    NASA Technical Reports Server (NTRS)

    Plant, W. J.; Keller, W. C.; Cross, A.

    1983-01-01

    Microwave techniques at X and L band were used to determine the dependence of ocean-wave radar modulation transfer functions (MTFs) on various environmental and radar parameters during the Marine Remote Sensing experiment of 1979 (MARSEN 79). These MIF are presented, as are coherence functions between the AM and FM parts of the backscattered microwave signal. It is shown that they both depend on several of these parameters. Besides confirming many of the properties of transfer functions reported by previous authors, indications are found that MTFs decrease with increasing angle between wave propagation and antenna-look directions but are essentially independent of small changes in air-sea temperature difference. However, coherence functions are much smaller when the antennas are pointed perpendicular to long waves. It is found that X band transfer functions measured with horizontally polarized microwave radiation have larger magnitudes than those obtained by using vertical polarization.

  11. Space Radar Image of Belgrade, Serbia

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This spaceborne radar image of Belgrade, Serbia, illustrates the variety of land use patterns that can be observed with a multiple wavelength radar system. Belgrade, the capital of Serbia and former capital of Yugoslavia, is the bright area in the center of the image. The Danube River flows from the top to the bottom of the image, and the Sava River flows into the Danube from the left. Agricultural fields appear in shades of dark blue, purple and brown in outlying areas. Vegetated areas along the rivers appear in light blue-green, while dense forests in hillier areas in the lower left appear in a darker shade of green. The image was acquired by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 2, 1994. The image is centered at 44.5 degrees north latitude and 20.5 degrees east longitude. North is toward the upper right. The image shows an area 36 kilometers by 32 kilometers 22 miles by 20 miles). The colors are assigned to different frequencies and polarizations of the radar as follows: red is L-band, horizontally transmitted, horizontally received; green is L-band, horizontally transmitted, vertically received; blue is C-band, horizontally transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.

  12. A prototype fully polarimetric 160-GHz bistatic ISAR compact radar range

    NASA Astrophysics Data System (ADS)

    Beaudoin, C. J.; Horgan, T.; DeMartinis, G.; Coulombe, M. J.; Goyette, T.; Gatesman, A. J.; Nixon, William E.

    2017-05-01

    We present a prototype bistatic compact radar range operating at 160 GHz and capable of collecting fullypolarimetric radar cross-section and electromagnetic scattering measurements in a true far-field facility. The bistatic ISAR system incorporates two 90-inch focal length, 27-inch-diameter diamond-turned mirrors fed by 160 GHz transmit and receive horns to establish the compact range. The prototype radar range with its modest sized quiet zone serves as a precursor to a fully developed compact radar range incorporating a larger quiet zone capable of collecting X-band bistatic RCS data and 3D imagery using 1/16th scale objects. The millimeter-wave transmitter provides 20 GHz of swept bandwidth in the single linear (Horizontal/Vertical) polarization while the millimeter-wave receiver, that is sensitive to linear Horizontal and Vertical polarization, possesses a 7 dB noise figure. We present the design of the compact radar range and report on test results collected to validate the system's performance.

  13. Auroral-Region Dynamics Determined with the Chatanika Radar.

    DTIC Science & Technology

    1982-11-01

    report) 17. DISTRIBUTION STATEMENT (of the abstract entered In Block 20, If different from report) 18 . SUPPLEMENTARY NOTES 19. KEY WORDS (Continue on...for 1 April 1973 .......... ... 41 18 Vertical Neutral Wind Measured with the Fabry-Perot Interferometer ......... ........................ ... 44 vii...Waves Determined from Radar Observations on 18 January 1976 ..... ............... ... 50 23 Meridional Wind and Gravity Waves Determined from Radar

  14. Estimating Forest Vertical Structure from Multialtitude, Fixed-Baseline Radar Interferometric and Polarimetric Data

    NASA Technical Reports Server (NTRS)

    Treuhaft, Robert N.; Law, Beverly E.; Siqueira, Paul R.

    2000-01-01

    Parameters describing the vertical structure of forests, for example tree height, height-to-base-of-live-crown, underlying topography, and leaf area density, bear on land-surface, biogeochemical, and climate modeling efforts. Single, fixed-baseline interferometric synthetic aperture radar (INSAR) normalized cross-correlations constitute two observations from which to estimate forest vertical structure parameters: Cross-correlation amplitude and phase. Multialtitude INSAR observations increase the effective number of baselines potentially enabling the estimation of a larger set of vertical-structure parameters. Polarimetry and polarimetric interferometry can further extend the observation set. This paper describes the first acquisition of multialtitude INSAR for the purpose of estimating the parameters describing a vegetated land surface. These data were collected over ponderosa pine in central Oregon near longitude and latitude -121 37 25 and 44 29 56. The JPL interferometric TOPSAR system was flown at the standard 8-km altitude, and also at 4-km and 2-km altitudes, in a race track. A reference line including the above coordinates was maintained at 35 deg for both the north-east heading and the return southwest heading, at all altitudes. In addition to the three altitudes for interferometry, one line was flown with full zero-baseline polarimetry at the 8-km altitude. A preliminary analysis of part of the data collected suggests that they are consistent with one of two physical models describing the vegetation: 1) a single-layer, randomly oriented forest volume with a very strong ground return or 2) a multilayered randomly oriented volume; a homogeneous, single-layer model with no ground return cannot account for the multialtitude correlation amplitudes. Below the inconsistency of the data with a single-layer model is followed by analysis scenarios which include either the ground or a layered structure. The ground returns suggested by this preliminary analysis seem

  15. Space Radar Image of Central Sumatra, Indonesia

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This is a radar image of the central part of the island of Sumatra in Indonesia that shows how the tropical rainforest typical of this country is being impacted by human activity. Native forest appears in green in this image, while prominent pink areas represent places where the native forest has been cleared. The large rectangular areas have been cleared for palm oil plantations. The bright pink zones are areas that have been cleared since 1989, while the dark pink zones are areas that were cleared before 1989. These radar data were processed as part of an effort to assist oil and gas companies working in the area to assess the environmental impact of both their drilling operations and the activities of the local population. Radar images are useful in these areas because heavy cloud cover and the persistent smoke and haze associated with deforestation have prevented usable visible-light imagery from being acquired since 1989. The dark shapes in the upper right (northeast) corner of the image are a chain of lakes in flat coastal marshes. This image was acquired in October 1994 by the Spaceborne Imaging Radar C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour. Environmental changes can be easily documented by comparing this image with visible-light data that were acquired in previous years by the Landsat satellite. The image is centered at 0.9 degrees north latitude and 101.3 degrees east longitude. The area shown is 50 kilometers by 100 kilometers (31 miles by 62 miles). The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, vertically received; blue is L-band vertically transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.

  16. Space Radar Image of Central Sumatra, Indonesia

    NASA Image and Video Library

    1999-04-15

    This is a radar image of the central part of the island of Sumatra in Indonesia that shows how the tropical rainforest typical of this country is being impacted by human activity. Native forest appears in green in this image, while prominent pink areas represent places where the native forest has been cleared. The large rectangular areas have been cleared for palm oil plantations. The bright pink zones are areas that have been cleared since 1989, while the dark pink zones are areas that were cleared before 1989. These radar data were processed as part of an effort to assist oil and gas companies working in the area to assess the environmental impact of both their drilling operations and the activities of the local population. Radar images are useful in these areas because heavy cloud cover and the persistent smoke and haze associated with deforestation have prevented usable visible-light imagery from being acquired since 1989. The dark shapes in the upper right (northeast) corner of the image are a chain of lakes in flat coastal marshes. This image was acquired in October 1994 by the Spaceborne Imaging Radar C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour. Environmental changes can be easily documented by comparing this image with visible-light data that were acquired in previous years by the Landsat satellite. The image is centered at 0.9 degrees north latitude and 101.3 degrees east longitude. The area shown is 50 kilometers by 100 kilometers (31 miles by 62 miles). The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, vertically received; blue is L-band vertically transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program. http

  17. Space Radar Image of Boston, Massachusetts

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This radar image of the area surrounding Boston, Mass., shows how a spaceborne radar system distinguishes between densely populated urban areas and nearby areas that are relatively unsettled. The bright white area at the right center of the image is downtown Boston. The wide river below and to the left of the city is the Charles River in Boston's Back Bay neighborhood. The dark green patch to the right of the Back Bay is Boston Common. A bridge across the north end of Back Bay connects the cities of Boston and Cambridge. The light green areas that dominate most of the image are the suburban communities surrounding Boston. The many ponds that dot the region appear as dark irregular spots. Many densely populated urban areas show up as red in the image due to the alignment of streets and buildings to the incoming radar beam. North is toward the upper left. The image was acquired on October 9, 1994, by the Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) as it flew aboard the space shuttle Endeavour. This area is centered at 42.4 degrees north latitude, 71.2 degrees west longitude. The area shown is approximately 37 km by 18 km (23 miles by 11 miles). Colors are assigned to different radar frequencies and polarizations as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, vertically received; blue is C-band horizontally transmitted, vertically received. SIR-C/X-SAR, a cooperative mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.

  18. Space Radar Image of Sydney, Australia

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This spaceborne radar image is dominated by the metropolitan area of Australia's largest city, Sydney. Sydney Harbour, with numerous coves and inlets, is seen in the upper center of the image, and the roughly circular Botany Bay is shown in the lower right. The downtown business district of Sydney appears as a bright white area just above the center of the image. The Sydney Harbour Bridge is a white line adjacent to the downtown district. The well-known Sydney Opera House is the small, white dot to the right of the bridge. Urban areas appear yellow, blue and brown. The purple areas are undeveloped areas and park lands. Manly, the famous surfing beach, is shown in yellow at the top center of the image. Runways from the Sydney Airport are the dark features that extend into Botany Bay in the lower right. Botany Bay is the site where Captain James Cook first landed his ship, Endeavour, in 1770. The image was acquired by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) on April 20, 1994, onboard the space shuttle Endeavour. The area shown is 33 kilometers by 38kilometers (20 miles by 23 miles) and is centered at 33.9 degrees south latitude, 151.2 degrees east longitude. North is toward the upper left. The colors are assigned to different radar frequenciesand polarizations as follows: red is L-band, vertically transmittedand horizontally received; green is C-band, vertically transmitted and horizontally received; and blue is C-band, vertically transmittedand received. SIR-C/X-SAR, a joint mission of the German, Italianand United States space agencies, is part of NASA's Mission to Planet Earth. #####

  19. Dual-Polarization Observations of Slowly Varying Solar Emissions from a Mobile X-Band Radar

    PubMed Central

    Gabella, Marco; Leuenberger, Andreas

    2017-01-01

    The radio noise that comes from the Sun has been reported in literature as a reference signal to check the quality of dual-polarization weather radar receivers for the S-band and C-band. In most cases, the focus was on relative calibration: horizontal and vertical polarizations were evaluated versus the reference signal mainly in terms of standard deviation of the difference. This means that the investigated radar receivers were able to reproduce the slowly varying component of the microwave signal emitted by the Sun. A novel method, aimed at the absolute calibration of dual-polarization receivers, has recently been presented and applied for the C-band. This method requires the antenna beam axis to be pointed towards the center of the Sun for less than a minute. Standard deviations of the difference as low as 0.1 dB have been found for the Swiss radars. As far as the absolute calibration is concerned, the average differences were of the order of −0.6 dB (after noise subtraction). The method has been implemented on a mobile, X-band radar, and this paper presents the successful results that were obtained during the 2016 field campaign in Payerne (Switzerland). Despite a relatively poor Sun-to-Noise ratio, the “small” (~0.4 dB) amplitude of the slowly varying emission was captured and reproduced; the standard deviation of the difference between the radar and the reference was ~0.2 dB. The absolute calibration of the vertical and horizontal receivers was satisfactory. After the noise subtraction and atmospheric correction a, the mean difference was close to 0 dB. PMID:28531164

  20. Dual-Polarization Observations of Slowly Varying Solar Emissions from a Mobile X-Band Radar.

    PubMed

    Gabella, Marco; Leuenberger, Andreas

    2017-05-22

    The radio noise that comes from the Sun has been reported in literature as a reference signal to check the quality of dual-polarization weather radar receivers for the S-band and C-band. In most cases, the focus was on relative calibration: horizontal and vertical polarizations were evaluated versus the reference signal mainly in terms of standard deviation of the difference. This means that the investigated radar receivers were able to reproduce the slowly varying component of the microwave signal emitted by the Sun. A novel method, aimed at the absolute calibration of dual-polarization receivers, has recently been presented and applied for the C-band. This method requires the antenna beam axis to be pointed towards the center of the Sun for less than a minute. Standard deviations of the difference as low as 0.1 dB have been found for the Swiss radars. As far as the absolute calibration is concerned, the average differences were of the order of -0.6 dB (after noise subtraction). The method has been implemented on a mobile, X-band radar, and this paper presents the successful results that were obtained during the 2016 field campaign in Payerne (Switzerland). Despite a relatively poor Sun-to-Noise ratio, the "small" (~0.4 dB) amplitude of the slowly varying emission was captured and reproduced; the standard deviation of the difference between the radar and the reference was ~0.2 dB. The absolute calibration of the vertical and horizontal receivers was satisfactory. After the noise subtraction and atmospheric correction a, the mean difference was close to 0 dB.

  1. Borehole radar interferometry revisited

    USGS Publications Warehouse

    Liu, Lanbo; Ma, Chunguang; Lane, John W.; Joesten, Peter K.

    2014-01-01

    Single-hole, multi-offset borehole-radar reflection (SHMOR) is an effective technique for fracture detection. However, commercial radar system limitations hinder the acquisition of multi-offset reflection data in a single borehole. Transforming cross-hole transmission mode radar data to virtual single-hole, multi-offset reflection data using a wave interferometric virtual source (WIVS) approach has been proposed but not fully demonstrated. In this study, we compare WIVS-derived virtual single-hole, multi-offset reflection data to real SHMOR radar reflection profiles using cross-hole and single-hole radar data acquired in two boreholes located at the University of Connecticut (Storrs, CT USA). The field data results are similar to full-waveform numerical simulations developed for a two-borehole model. The reflection from the adjacent borehole is clearly imaged by both the real and WIVS-derived virtual reflection profiles. Reflector travel-time changes induced by deviation of the two boreholes from the vertical can also be observed on the real and virtual reflection profiles. The results of this study demonstrate the potential of the WIVS approach to improve bedrock fracture imaging for hydrogeological and petroleum reservoir development applications.

  2. Retrieval of Raindrop Size Distribution, Vertical Air Velocity and Water Vapor Attenuation Using Dual-Wavelength Doppler Radar Observations

    NASA Technical Reports Server (NTRS)

    Heymsfield, Gerald M.; Tian, Lin; Li, Lihua; Srivastava, C.

    2005-01-01

    Two techniques for retrieving the slope and intercept parameters of an assumed exponential raindrop size distribution (RSD), vertical air velocity, and attenuation by precipitation and water vapor in light stratiform rain using observations by airborne, nadir looking dual-wavelength (X-band, 3.2 cm and W-band, 3.2 mm) radars are presented. In both techniques, the slope parameter of the RSD and the vertical air velocity are retrieved using only the mean Doppler velocities at the two wavelengths. In the first method, the intercept of the RSD is estimated from the observed reflectivity at the longer wavelength assuming no attenuation at that wavelength. The attenuation of the shorter wavelength radiation by precipitation and water vapor are retrieved using the observed reflectivity at the shorter wavelength. In the second technique, it is assumed that the longer wavelength suffers attenuation only in the melting band. Then, assuming a distribution of water vapor, the melting band attenuation at both wavelengths and the rain attenuation at the shorter wavelength are retrieved. Results of the retrievals are discussed and several physically meaningful results are presented.

  3. Effectiveness of glues for harmonic radar tag attachment on Halyomorpha halys (Hemiptera: Pentatomidae) and their impact on adult survivorship and mobility.

    PubMed

    Lee, Doo-Hyung; Wright, Starker E; Boiteau, Gilles; Vincent, Charles; Leskey, Tracy C

    2013-06-01

    We evaluated the effectiveness of three cyanoacrylate glues (trade names: Krazy [Elmer's Products Inc., Westerville, OH], Loctite [Henkel Corporation, Rocky Hill, CT], and FSA [Barnes Distribution, Cleveland, OH]) to attach harmonic radar tags securely on adult Halyomorpha halys (Stål) (Hemiptera: Pentatomidae) and quantified the effect of the radar tag attachment on insect survivorship and mobility. In the laboratory, the strength of the glue bond between the radar tag and H. halys pronotum was significantly increased when the pronotum was sanded to remove cuticular waxes. The adhesive bond of the radar tag to the sanded pronotum of H. halys had strength of 160-190-g force and there was no significant difference among the three types of glue tested. The three glues had no measurable effect on the survivorship of radar-tagged H. halys over 7 d, compared with untagged insects. Over a 7-d period in the laboratory, horizontal distance traveled, horizontal walking velocity, and vertical climbing distance were all unaffected by the presence of the tags regardless of glue. A field experiment was conducted to compare the free flight behavior of untagged and radar-tagged H. halys. Adults were released on a vertical dowel and their flights were tracked visually up to ≍200 m from the release point. There was no significant difference in take-off time or in flight distance, time, or speed between untagged and radar-tagged individuals. In addition, prevailing flight direction was not significantly different between untagged and radar-tagged individuals. The absence of measurable impact of the radar tag attachment on H. halys survivorship or mobility validates the use of harmonic radar tags to study the dispersal ecology of this insect in field conditions.

  4. Space Radar Image of Victoria, Canada

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This three-frequency spaceborne radar image shows the southern end of Vancouver Island on the west coast of Canada. The white area in the lower right is the city of Victoria, the capital of the province of British Columbia. The three radar frequencies help to distinguish different land use patterns. The bright pink areas are suburban regions, the brownish areas are forested regions, and blue areas are agricultural fields or forest clear-cuts. Founded in 1843 as a fur trading post, Victoria has grown to become one of western Canada's largest commercial centers. In the upper right is San Juan Island, in the state of Washington. The Canada/U.S. border runs through Haro Strait, on the right side of the image, between San Juan Island and Vancouver Island. The image was acquired by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) on October 6, 1994, onboard the space shuttle Endeavour. The area shown is 37 kilometers by 42 kilometers (23 miles by 26 miles) and is centered at 48.5 degrees north latitude, 123.3 degrees west longitude. North is toward the upper left. The colors are assigned to different radar frequencies and polarizations as follows: red is L-band horizontally transmitted and received; green is C-band, vertically transmitted and received; and blue is X-band, vertically transmitted and received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.

  5. Observation of snowfall with a low-power FM-CW K-band radar (Micro Rain Radar)

    NASA Astrophysics Data System (ADS)

    Kneifel, Stefan; Maahn, Maximilian; Peters, Gerhard; Simmer, Clemens

    2011-06-01

    Quantifying snowfall intensity especially under arctic conditions is a challenge because wind and snow drift deteriorate estimates obtained from both ground-based gauges and disdrometers. Ground-based remote sensing with active instruments might be a solution because they can measure well above drifting snow and do not suffer from flow distortions by the instrument. Clear disadvantages are, however, the dependency of e.g. radar returns on snow habit which might lead to similar large uncertainties. Moreover, high sensitivity radars are still far too costly to operate in a network and under harsh conditions. In this paper we compare returns from a low-cost, low-power vertically pointing FM-CW radar (Micro Rain Radar, MRR) operating at 24.1 GHz with returns from a 35.5 GHz cloud radar (MIRA36) for dry snowfall during a 6-month observation period at an Alpine station (Environmental Research Station Schneefernerhaus, UFS) at 2,650 m height above sea level. The goal was to quantify the potential and limitations of the MRR in relation to what is achievable by a cloud radar. The operational MRR procedures to derive standard radar variables like effective reflectivity factor ( Z e) or the mean Doppler velocity ( W) had to be modified for snowfall since the MRR was originally designed for rain observations. Since the radar returns from snowfall are weaker than from comparable rainfall, the behavior of the MRR close to its detection threshold has been analyzed and a method is proposed to quantify the noise level of the MRR based on clear sky observations. By converting the resulting MRR- Z e into 35.5 GHz equivalent Z e values, a remaining difference below 1 dBz with slightly higher values close to the noise threshold could be obtained. Due to the much higher sensitivity of MIRA36, the transition of the MRR from the true signal to noise can be observed, which agrees well with the independent clear sky noise estimate. The mean Doppler velocity differences between both radars

  6. Comparing helicopter-borne profiling radar with airborne laser scanner data for forest structure estimation.

    NASA Astrophysics Data System (ADS)

    Piermattei, Livia; Hollaus, Markus; Pfeifer, Norbert; Chen, Yuwei; Karjalainen, Mika; Hakala, Teemu; Hyyppä, Juha; Wagner, Wolfgang

    2017-04-01

    Forests are complex ecosystems that show substantial variation with respect to climate, management regime, stand history, disturbance, and needs of local communities. The dynamic processes of growth and disturbance are reflected in the structural components of forests that include the canopy vertical structure and geometry (e.g. size, height, and form), tree position and species diversity. Current remote-sensing systems to measure forest structural attributes include passive optical sensors and active sensors. The technological capabilities of active remote sensing like the ability to penetrate the vegetation and provide information about its vertical structure has promoted an extensive use of LiDAR (Light Detection And Ranging) and radar (RAdio Detection And Ranging) system over the last 20 years. LiDAR measurements from aircraft (airborne laser scanning, ALS) currently represents the primary data source for three-dimensional information on forest vertical structure. Contrary, despite the potential of radar remote sensing, their use is not yet established in forest monitoring. In order to better understand the interaction of pulsed radar with the forest canopy, and to increase the feasibility of this system, the Finnish Geospatial Research Institute has developed a helicopter-borne profiling radar system, called TomoRadar. TomoRadar is capable of recording a canopy-penetrating profile of forests. To georeference the radar measurements the system was equipped with a global navigation satellite system and an inertial measurement unit with a centimeter level accuracy of the flight trajectory. The TomoRadar operates at Ku-band, (wave lengths λ 1.5cm) with two separated parabolic antennas providing co- and cross-polarization modes. The purpose of this work is to investigate the capability of the TomoRadar system, for estimating the forest vertical profile, terrain topography and tree height. We analysed 600 m TomoRadar crosspolarized (i.e. horizontal - vertical

  7. A new system model for radar polarimeters

    NASA Technical Reports Server (NTRS)

    Freeman, Anthony

    1991-01-01

    The validity of the 2 x 2 receive R and transmit T model for radar polarimeter systems, first proposed by Zebker et al. (1987), is questioned. The model is found to be invalid for many practical realizations of radar polarimeters, which can lead to significant errors in the calibration of polarimetric radar images. A more general model is put forward, which addresses the system defects which cause the 2 x 2 model to break down. By measuring one simple parameter from a polarimetric active radar calibration (PARC), it is possible to transform the scattering matrix measurements made by a radar polarimeter to a format compatible with a 2 x 2 R and T matrix model. Alternatively, the PARC can be used to verify the validity of the 2 x 2 model for any polarimetric radar system. Recommendations for the use of PARCs in polarimetric calibration and to measure the orientation angle of the horizontal (H) and vertical (V) coordinate system are also presented.

  8. A new system model for radar polarimeters

    NASA Astrophysics Data System (ADS)

    Freeman, Anthony

    1991-09-01

    The validity of the 2 x 2 receive R and transmit T model for radar polarimeter systems, first proposed by Zebker et al. (1987), is questioned. The model is found to be invalid for many practical realizations of radar polarimeters, which can lead to significant errors in the calibration of polarimetric radar images. A more general model is put forward, which addresses the system defects which cause the 2 x 2 model to break down. By measuring one simple parameter from a polarimetric active radar calibration (PARC), it is possible to transform the scattering matrix measurements made by a radar polarimeter to a format compatible with a 2 x 2 R and T matrix model. Alternatively, the PARC can be used to verify the validity of the 2 x 2 model for any polarimetric radar system. Recommendations for the use of PARCs in polarimetric calibration and to measure the orientation angle of the horizontal (H) and vertical (V) coordinate system are also presented.

  9. Observations of the structure and vertical transport of the polar upper ionosphere with the EISCAT VHF radar. I - Is EISCAT able to determine O(+) and H(+) polar wind characteristic? A simulation study

    NASA Technical Reports Server (NTRS)

    Blelly, Pierre-Louis; Barakat, Abdullah R.; Fontanari, Jean; Alcayde, Denis; Blanc, Michel; Wu, Jian; Lathuillere, C.

    1992-01-01

    A method presented by Wu et al. (1992) for computing the H(+) vertical velocity from the main ionospheric parameters measured by the EISCAT VHF radar is tested in a fully controlled sequence which consists of generating an ideal ionospheric model by solving the coupled continuity and momentum equations for a two-ion plasma (O(+) and H(+)). Synthetic autocorrelation functions are generated from this model with the radar characteristics and used as actual measurements to compute the H(+) vertical velocities. Results of these simulations are shown and discussed for three cases of typical and low SNR and for low and increased mixing ratios. In most cases general agreement is found between computed H(+) velocities and generic ones with the altitude range considered, i.e., 200-1000 km. The method is shown to be reliable.

  10. Space Radar Image of Long Valley, California - 3-D view

    NASA Image and Video Library

    1999-05-01

    This is a three-dimensional perspective view of Long Valley, California by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar on board the space shuttle Endeavour. This view was constructed by overlaying a color composite SIR-C image on a digital elevation map. The digital elevation map was produced using radar interferometry, a process by which radar data are acquired on different passes of the space shuttle and, which then, are compared to obtain elevation information. The data were acquired on April 13, 1994 and on October 3, 1994, during the first and second flights of the SIR-C/X-SAR radar instrument. The color composite radar image was produced by assigning red to the C-band (horizontally transmitted and vertically received) polarization; green to the C-band (vertically transmitted and received) polarization; and blue to the ratio of the two data sets. Blue areas in the image are smooth and yellow areas are rock outcrops with varying amounts of snow and vegetation. The view is looking north along the northeastern edge of the Long Valley caldera, a volcanic collapse feature created 750,000 years ago and the site of continued subsurface activity. Crowley Lake is off the image to the left. http://photojournal.jpl.nasa.gov/catalog/PIA01757

  11. Space Radar Image of Samara, Russia

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This three-frequency space radar image shows the city of Samara, Russia in pink and light green right of center. Samara is at the junction of the Volga and Samara Rivers approximately 800 kilometers (500 miles) southeast of Moscow. The wide river in the center of the image is the Volga. Samara, formerly Kuybyshev, is a busy industrial city known for its chemical, mechanical and petroleum industries. Northwest of the Volga (upper left corner of the image) are deciduous forests of the Samarskaya Luka National Park. Complex patterns in the floodplain of the Volga are caused by 'cut-off' lakes and channels from former courses of the meandering river. The three radar frequencies allow scientists to distinguish different types of agricultural fields in the lower right side of the image. For example, fields which appear light blue are short grass or cleared fields. Purple and green fields contain taller plants or rough plowed soil. Scientists hope to use radar data such as these to understand the environmental consequences of industrial, agricultural and natural preserve areas coexisting in close proximity. This image is 50 kilometers by 26 kilometers (31 by 16 miles) and is centered at 53.2 degrees north latitude, 50.1 degrees east longitude. North is toward the top of the image. The colors are assigned to different radar frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is C-band, horizontally transmitted and vertically received; and blue is X-band, vertically transmitted and received. The image was acquired by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) on October 1, 1994 onboard the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth.

  12. a Study of Precipitation Using Dual-Frequency and Interferometric Doppler Radars.

    NASA Astrophysics Data System (ADS)

    Chilson, Phillip Bruce

    The primary focus of this dissertation involves the investigation of precipitation using Doppler radar but using distinctly different methods. Each method will be treated separately. The first part describes an investigation of a tropical thunderstorm that occurred in the summer of 1991 over the National Astronomy and Ionosphere Center in Arecibo, Puerto Rico. Observations were made using a vertically pointing, dual-wavelength, collinear beam Doppler radar which permits virtually simultaneous observations of the same pulse volume using transmission and reception of coherent UHF and VHF signals on alternate pulses. This made it possible to measure directly the vertical wind within the sampling volume using the VHF signal while using the UHF signal to study the nature of the precipitation. The observed storm showed strong similarities with systems observed in the Global Atmospheric Research Program's (GARP) Atlantic Tropical Experiment (GATE) study. The experiment provided a means of determining various parameters associated with the storm, such as the vertical air velocity, the mean fall speeds of the precipitation, and the reflectivity. Rogers proposed a means of deducing the mean fall speed of precipitation particles using the radar reflectivity factor. Using the data from our experiment, the mean precipitation fall speeds were calculated and compared with those that would be inferred from Rogers' method. The results suggest the Rogers method of estimating mean precipitation fall speeds to be unreliable in turbulent environments. The second part reports observations made with the 50 MHz Middle and Upper Atmosphere (MU) radar located at Shigaraki, Japan during May of 1992. The facility was operated in a spatial interferometry (SI) mode while observing frontal precipitation. The data suggest that the presence of precipitation can produce a bias in the SI cross-spectral phase that in turn creates an overestimation of the horizontal wind. The process is likened to

  13. Comparison of vertical hydraulic conductivity in a streambed-point bar system of a gaining stream

    NASA Astrophysics Data System (ADS)

    Dong, Weihong; Chen, Xunhong; Wang, Zhaowei; Ou, Gengxin; Liu, Can

    2012-07-01

    SummaryVertical hydraulic conductivities (Kv) of both streambed and point bars can influence water and solute exchange between streams and surrounding groundwater systems. The sediments in point bars are relatively young compared to the older sediments in the adjacent aquifers but slightly older compared to submerged streambeds. Thus, the permeability in point bar sediments can be different not only from regional aquifer but also from modern streambed. However, there is a lack of detailed studies that document spatial variability of vertical hydraulic conductivity in point bars of meandering streams. In this study, the authors proposed an in situ permeameter test method to measure vertical hydraulic conductivity of the two point bars in Clear Creek, Nebraska, USA. We compared the Kv values in streambed and adjacent point bars through 45 test locations in the two point bars and 51 test locations in the streambed. The Kv values in the point bars were lower than those in the streambed. Kruskal-Wallis test confirmed that the Kv values from the point bars and from the channel came from two statistically different populations. Within a point bar, the Kv values were higher along the point bar edges than those from inner point bars. Grain size analysis indicated that slightly more silt and clay particles existed in sediments from inner point bars, compared to that from streambed and from locations near the point bar edges. While point bars are the deposits of the adjacent channel, the comparison of two groups of Kv values suggests that post-depositional processes had an effect on the evolution of Kv from channel to point bars in fluvial deposits. We believed that the transport of fine particles and the gas ebullition in this gaining stream had significant effects on the distribution of Kv values in a streambed-point bar system. With the ageing of deposition in a floodplain, the permeability of point bar sediments can likely decrease due to reduced effects of the upward

  14. Space Radar Image of Ubar Optical/Radar

    NASA Image and Video Library

    1998-04-28

    This pair of images from space shows a portion of the southern Empty Quarter of the Arabian Peninsula in the country of Oman. On the left is a radar image of the region around the site of the fabled Lost City of Ubar, discovered in 1992 with the aid of remote sensing data. On the right is an enhanced optical image taken by the shuttle astronauts. Ubar existed from about 2800 BC to about 300 AD. and was a remote desert outpost where caravans were assembled for the transport of frankincense across the desert. The actual site of the fortress of the Lost City of Ubar, currently under excavation, is too small to show in either image. However, tracks leading to the site, and surrounding tracks, show as prominent, but diffuse, reddish streaks in the radar image. Although used in modern times, field investigations show many of these tracks were in use in ancient times as well. Mapping of these tracks on regional remote sensing images provided by the Landsat satellite was a key to recognizing the site as Ubar. The prominent magenta colored area is a region of large sand dunes. The green areas are limestone rocks, which form a rocky desert floor. A major wadi, or dry stream bed, runs across the scene and appears as a white line. The radar images, and ongoing field investigations, will help shed light on an early civilization about which little in known. The radar image was taken by the Spaceborne Imaging Radar C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) and is centered at 18 degrees North latitude and 53 degrees East longitude. The image covers an area about 50 kilometers by 100 kilometers (31 miles by 62 miles). The colors in the image are assigned to different frequencies and polarizations of the radar as follows: red is L-band, horizontally transmitted, horizontally received; blue is C-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian and the United

  15. Space Radar Image of Wenatchee, Washington

    NASA Technical Reports Server (NTRS)

    1994-01-01

    This spaceborne radar image shows a segment of the Columbia River as it passes through the area of Wenatchee, Washington, about 220 kilometers (136 miles) east of Seattle. The Wenatchee Mountains, part of the Cascade Range, are shown in green at the lower left of the image. The Cascades create a 'rain shadow' for the region, limiting rainfall east of the range to less than 26 centimeters (10 inches) per year. The radar's ability to see different types of vegetation is highlighted in the contrast between the pine forests, that appear in green and the dry valley plain that shows up as dark purple. The cities of Wenatchee and East Wenatchee are the grid-like areas straddling the Columbia River in the left center of the image. With a population of about 60,000, the region produces about half of Washington state's lucrative apple crop. Several orchard areas appear as green rectangular patches to the right of the river in the lower right center. Radar images such as these can be used to monitor land use patterns in areas such as Wenatchee, that have diverse and rapidly changing urban, agricultural and wild land pressures. This image was acquired by Spaceborne Imaging Radar-C/X-Band Synthetic Aperture Radar (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 10, 1994. The image is 38 kilometers by 45 kilometers (24 miles by 30 miles) and is centered at 47.3 degrees North latitude, 120.1 degrees West longitude. North is toward the upper left. The colors are assigned to different radar frequencies and polarizations of the radar as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted, vertically received; and blue is C-band, horizontally transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian, and United States space agencies, is part of NASA's Mission to Planet Earth.

  16. Optimal integration of gravity in trajectory planning of vertical pointing movements.

    PubMed

    Crevecoeur, Frédéric; Thonnard, Jean-Louis; Lefèvre, Philippe

    2009-08-01

    The planning and control of motor actions requires knowledge of the dynamics of the controlled limb to generate the appropriate muscular commands and achieve the desired goal. Such planning and control imply that the CNS must be able to deal with forces and constraints acting on the limb, such as the omnipresent force of gravity. The present study investigates the effect of hypergravity induced by parabolic flights on the trajectory of vertical pointing movements to test the hypothesis that motor commands are optimized with respect to the effect of gravity on the limb. Subjects performed vertical pointing movements in normal gravity and hypergravity. We use a model based on optimal control to identify the role played by gravity in the optimal arm trajectory with minimal motor costs. First, the simulations in normal gravity reproduce the asymmetry in the velocity profiles (the velocity reaches its maximum before half of the movement duration), which typically characterizes the vertical pointing movements performed on Earth, whereas the horizontal movements present symmetrical velocity profiles. Second, according to the simulations, the optimal trajectory in hypergravity should present an increase in the peak acceleration and peak velocity despite the increase in the arm weight. In agreement with these predictions, the subjects performed faster movements in hypergravity with significant increases in the peak acceleration and peak velocity, which were accompanied by a significant decrease in the movement duration. This suggests that movement kinematics change in response to an increase in gravity, which is consistent with the hypothesis that motor commands are optimized and the action of gravity on the limb is taken into account. The results provide evidence for an internal representation of gravity in the central planning process and further suggest that an adaptation to altered dynamics can be understood as a reoptimization process.

  17. Space Radar Image of Craters of the Moon, Idaho

    NASA Technical Reports Server (NTRS)

    1994-01-01

    Ancient lava flows dating back 2,000 to 15,000 years are shown in light green and red on the left side of this space radar image of the Craters of the Moon National Monument area in Idaho. The volcanic cones that produced these lava flows are the dark points shown within the light green area. Craters of the Moon National Monument is part of the Snake River Plain volcanic province. Geologists believe this area was formed as the North American tectonic plate moved across a 'hot spot' which now lies beneath Yellowstone National Park. The irregular patches, shown in red, green and purple in the lower half of the image are lava flows of different ages and surface roughnesses. One of these lava flows is surrounded by agricultural fields, the blue and purple geometric features, in the right center of the image. The town of Arco, Idaho is the bright yellow area on the right side of the agricultural area. The peaks along the top of the image are the White Knob Mountains. The Big Lost River flows out of the canyon at the top right of the image. The image was acquired by the Spaceborne Imaging Radar-C/X-band Synthetic Aperture Radar (SIR-C/X-SAR) when it flew aboard the space shuttle Endeavour on October 5, 1994. This image is centered at 43.58 degrees north latitude, 113.42 degrees west longitude. The area shown is approximately 33 kilometers by 48 kilometers 20.5 miles by 30 miles). Colors are assigned to different frequencies and polarizations of the radar as follows: red is the L-band horizontally transmitted, horizontally received; green is the L-band horizontally transmitted, vertically received; blue is the C-band horizontally transmitted, vertically received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.

  18. Accuracy aspects of stereo side-looking radar. [analysis of its visual perception and binocular vision

    NASA Technical Reports Server (NTRS)

    Leberl, F. W.

    1979-01-01

    The geometry of the radar stereo model and factors affecting visual radar stereo perception are reviewed. Limits to the vertical exaggeration factor of stereo radar are defined. Radar stereo model accuracies are analyzed with respect to coordinate errors caused by errors of radar sensor position and of range, and with respect to errors of coordinate differences, i.e., cross-track distances and height differences.

  19. The Earthcare Cloud Profiling Radar, its PFM development status (Conference Presentation)

    NASA Astrophysics Data System (ADS)

    Nakatsuka, Hirotaka; Tomita, Eichi; Aida, Yoshihisa; Seki, Yoshihiro; Okada, Kazuyuki; Maruyama, Kenta; Ishii, Yasuyuki; Tomiyama, Nobuhiro; Ohno, Yuichi; Horie, Hiroaki; Sato, Kenji

    2016-10-01

    The Earth Clouds, Aerosols and Radiation Explorer (EarthCARE) mission is joint mission between Europe and Japan for the launch year of 2018. Mission objective is to improve scientific understanding of cloud-aerosol-radiation interactions that is one of the biggest uncertain factors for numerical climate and weather predictions. The EarthCARE spacecraft equips four instruments such as an ultra violet lidar (ATLID), a cloud profiling radar (CPR), a broadband radiometer (BBR), and a multi-spectral imager (MSI) and perform complete synergy observation to observe aerosols, clouds and their interactions simultaneously from the orbit. Japan Aerospace Exploration Agency (JAXA) is responsible for development of the CPR in this EarthCARE mission and the CPR will be the first space-borne W-band Doppler radar. The CPR is defined with minimum radar sensitivity of -35dBz (6dB better than current space-borne cloud radar, i.e. CloudSat, NASA), radiometric accuracy of 2.7 dB, and Doppler velocity measurement accuracy of less than 1.3 m/s. These specifications require highly accurate pointing technique in orbit and high power source with large antenna dish. JAXA and National Institute of Information and Communications Technology (NICT) have been jointly developed this CPR to meet these strict requirements so far and then achieved the development such as new CFRP flex-core structure, long life extended interaction klystron, low loss quasi optical feed technique, and so on. Through these development successes, CPR development phase has been progressed to critical design phase. In addition, new ground calibration technique is also being progressed for launch of EarthCARE/CPR. The unique feature of EarthCARE CPR is vertical Doppler velocity measurement capability. Vertical Doppler velocity measurement is very attractive function from the science point of view, because vertical motions of cloud particles are related with cloud microphysics and dynamics. However, from engineering point of

  20. Vertical rise velocity of equatorial plasma bubbles estimated from Equatorial Atmosphere Radar (EAR) observations and HIRB model simulations

    NASA Astrophysics Data System (ADS)

    Tulasi Ram, S.; Ajith, K. K.; Yokoyama, T.; Yamamoto, M.; Niranjan, K.

    2017-06-01

    The vertical rise velocity (Vr) and maximum altitude (Hm) of equatorial plasma bubbles (EPBs) were estimated using the two-dimensional fan sector maps of 47 MHz Equatorial Atmosphere Radar (EAR), Kototabang, during May 2010 to April 2013. A total of 86 EPBs were observed out of which 68 were postsunset EPBs and remaining 18 EPBs were observed around midnight hours. The vertical rise velocities of the EPBs observed around the midnight hours are significantly smaller ( 26-128 m/s) compared to those observed in postsunset hours ( 45-265 m/s). Further, the vertical growth of the EPBs around midnight hours ceases at relatively lower altitudes, whereas the majority of EPBs at postsunset hours found to have grown beyond the maximum detectable altitude of the EAR. The three-dimensional numerical high-resolution bubble (HIRB) model with varying background conditions are employed to investigate the possible factors that control the vertical rise velocity and maximum attainable altitudes of EPBs. The estimated rise velocities from EAR observations at both postsunset and midnight hours are, in general, consistent with the nonlinear evolution of EPBs from the HIRB model. The smaller vertical rise velocities (Vr) and lower maximum altitudes (Hm) of EPBs during midnight hours are discussed in terms of weak polarization electric fields within the bubble due to weaker background electric fields and reduced background ion density levels.Plain Language SummaryEquatorial plasma bubbles are plasma density irregularities in the ionosphere. The radio waves passing through these irregular density structures undergo severe degradation/scintillation that could cause severe disruption of satellite-based communication and augmentation systems such as GPS navigation. These bubbles develop at geomagnetic equator, grow <span class="hlt">vertically</span>, and elongate along the field lines to latitudes away from the equator. The knowledge on bubble rise velocities and their</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01309&hterms=river+urban+city&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Driver%2Burban%2Bcity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01309&hterms=river+urban+city&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Driver%2Burban%2Bcity"><span>Space <span class="hlt">radar</span> image of New York City</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1995-01-01</p> <p>This <span class="hlt">radar</span> image of the New York city metropolitan area. The island of Manhattan appears in the center of the image. The green-colored rectangle on Manhattan is Central Park. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/ X-SAR) aboard the space shuttle Endeavour on October 10, 1994. North is toward the upper right. The area shown is 75.0 kilometers by 48.8 kilometers (46.5 miles by 30.2 miles). The image is centered at 40.7 degrees north latitude and 73.8 degrees west longitude. In general, light blue areas correspond to dense urban development, green areas to moderately vegetated zones and black areas to bodies of water. The Hudson River is the black strip that runs from the left edge to the upper right corner of the image. It separates New Jersey, in the upper left of the image, from New York. The Atlantic Ocean is at the bottom of the image where two barrier islands along the southern shore of Long Island are also visible. John F. Kennedy International Airport is visible above these islands. Long Island Sound, separating Long Island from Connecticut, is the dark area right of the center of the image. Many bridges are visible in the image, including the Verrazano Narrows, George Washington and Brooklyn bridges. The <span class="hlt">radar</span> illumination is from the left of the image; this causes some urban zones to appear red because the streets are at a perpendicular angle to the <span class="hlt">radar</span> pulse. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted, <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted, <span class="hlt">vertically</span> received). <span class="hlt">Radar</span> images like this one could be used as a tool for city planners and resource managers to map and monitor land use patterns. The <span class="hlt">radar</span> imaging systems can clearly detect the variety of landscapes in the area, as well as the density of urban</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040028085','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040028085"><span>Preliminary Analysis of X-Band and Ka-Band <span class="hlt">Radar</span> for Use in the Detection of Icing Conditions Aloft</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Reehorst, Andrew L.; Koenig, George G.</p> <p>2004-01-01</p> <p>NASA and the U.S. Army Cold Regions Research and Engineering Laboratory (CRREL) have an on-going activity to develop remote sensing technologies for the detection and measurement of icing conditions aloft. <span class="hlt">Radar</span> has been identified as a strong tool for this work. However, since the remote detection of icing conditions with the intent to identify areas of icing hazard is a new and evolving capability, there are no set requirements for <span class="hlt">radar</span> sensitivity. This work is an initial attempt to quantify, through analysis, the sensitivity requirements for an icing remote sensing <span class="hlt">radar</span>. The primary <span class="hlt">radar</span> of interest for cloud measurements is Ka-band, however, since NASA is currently using an X-band unit, this frequency is also examined. Several aspects of <span class="hlt">radar</span> signal analysis were examined. Cloud reflectivity was calculated for several forms of cloud using two different techniques. The Air Force Geophysical Laboratory (AFGL) cloud models, with different drop spectra represented by a modified gamma distribution, were utilized to examine several categories of cloud formation. Also a fundamental methods approach was used to allow manipulation of the cloud droplet size spectra. And an analytical icing <span class="hlt">radar</span> simulator was developed to examine the complete <span class="hlt">radar</span> system response to a configurable multi-layer cloud environment. Also discussed is the NASA <span class="hlt">vertical</span> <span class="hlt">pointing</span> X-band <span class="hlt">radar</span>. The <span class="hlt">radar</span> and its data system are described, and several summer weather events are reviewed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950011783','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950011783"><span>Microburst <span class="hlt">vertical</span> wind estimation from horizontal wind measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vicroy, Dan D.</p> <p>1994-01-01</p> <p>The <span class="hlt">vertical</span> wind or downdraft component of a microburst-generated wind shear can significantly degrade airplane performance. Doppler <span class="hlt">radar</span> and lidar are two sensor technologies being tested to provide flight crews with early warning of the presence of hazardous wind shear. An inherent limitation of Doppler-based sensors is the inability to measure velocities perpendicular to the line of sight, which results in an underestimate of the total wind shear hazard. One solution to the line-of-sight limitation is to use a <span class="hlt">vertical</span> wind model to estimate the <span class="hlt">vertical</span> component from the horizontal wind measurement. The objective of this study was to assess the ability of simple <span class="hlt">vertical</span> wind models to improve the hazard prediction capability of an airborne Doppler sensor in a realistic microburst environment. Both simulation and flight test measurements were used to test the <span class="hlt">vertical</span> wind models. The results indicate that in the altitude region of interest (at or below 300 m), the simple <span class="hlt">vertical</span> wind models improved the hazard estimate. The <span class="hlt">radar</span> simulation study showed that the magnitude of the performance improvement was altitude dependent. The altitude of maximum performance improvement occurred at about 300 m.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090032059','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090032059"><span><span class="hlt">Vertical</span> Cloud Climatology During TC4 Derived from High-Altitude Aircraft Merged Lidar and <span class="hlt">Radar</span> Profiles</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hlavka, Dennis; Tian, Lin; Hart, William; Li, Lihua; McGill, Matthew; Heymsfield, Gerald</p> <p>2009-01-01</p> <p>Aircraft lidar works by shooting laser pulses toward the earth and recording the return time and intensity of any of the light returning to the aircraft after scattering off atmospheric particles and/or the Earth s surface. The scattered light signatures can be analyzed to tell the exact location of cloud and aerosol layers and, with the aid of a few optical assumptions, can be analyzed to retrieve estimates of optical properties such as atmospheric transparency. <span class="hlt">Radar</span> works in a similar fashion except it sends pulses toward earth at a much larger wavelength than lidar. <span class="hlt">Radar</span> records the return time and intensity of cloud or rain reflection returning to the aircraft. Lidar can measure scatter from optically thin cirrus and aerosol layers whose particles are too small for the <span class="hlt">radar</span> to detect. <span class="hlt">Radar</span> can provide reflection profiles through thick cloud layers of larger particles that lidar cannot penetrate. Only after merging the two instrument products can accurate measurements of the locations of all layers in the full atmospheric column be achieved. Accurate knowledge of the <span class="hlt">vertical</span> distribution of clouds is important information for understanding the Earth/atmosphere radiative balance and for improving weather/climate forecast models. This paper describes one such merged data set developed from the Tropical Composition, Cloud and Climate Coupling (TC4) experiment based in Costa Rica in July-August 2007 using the nadir viewing Cloud Physics Lidar (CPL) and the Cloud <span class="hlt">Radar</span> System (CRS) on board the NASA ER-2 aircraft. Statistics were developed concerning cloud probability through the atmospheric column and frequency of the number of cloud layers. These statistics were calculated for the full study area, four sub-regions, and over land compared to over ocean across all available flights. The results are valid for the TC4 experiment only, as preferred cloud patterns took priority during mission planning. The TC4 Study Area was a very cloudy region, with cloudy</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1393573','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1393573"><span>Cloud <span class="hlt">vertical</span> distribution from combined surface and space <span class="hlt">radar</span>-lidar observations at two Arctic atmospheric observatories</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>Liu, Yinghui; Shupe, Matthew D.; Wang, Zhien</p> <p></p> <p>Detailed and accurate <span class="hlt">vertical</span> distributions of cloud properties (such as cloud fraction, cloud phase, and cloud water content) and their changes are essential to accurately calculate the surface radiative flux and to depict the mean climate state. Surface and space-based active sensors including <span class="hlt">radar</span> and lidar are ideal to provide this information because of their superior capability to detect clouds and retrieve cloud microphysical properties. In this study, we compare the annual cycles of cloud property <span class="hlt">vertical</span> distributions from space-based active sensors and surface-based active sensors at two Arctic atmospheric observatories, Barrow and Eureka. Based on the comparisons, we identifymore » the sensors' respective strengths and limitations, and develop a blended cloud property <span class="hlt">vertical</span> distribution by combining both sets of observations. Results show that surface-based observations offer a more complete cloud property <span class="hlt">vertical</span> distribution from the surface up to 11 km above mean sea level (a.m.s.l.) with limitations in the middle and high altitudes; the annual mean total cloud fraction from space-based observations shows 25-40 % fewer clouds below 0.5 km than from surface-based observations, and space-based observations also show much fewer ice clouds and mixed-phase clouds, and slightly more liquid clouds, from the surface to 1 km. In general, space-based observations show comparable cloud fractions between 1 and 2 km a.m.s.l., and larger cloud fractions above 2 km a.m.s.l. than from surface-based observations. A blended product combines the strengths of both products to provide a more reliable annual cycle of cloud property <span class="hlt">vertical</span> distributions from the surface to 11 km a.m.s.l. This information can be valuable for deriving an accurate surface radiative budget in the Arctic and for cloud parameterization evaluation in weather and climate models. Cloud annual cycles show similar evolutions in total cloud fraction and ice cloud fraction, and lower liquid</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1393573-cloud-vertical-distribution-from-combined-surface-space-radar-lidar-observations-two-arctic-atmospheric-observatories','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1393573-cloud-vertical-distribution-from-combined-surface-space-radar-lidar-observations-two-arctic-atmospheric-observatories"><span>Cloud <span class="hlt">vertical</span> distribution from combined surface and space <span class="hlt">radar</span>-lidar observations at two Arctic atmospheric observatories</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Liu, Yinghui; Shupe, Matthew D.; Wang, Zhien; ...</p> <p>2017-05-16</p> <p>Detailed and accurate <span class="hlt">vertical</span> distributions of cloud properties (such as cloud fraction, cloud phase, and cloud water content) and their changes are essential to accurately calculate the surface radiative flux and to depict the mean climate state. Surface and space-based active sensors including <span class="hlt">radar</span> and lidar are ideal to provide this information because of their superior capability to detect clouds and retrieve cloud microphysical properties. In this study, we compare the annual cycles of cloud property <span class="hlt">vertical</span> distributions from space-based active sensors and surface-based active sensors at two Arctic atmospheric observatories, Barrow and Eureka. Based on the comparisons, we identifymore » the sensors' respective strengths and limitations, and develop a blended cloud property <span class="hlt">vertical</span> distribution by combining both sets of observations. Results show that surface-based observations offer a more complete cloud property <span class="hlt">vertical</span> distribution from the surface up to 11 km above mean sea level (a.m.s.l.) with limitations in the middle and high altitudes; the annual mean total cloud fraction from space-based observations shows 25-40 % fewer clouds below 0.5 km than from surface-based observations, and space-based observations also show much fewer ice clouds and mixed-phase clouds, and slightly more liquid clouds, from the surface to 1 km. In general, space-based observations show comparable cloud fractions between 1 and 2 km a.m.s.l., and larger cloud fractions above 2 km a.m.s.l. than from surface-based observations. A blended product combines the strengths of both products to provide a more reliable annual cycle of cloud property <span class="hlt">vertical</span> distributions from the surface to 11 km a.m.s.l. This information can be valuable for deriving an accurate surface radiative budget in the Arctic and for cloud parameterization evaluation in weather and climate models. Cloud annual cycles show similar evolutions in total cloud fraction and ice cloud fraction, and lower liquid</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01769.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01769.html"><span>Space <span class="hlt">Radar</span> Image of Long Valley, California in 3-D</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This three-dimensional perspective view of Long Valley, California was created from data taken by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> on board the space shuttle Endeavour. This image was constructed by overlaying a color composite SIR-C <span class="hlt">radar</span> image on a digital elevation map. The digital elevation map was produced using <span class="hlt">radar</span> interferometry, a process by which <span class="hlt">radar</span> data are acquired on different passes of the space shuttle. The two data passes are compared to obtain elevation information. The interferometry data were acquired on April 13,1994 and on October 3, 1994, during the first and second flights of the SIR-C/X-SAR instrument. The color composite <span class="hlt">radar</span> image was taken in October and was produced by assigning red to the C-band (horizontally transmitted and <span class="hlt">vertically</span> received) polarization; green to the C-band (<span class="hlt">vertically</span> transmitted and received) polarization; and blue to the ratio of the two data sets. Blue areas in the image are smooth and yellow areas are rock outcrops with varying amounts of snow and vegetation. The view is looking north along the northeastern edge of the Long Valley caldera, a volcanic collapse feature created 750,000 years ago and the site of continued subsurface activity. Crowley Lake is the large dark feature in the foreground. http://photojournal.jpl.nasa.gov/catalog/PIA01769</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030071078&hterms=break+even+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dbreak%2Beven%2Banalysis','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030071078&hterms=break+even+analysis&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dbreak%2Beven%2Banalysis"><span>Archetypal TRMM <span class="hlt">Radar</span> Profiles Identified Through Cluster Analysis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Boccippio, Dennis J.</p> <p>2003-01-01</p> <p>It is widely held that identifiable 'convective regimes' exist in nature, although precise definitions of these are elusive. Examples include land / Ocean distinctions, break / monsoon beahvior, seasonal differences in the Amazon (SON vs DJF), etc. These regimes are often described by differences in the realized local convective spectra, and measured by various metrics of convective intensity, depth, areal coverage and rainfall amount. Objective regime identification may be valuable in several ways: regimes may serve as natural 'branch <span class="hlt">points</span>' in satellite retrieval algorithms or data assimilation efforts; one example might be objective identification of regions that 'should' share a similar 2-R relationship. Similarly, objectively defined regimes may provide guidance on optimal siting of ground validation efforts. Objectively defined regimes could also serve as natural (rather than arbitrary geographic) domain 'controls' in studies of convective response to environmental forcing. Quantification of convective <span class="hlt">vertical</span> structure has traditionally involved parametric study of prescribed quantities thought to be important to convective dynamics: maximum <span class="hlt">radar</span> reflectivity, cloud top height, 30-35 dBZ echo top height, rain rate, etc. Individually, these parameters are somewhat deficient as their interpretation is often nonunique (the same metric value may signify different physics in different storm realizations). Individual metrics also fail to capture the coherence and interrelationships between <span class="hlt">vertical</span> levels available in full 3-D <span class="hlt">radar</span> datasets. An alternative approach is discovery of natural partitions of <span class="hlt">vertical</span> structure in a globally representative dataset, or 'archetypal' reflectivity profiles. In this study, this is accomplished through cluster analysis of a very large sample (0[107) of TRMM-PR reflectivity columns. Once achieved, the rainconditional and unconditional 'mix' of archetypal profile types in a given location and/or season provides a description</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70024856','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70024856"><span><span class="hlt">Vertical</span> velocity variance in the mixed layer from <span class="hlt">radar</span> wind profilers</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Eng, K.; Coulter, R.L.; Brutsaert, W.</p> <p>2003-01-01</p> <p><span class="hlt">Vertical</span> velocity variance data were derived from remotely sensed mixed layer turbulence measurements at the Atmospheric Boundary Layer Experiments (ABLE) facility in Butler County, Kansas. These measurements and associated data were provided by a collection of instruments that included two 915 MHz wind profilers, two radio acoustic sounding systems, and two eddy correlation devices. The data from these devices were available through the Atmospheric Boundary Layer Experiment (ABLE) database operated by Argonne National Laboratory. A signal processing procedure outlined by Angevine et al. was adapted and further built upon to derive <span class="hlt">vertical</span> velocity variance, w_pm???2, from 915 MHz wind profiler measurements in the mixed layer. The proposed procedure consisted of the application of a height-dependent signal-to-noise ratio (SNR) filter, removal of outliers plus and minus two standard deviations about the mean on the spectral width squared, and removal of the effects of beam broadening and <span class="hlt">vertical</span> shearing of horizontal winds. The scatter associated with w_pm???2 was mainly affected by the choice of SNR filter cutoff values. Several different sets of cutoff values were considered, and the optimal one was selected which reduced the overall scatter on w_pm???2 and yet retained a sufficient number of data <span class="hlt">points</span> to average. A similarity relationship of w_pm???2 versus height was established for the mixed layer on the basis of the available data. A strong link between the SNR and growth/decay phases of turbulence was identified. Thus, the mid to late afternoon hours, when strong surface heating occurred, were observed to produce the highest quality signals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800015040','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800015040"><span>The Urbana coherent-scatter <span class="hlt">radar</span>: Synthesis and first results</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Gibbs, K. P.; Bowhill, S. A.</p> <p>1979-01-01</p> <p>A coherent scatter <span class="hlt">radar</span> system was synthesized and several hundred hours of echo power and line of sight velocity data obtained. The coherent scatter <span class="hlt">radar</span> utilizes a diode array and components from meteor <span class="hlt">radar</span>. The receiving system permits a time resolution of one minute in the data. Echo power from the D region shows a high degree of variability from day to day. Examples of changes in power level at shorter time scales are observed. Velocity data show the existence of gravity waves and occasionally exhibit <span class="hlt">vertical</span> standing wave characteristics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1378031-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1378031-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia"><span>A case study of microphysical structures and hydrometeor phase in convection using <span class="hlt">radar</span> Doppler spectra at Darwin, Australia</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>Riihimaki, Laura D.; Comstock, Jennifer M.; Luke, Edward</p> <p></p> <p>To understand the microphysical processes that impact diabatic heating and cloud lifetimes in convection, we need to characterize the spatial distribution of supercooled liquid water. To address this observational challenge, <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors at the Darwin Atmospheric Radiation Measurement (ARM) site are used to classify cloud phase within a deep convective cloud in a shallow to deep convection transitional case. The cloud cannot be fully observed by a lidar due to signal attenuation. Thus we develop an objective method for identifying hydrometeor classes, including mixed-phase conditions, using k-means clustering on parameters that describe the shape of the Doppler spectramore » from <span class="hlt">vertically</span> <span class="hlt">pointing</span> Ka band cloud <span class="hlt">radar</span>. This approach shows that multiple, overlapping mixed-phase layers exist within the cloud, rather than a single region of supercooled liquid, indicating complexity to how ice growth and diabatic heating occurs in the <span class="hlt">vertical</span> structure of the cloud.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ak0486.photos.193535p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ak0486.photos.193535p/"><span>50. View of waveguides beginning to move toward two <span class="hlt">radar</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>50. View of waveguides beginning to move toward two <span class="hlt">radar</span> scanner switches (two per <span class="hlt">radar</span> scanner building) by <span class="hlt">vertical</span> bends; also tuning devices are located here. - Clear Air Force Station, Ballistic Missile Early Warning System Site II, One mile west of mile marker 293.5 on Parks Highway, 5 miles southwest of Anderson, Anderson, Denali Borough, AK</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ESASP.724E.105W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ESASP.724E.105W"><span>Fusion of Cross-Track TerraSAR-X PS <span class="hlt">Point</span> Clouds over Las Vegas</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Ziyun; Balz, Timo; Wei, Lianhuan; Liao, Mingsheng</p> <p>2014-11-01</p> <p>Persistent scatterer interferometry (PS-InSAR) is widely used in <span class="hlt">radar</span> remote sensing. However, because the surface motion is estimated in the line-of-sight (LOS) direction, it is not possible to differentiate between <span class="hlt">vertical</span> and horizontal surface motions from a single stack. Cross-track data, i.e. the combination of data from ascending and descending orbits, allows us to better analyze the deformation and to obtain 3d motion information. We implemented a cross-track fusion of PS-InSAR <span class="hlt">point</span> cloud data, making it possible to separate the <span class="hlt">vertical</span> and horizontal components of the surface motion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850024156','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850024156"><span>Observations of frontal zone structures with a VHF Doppler <span class="hlt">radar</span> and radiosondes, part 1.2A</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Larsen, M. F.; Rottger, J.</p> <p>1984-01-01</p> <p>The SOUSY-VHF-<span class="hlt">Radar</span> is a pulsed coherent <span class="hlt">radar</span> operating at 53.5 MHz and located near Bad Lauterbert, West Germany. Since 1977, the facility, operated by the Max-Planck-Institut fur Aeronomie, has been used to make a series of frontal passage observations in the spring and fall. Experiments in winter have been difficult because part of the transmitting and receiving array is usually covered by snow during that part of the year. Wavelengths around 6 m are known to be sensitive to the <span class="hlt">vertical</span> temperature structure of the atmosphere (GREEN and GAGE, 1980; RASTOGI and ROTTGER, 1982). Thus, it has been possible to use <span class="hlt">radars</span> operating at frequencies near 500 MHz to locate the tropopause. Comparisons between <span class="hlt">radar</span> data and radiosonde data have shown that there is a large gradient in the <span class="hlt">radar</span> reflectivity at the height where the radiosonde tropopause occurs. An experiment carried out by ROTTGER (1979) on March 15 to 16, 1977, showed that the <span class="hlt">radar</span>'s sensitivity to the <span class="hlt">vertical</span> temperature structure could also be used to locate the position of fronts. The SOUSY-VHF-<span class="hlt">Radar</span> consists of a transmitting array, also used for receiving in some configurations, that can be scanned in the off-<span class="hlt">vertical</span> direction but not at sufficiently low elevation angles to study the horizontal extent of structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01305&hterms=active+volcanoes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dactive%2Bvolcanoes','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01305&hterms=active+volcanoes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dactive%2Bvolcanoes"><span>Space <span class="hlt">radar</span> image of Galeras Volcano, Colombia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1995-01-01</p> <p>This <span class="hlt">radar</span> image of the area surrounding the Galeras volcano in southern Colombia shows the ability of a multi-frequency <span class="hlt">radar</span> to map volcanic structures that can be dangerous to study on the ground. Galeras has erupted more than 20 times since the area was first visited by European explorers in the 1500s. Volcanic activity levels have been high in the last five years, including an eruption in January 1993 that killed nine people on a scientific expedition to the volcano summit. Galeras is the light green area near the center of the image. The active cone, with a small summit pit, is the red feature nestled against the lower right edge of the caldera (crater) wall. The city of Pasto, with a population of 300,000, is shown in orange near the bottom of the image, just 8 kilometers (5 miles) from the volcano. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/ X-SAR) aboard the space shuttle Endeavour on its 96th orbit on April 15, 1994. North is toward the upper right. The area shown is 49.1 by 36.0 kilometers (30.5 by 22.3 miles), centered at 1.2 degrees north latitude and 77.4 degrees west longitude. The <span class="hlt">radar</span> illumination is from the top of the image. The false colors in this image were created using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted, <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted, <span class="hlt">vertically</span> received). Galeras is one of 15 volcanoes worldwide that are being monitored by the scientific community as an 'International Decade Volcano' because of the hazard that it represents to the local population.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A23I3355P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A23I3355P"><span>RaInCube: a proposed constellation of precipitation profiling <span class="hlt">Radars</span> In Cubesat</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Peral, E.; Tanelli, S.; Haddad, Z. S.; Stephens, G. L.; Im, E.</p> <p>2014-12-01</p> <p>Precipitation <span class="hlt">radars</span> in Low-Earth-Orbit provide <span class="hlt">vertically</span> resolved profiles of rain and snow on a global scale. With the recent advances in miniaturized <span class="hlt">radar</span> and CubeSat/SmallSat technologies, it would now be feasible to launch multiple copies of the same <span class="hlt">radar</span> instrument in desirable formations to allow measurements of short time scale evolution of atmospheric processes. One such concept is the novel <span class="hlt">radar</span> architecture compatible with the 6U CubeSat class that is being developed at JPL by exploiting simplification and miniaturization of the <span class="hlt">radar</span> subsystems. The RaInCube architecture would significantly reduce the number of components, power consumption and mass with respect to existing spaceborne <span class="hlt">radars</span>. The baseline RaInCube instrument configuration would be a fixed nadir-<span class="hlt">pointing</span> profiler at Ka-band with a minimum detectable reflectivity better than +10 dBZ at 250m range resolution and 5 km horizontal resolution. The low cost nature of the RaInCube platform would enable deployment of a constellation of identical copies of the same instrument in various relative positions in LEO to address specific observational gaps left open by the current missions that require high-resolution <span class="hlt">vertical</span> profiling capability. A constellation of only four RaInCubes would populate the precipitation statistics in a distributed fashion across the globe and across the times of day, and therefore, would enable substantially better sampling of the diurnal cycle statistics. One could extend this scheme by adding more RaInCubes in each of the orbital planes, and phase them once in orbit so that they would be separated by an arbitrary amount of time among them. Wide separations (say 20-30 min) would further extend the sampling of the diurnal cycle to sub-hourly scales. Narrower time separations between RaInCubes would allow studying the evolution of convective systems at the convective time scale in each region of interest and would reveal the dominant modes of evolution of each</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20130010396','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20130010396"><span><span class="hlt">Radar</span> Scan Strategies for the Patrick Air Force Base Weather Surveillance <span class="hlt">Radar</span>, Model-74C, Replacement</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Short, David</p> <p>2008-01-01</p> <p>The 45th Weather Squadron (45 WS) is replacing the Weather Surveillance <span class="hlt">Radar</span>, Model 74C (WSR-74C) at Patrick Air Force Base (PAFB), with a Doppler, dual polarization <span class="hlt">radar</span>, the Radtec 43/250. A new scan strategy is needed for the Radtec 43/250, to provide high <span class="hlt">vertical</span> resolution data over the Kennedy Space Center (KSC) and Cape Canaveral Air Force Station (CCAFS) launch pads, while taking advantage of the new <span class="hlt">radar</span>'s advanced capabilities for detecting severe weather phenomena associated with convection within the 45 WS area of responsibility. The Applied Meteorology Unit (AMU) developed several scan strategies customized for the operational needs of the 45 WS. The AMU also developed a plan for evaluating the scan strategies in the period prior to operational acceptance, currently scheduled for November 2008.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014WRR....50.8571D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014WRR....50.8571D"><span>Exploration of discrepancy between <span class="hlt">radar</span> and gauge rainfall estimates driven by wind fields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dai, Qiang; Han, Dawei</p> <p>2014-11-01</p> <p>Due to the fact that weather <span class="hlt">radar</span> is prone to several sources of errors, it is acknowledged that adjustment against ground observations such as rain gauges is crucial for <span class="hlt">radar</span> measurement. Spatial matching of precipitation patterns between <span class="hlt">radar</span> and rain gauge is a significant premise in <span class="hlt">radar</span> bias corrections. It is a conventional way to construct <span class="hlt">radar</span>-gauge pairs based on their <span class="hlt">vertical</span> locations. However, due to the wind effects, the raindrops observed by the <span class="hlt">radar</span> do not always fall <span class="hlt">vertically</span> to the ground, and the raindrops arriving at the ground may not all be caught by the rain gauge. This study proposes a fully formulated scheme to numerically simulate the movement of raindrops in a three-dimensional wind field in order to adjust the wind-induced errors. The Brue catchment (135 km2) in Southwest England covering 28 <span class="hlt">radar</span> pixels and 49 rain gauges is an experimental catchment, where the <span class="hlt">radar</span> central beam height varies between 500 and 700 m. The 20 typical events (with durations of 6-36 h) are chosen to assess the correlation between hourly <span class="hlt">radar</span> and gauge rainfall surfaces. It is found that for most events, the improved rates of correlation coefficients are greater than 10%, and nearly half of the events increase by 20%. With the proposed method, except four events, all the event-averaged correlation values are greater than 0.5. This work is the first study to tackle both wind effects on <span class="hlt">radar</span> and rain gauges, which could be considered as one of the essential components in processing <span class="hlt">radar</span> observational data in its hydrometeorological applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910040010&hterms=Per&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPer','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910040010&hterms=Per&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DPer"><span>Per-<span class="hlt">point</span> and per-field contextual classification of multipolarization and multiple incidence angle aircraft L-band <span class="hlt">radar</span> data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hoffer, Roger M.; Hussin, Yousif Ali</p> <p>1989-01-01</p> <p>Multipolarized aircraft L-band <span class="hlt">radar</span> data are classified using two different image classification algorithms: (1) a per-<span class="hlt">point</span> classifier, and (2) a contextual, or per-field, classifier. Due to the distinct variations in <span class="hlt">radar</span> backscatter as a function of incidence angle, the data are stratified into three incidence-angle groupings, and training and test data are defined for each stratum. A low-pass digital mean filter with varied window size (i.e., 3x3, 5x5, and 7x7 pixels) is applied to the data prior to the classification. A predominately forested area in northern Florida was the study site. The results obtained by using these image classifiers are then presented and discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20060042799&hterms=rain&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Drain','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060042799&hterms=rain&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Drain"><span>Backscattering enhancement for Marshall-Palmer distributed rains for a W-band nadir-<span class="hlt">pointing</span> <span class="hlt">radar</span> with a finite beam width</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kobayashi, Satoru; Tanelli, Simone; Im, Eastwood; Oguchi, Tomohiro</p> <p>2005-01-01</p> <p>In this paper, we expand the previous theory to be applied to a generic drop size distribution with spheroidal raindrops including spherical raindrops. Results will be used to discuss the multiple scattering effects on the backscatter measurements acquired by a W-band nadir-<span class="hlt">pointing</span> <span class="hlt">radar</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_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01796&hterms=image+alignment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dimage%2Balignment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01796&hterms=image+alignment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dimage%2Balignment"><span>Space <span class="hlt">Radar</span> Image of Pishan, China</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image is centered near the small town of Pishan in northwest China, about 280 km (174 miles) southeast of the city of Kashgar along the ancient Silk Route in the Taklamakan desert of the Xinjiang Province. Geologists are using this <span class="hlt">radar</span> image as a map to study past climate changes and tectonics of the area. The irregular lavender branching patterns in the center of the image are the remains of ancient alluvial fans, gravel deposits that have accumulated at the base of the mountains during times of wetter climate. The subtle striped pattern cutting across the ancient fans are caused by thrusting of the Kun Lun Mountains north. This motion is caused by the continuing plate-tectonic collision of India with Asia. Modern fans show up as large lavender triangles above the ancient fan deposits. Yellow areas on the modern fans are vegetated oases. The gridded pattern results from the alignment of poplar trees that have been planted as wind breaks. The reservoir at the top of the image is part of a sophisticated irrigation system that supplies water to the oases. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour in April 1994. This image is centered at 37.4 degrees north latitude, 78.3 degrees east longitude and shows an area approximately 50 km by 100 km (31 miles by 62 miles). The colors are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: Red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, <span class="hlt">vertically</span> received; and blue is C-band horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian, and the United States space agencies, is part of NASA's Mission to Planet Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001AtmRe..59..231B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001AtmRe..59..231B"><span>Analysis of polarization <span class="hlt">radar</span> returns from ice clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Battaglia, A.; Sturniolo, O.; Prodi, F.</p> <p></p> <p>Using a modified T-matrix code, some polarimetric single-scattering <span class="hlt">radar</span> parameters ( Zh,v, LDR h,v, ρhv, ZDR and δhv) from populations of ice crystals in ice phase at 94 GHz, modeled with axisymmetric prolate and oblate spheroidal shapes for a Γ-size distribution with different α parameter ( α=0, 1, 2) and characteristic dimension Lm varying from 0.1 to 1.8 mm, have been computed. Some of the results for different <span class="hlt">radar</span> elevation angles and different orientation distribution for fixed water content are shown. Deeper analysis has been carried out for pure extensive <span class="hlt">radar</span> polarimetric variables; all of them are strongly dependent on the shapes (characterised by the aspect ratio), the canting angle and the <span class="hlt">radar</span> elevation angle. Quantities like ZDR or δhv at side incidence or LDR h and ρhv at <span class="hlt">vertical</span> incidence can be used to investigate the preferred orientation of the particles and, in some cases, their habits. We analyze scatterplots using couples of pure extensive variables. The scatterplots with the most evident clustering properties for the different habits seem to be those in the ( ZDR [ χ=0°], δhv [ χ=0°]), in the ( ZDR [ χ=0°], LDR h [ χ=90°]) and in the ( ZDR [ χ=0°], ρhv [ χ=90°]) plane. Among these, the most appealing one seems to be that involving ZDR and ρhv variables. To avoid the problem of having simultaneous measurements with a side and a <span class="hlt">vertical</span>-looking <span class="hlt">radar</span>, we believe that measurements of these two extensive variables using a <span class="hlt">radar</span> with an elevation angle around 45° can be an effective instrument to identify different habits. In particular, this general idea can be useful for future space-borne polarimetric <span class="hlt">radars</span> involved in the studies of high ice clouds. It is also believed that these results can be used in next challenge of developing probabilistic and expert methods for identifying hydrometeor types by W-band <span class="hlt">radars</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A42C..06L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A42C..06L"><span>Potential of Higher Moments of the <span class="hlt">Radar</span> Doppler Spectrum for Studying Ice Clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Loehnert, U.; Maahn, M.</p> <p>2015-12-01</p> <p>More observations of ice clouds are required to fill gaps in understanding of microphysical properties and processes. However, in situ observations by aircraft are costly and cannot provide long term observations which are required for a deeper understanding of the processes. Ground based remote sensing observations have the potential to fill this gap, but their observations do not contain sufficient information to unambiguously constrain ice cloud properties which leads to high uncertainties. For <span class="hlt">vertically</span> <span class="hlt">pointing</span> cloud <span class="hlt">radars</span>, usually only reflectivity and mean Doppler velocity are used for retrievals; some studies proposed also the use of Doppler spectrum width.In this study, it is investigated whether additional information can be obtained by exploiting also higher moments of the Doppler spectrum such as skewness and kurtosis together with the slope of the Doppler peak. For this, observations of pure ice clouds from the Indirect and Semi-Direct Aerosol Campaign (ISDAC) in Alaska 2008 are analyzed. Using the ISDAC data set, an Optimal Estimation based retrieval is set up based on synthetic and real <span class="hlt">radar</span> observations. The passive and active microwave radiative transfer model (PAMTRA) is used as a forward model together with the Self-Similar Rayleigh-Gans approximation for estimation of the scattering properties. The state vector of the retrieval consists of the parameters required to simulate the <span class="hlt">radar</span> Doppler spectrum and describes particle mass, cross section area, particle size distribution, and kinematic conditions such as turbulence and <span class="hlt">vertical</span> air motion. Using the retrieval, the information content (degrees of freedom for signal) is quantified that higher moments and slopes can contribute to an ice cloud retrieval. The impact of multiple frequencies, <span class="hlt">radar</span> sensitivity and <span class="hlt">radar</span> calibration is studied. For example, it is found that a single-frequency measurement using all moments and slopes contains already more information content than a dual</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A42C..06L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A42C..06L"><span>Potential of Higher Moments of the <span class="hlt">Radar</span> Doppler Spectrum for Studying Ice Clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lunt, M. F.; Rigby, M. L.; Ganesan, A.; Manning, A.; O'Doherty, S.; Prinn, R. G.; Saito, T.; Harth, C. M.; Muhle, J.; Weiss, R. F.; Salameh, P.; Arnold, T.; Yokouchi, Y.; Krummel, P. B.; Steele, P.; Fraser, P. J.; Li, S.; Park, S.; Kim, J.; Reimann, S.; Vollmer, M. K.; Lunder, C. R.; Hermansen, O.; Schmidbauer, N.; Young, D.; Simmonds, P. G.</p> <p>2014-12-01</p> <p>More observations of ice clouds are required to fill gaps in understanding of microphysical properties and processes. However, in situ observations by aircraft are costly and cannot provide long term observations which are required for a deeper understanding of the processes. Ground based remote sensing observations have the potential to fill this gap, but their observations do not contain sufficient information to unambiguously constrain ice cloud properties which leads to high uncertainties. For <span class="hlt">vertically</span> <span class="hlt">pointing</span> cloud <span class="hlt">radars</span>, usually only reflectivity and mean Doppler velocity are used for retrievals; some studies proposed also the use of Doppler spectrum width.In this study, it is investigated whether additional information can be obtained by exploiting also higher moments of the Doppler spectrum such as skewness and kurtosis together with the slope of the Doppler peak. For this, observations of pure ice clouds from the Indirect and Semi-Direct Aerosol Campaign (ISDAC) in Alaska 2008 are analyzed. Using the ISDAC data set, an Optimal Estimation based retrieval is set up based on synthetic and real <span class="hlt">radar</span> observations. The passive and active microwave radiative transfer model (PAMTRA) is used as a forward model together with the Self-Similar Rayleigh-Gans approximation for estimation of the scattering properties. The state vector of the retrieval consists of the parameters required to simulate the <span class="hlt">radar</span> Doppler spectrum and describes particle mass, cross section area, particle size distribution, and kinematic conditions such as turbulence and <span class="hlt">vertical</span> air motion. Using the retrieval, the information content (degrees of freedom for signal) is quantified that higher moments and slopes can contribute to an ice cloud retrieval. The impact of multiple frequencies, <span class="hlt">radar</span> sensitivity and <span class="hlt">radar</span> calibration is studied. For example, it is found that a single-frequency measurement using all moments and slopes contains already more information content than a dual</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01848&hterms=route+tourist&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Droute%2Btourist','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01848&hterms=route+tourist&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Droute%2Btourist"><span>Space <span class="hlt">Radar</span> Image of County Kerry, Ireland</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>The Iveragh Peninsula, one of the four peninsulas in southwestern Ireland, is shown in this spaceborne <span class="hlt">radar</span> image. The lakes of Killarney National Park are the green patches on the left side of the image. The mountains to the right of the lakes include the highest peaks (1,036 meters or 3,400 feet) in Ireland. The patchwork patterns between the mountains are areas of farming and grazing. The delicate patterns in the water are caused by refraction of ocean waves around the peninsula edges and islands, including Skellig Rocks at the right edge of the image. The Skelligs are home to a 15th century monastery and flocks of puffins. The region is part of County Kerry and includes a road called the 'Ring of Kerry' that is one of the most famous tourist routes in Ireland. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the Space Shuttle Endeavour on April 12, 1994. The image is 82 kilometers by 42 kilometers (51 miles by 26 miles) and is centered at 52.0 degrees north latitude, 9.9 degrees west longitude. North is toward the lower left. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band, horizontally transmitted and received; green is L-band, <span class="hlt">vertically</span> transmitted and received; and blue is C-band, <span class="hlt">vertically</span> transmitted and received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940011417','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940011417"><span>Spaceborne Imaging <span class="hlt">Radar</span>-C instrument</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Huneycutt, Bryan L.</p> <p>1993-01-01</p> <p>The Spaceborne Imaging <span class="hlt">Radar</span>-C is the next <span class="hlt">radar</span> in the series of spaceborne <span class="hlt">radar</span> experiments, which began with Seasat and continued with SIR-A and SIR-B. The SIR-C instrument has been designed to obtain simultaneous multifrequency and simultaneous multipolarization <span class="hlt">radar</span> images from a low earth orbit. It is a multiparameter imaging <span class="hlt">radar</span> that will be flown during at least two different seasons. The instrument operates in the squint alignment mode, the extended aperture mode, the scansar mode, and the interferometry mode. The instrument uses engineering techniques such as beam nulling for echo tracking, pulse repetition frequency hopping for Doppler centroid tracking, generating the frequency step chirp for <span class="hlt">radar</span> parameter flexibility, block floating-<span class="hlt">point</span> quantizing for data rate compression, and elevation beamwidth broadening for increasing the swath illumination.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmRe.201..116L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmRe.201..116L"><span>Intercomparison of attenuation correction algorithms for single-polarized X-band <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lengfeld, K.; Berenguer, M.; Sempere Torres, D.</p> <p>2018-03-01</p> <p>Attenuation due to liquid water is one of the largest uncertainties in <span class="hlt">radar</span> observations. The effects of attenuation are generally inversely proportional to the wavelength, i.e. observations from X-band <span class="hlt">radars</span> are more affected by attenuation than those from C- or S-band systems. On the other hand, X-band <span class="hlt">radars</span> can measure precipitation fields in higher temporal and spatial resolution and are more mobile and easier to install due to smaller antennas. A first algorithm for attenuation correction in single-polarized systems was proposed by Hitschfeld and Bordan (1954) (HB), but it gets unstable in case of small errors (e.g. in the <span class="hlt">radar</span> calibration) and strong attenuation. Therefore, methods have been developed that restrict attenuation correction to keep the algorithm stable, using e.g. surface echoes (for space-borne <span class="hlt">radars</span>) and mountain returns (for ground <span class="hlt">radars</span>) as a final value (FV), or adjustment of the <span class="hlt">radar</span> constant (C) or the coefficient α. In the absence of mountain returns, measurements from C- or S-band <span class="hlt">radars</span> can be used to constrain the correction. All these methods are based on the statistical relation between reflectivity and specific attenuation. Another way to correct for attenuation in X-band <span class="hlt">radar</span> observations is to use additional information from less attenuated <span class="hlt">radar</span> systems, e.g. the ratio between X-band and C- or S-band <span class="hlt">radar</span> measurements. Lengfeld et al. (2016) proposed such a method based isotonic regression of the ratio between X- and C-band <span class="hlt">radar</span> observations along the <span class="hlt">radar</span> beam. This study presents a comparison of the original HB algorithm and three algorithms based on the statistical relation between reflectivity and specific attenuation as well as two methods implementing additional information of C-band <span class="hlt">radar</span> measurements. Their performance in two precipitation events (one mainly convective and the other one stratiform) shows that a restriction of the HB is necessary to avoid instabilities. A comparison with <span class="hlt">vertically</span> <span class="hlt">pointing</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090042963&hterms=phi&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphi','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090042963&hterms=phi&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dphi"><span><span class="hlt">Radar</span> Differential Phase Signatures of Ice Orientation for the Prediction of Lightning Initiation and Cessation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Carey, L.D.; Petersen, W.A.; Deierling, W.</p> <p>2009-01-01</p> <p>The majority of lightning-related casualties typically occur during thunderstorm initiation (e.g., first flash) or dissipation (e.g., last flash). The physics of electrification and lightning production during thunderstorm initiation is fairly well understood. As such, the literature includes a number of studies presenting various <span class="hlt">radar</span> techniques (using reflectivity and, if available, other dual-polarimetric parameters) for the anticipation of initial electrification and first lightning flash. These <span class="hlt">radar</span> techniques have shown considerable skill at forecasting first flash. On the other hand, electrical processes and lightning production during thunderstorm dissipation are not nearly as well understood and few, if any, successful techniques have been developed to anticipate the last flash and subsequent cessation of lightning. One promising approach involves the use of dual-polarimetric <span class="hlt">radar</span> variables to infer the presence of oriented ice crystals in lightning producing storms. In the absence of strong <span class="hlt">vertical</span> electric fields, ice crystals fall with their largest (semi-major) axis in the horizontal associated with gravitational and aerodynamic forces. In thunderstorms, strong <span class="hlt">vertical</span> electric fields (100-200 kV m(sup -1)) have been shown to orient small (less than 2 mm) ice crystals such that their semi-major axis is <span class="hlt">vertical</span> (or nearly <span class="hlt">vertical</span>). After a lightning flash, the electric field is typically relaxed and prior <span class="hlt">radar</span> research suggests that ice crystals rapidly resume their preferred horizontal orientation. In active thunderstorms, the <span class="hlt">vertical</span> electric field quickly recovers and the ice crystals repeat this cycle of orientation for each nearby flash. This change in ice crystal orientation from primarily horizontal to <span class="hlt">vertical</span> during the development of strong <span class="hlt">vertical</span> electric fields prior to a lightning flash forms the physical basis for anticipating lightning initiation and, potentially, cessation. Research has shown that <span class="hlt">radar</span> reflectivity (Z) and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405960-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-bullen-point-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405960-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-bullen-point-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron Elmendorf AFB, Alaska. Risk assessment Bullen <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-03-18</p> <p>This document contains the baseline human health risk assessment and the ecological risk assessment (ERA) for the Bullen <span class="hlt">Point</span> Distant Early Warning (DEW) Line <span class="hlt">radar</span> installation. Five sites at the Bullen <span class="hlt">Point</span> <span class="hlt">radar</span> installation underwent remedial investigations (RIs) during the summer of 1993. The presence of chemical contamination in the soil, sediments, and surface water at the installation was evaluated and reported in the Bullen <span class="hlt">Point</span> Remedial Investigation/Feasibility Study (RI/FS) (U.S. Air Force 1996). The analytical data reported in the RI/FS form the basis for the human health and ecological risk assessments. The primary chemicals of concern (COCs) at themore » five sites are diesel and gasoline from past spills and/or leaks.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000011581','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000011581"><span>TRMM Precipitation <span class="hlt">Radar</span> Reflectivity Profiles Compared to High-Resolution Airborne and Ground-Based <span class="hlt">Radar</span> Measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heymsfield, G. M.; Geerts, B.; Tian, L.</p> <p>1999-01-01</p> <p>In this paper, TRMM (Tropical Rainfall Measuring Mission Satellite) Precipitation <span class="hlt">Radar</span> (PR) products are evaluated by means of simultaneous comparisons with data from the high-altitude ER-2 Doppler <span class="hlt">Radar</span> (EDOP), as well as ground-based <span class="hlt">radars</span>. The comparison is aimed primarily at the <span class="hlt">vertical</span> reflectivity structure, which is of key importance in TRMM rain type classification and latent heating estimation. The <span class="hlt">radars</span> used in this study have considerably different viewing geometries and resolutions, demanding non-trivial mapping procedures in common earth-relative coordinates. Mapped <span class="hlt">vertical</span> cross sections and mean profiles of reflectivity from the PR, EDOP, and ground-based <span class="hlt">radars</span> are compared for six cases. These cases cover a stratiform frontal rainband, convective cells of various sizes and stages, and a hurricane. For precipitating systems that are large relative to the PR footprint size, PR reflectivity profiles compare very well to high-resolution measurements thresholded to the PR minimum reflectivity, and derived variables such as bright band height and rain types are accurate, even at high PR incidence angles. It was found that for, the PR reflectivity of convective cells small relative to the PR footprint is weaker than in reality. Some of these differences can be explained by non-uniform beam filling. For other cases where strong reflectivity gradients occur within a PR footprint, the reflectivity distribution is spread out due to filtering by the PR antenna illumination pattern. In these cases, rain type classification may err and be biased towards the stratiform type, and the average reflectivity tends to be underestimated. The limited sensitivity of the PR implies that the upper regions of precipitation systems remain undetected and that the PR storm top height estimate is unreliable, usually underestimating the actual storm top height. This applies to all cases but the discrepancy is larger for smaller cells where limited sensitivity is compounded</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AMT....11.1417K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AMT....11.1417K"><span>A simple biota removal algorithm for 35 GHz cloud <span class="hlt">radar</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>Kalapureddy, Madhu Chandra R.; Sukanya, Patra; Das, Subrata K.; Deshpande, Sachin M.; Pandithurai, Govindan; Pazamany, Andrew L.; Ambuj K., Jha; Chakravarty, Kaustav; Kalekar, Prasad; Krishna Devisetty, Hari; Annam, Sreenivas</p> <p>2018-03-01</p> <p>Cloud <span class="hlt">radar</span> reflectivity profiles can be an important measurement for the investigation of cloud <span class="hlt">vertical</span> structure (CVS). However, extracting intended meteorological cloud content from the measurement often demands an effective technique or algorithm that can reduce error and observational uncertainties in the recorded data. In this work, a technique is proposed to identify and separate cloud and non-hydrometeor echoes using the <span class="hlt">radar</span> Doppler spectral moments profile measurements. The <span class="hlt">point</span> and volume target-based theoretical <span class="hlt">radar</span> sensitivity curves are used for removing the receiver noise floor and identified <span class="hlt">radar</span> echoes are scrutinized according to the signal decorrelation period. Here, it is hypothesized that cloud echoes are observed to be temporally more coherent and homogenous and have a longer correlation period than biota. That can be checked statistically using ˜ 4 s sliding mean and standard deviation value of reflectivity profiles. The above step helps in screen out clouds critically by filtering out the biota. The final important step strives for the retrieval of cloud height. The proposed algorithm potentially identifies cloud height solely through the systematic characterization of Z variability using the local atmospheric <span class="hlt">vertical</span> structure knowledge besides to the theoretical, statistical and echo tracing tools. Thus, characterization of high-resolution cloud <span class="hlt">radar</span> reflectivity profile measurements has been done with the theoretical echo sensitivity curves and observed echo statistics for the true cloud height tracking (TEST). TEST showed superior performance in screening out clouds and filtering out isolated insects. TEST constrained with polarimetric measurements was found to be more promising under high-density biota whereas TEST combined with linear depolarization ratio and spectral width perform potentially to filter out biota within the highly turbulent shallow cumulus clouds in the convective boundary layer (CBL). This TEST technique is</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770024409','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770024409"><span>The estimation of <span class="hlt">pointing</span> angle and normalized surface scattering cross section from GEOS-3 <span class="hlt">radar</span> altimeter measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brown, G. S.; Curry, W. J.</p> <p>1977-01-01</p> <p>The statistical error of the <span class="hlt">pointing</span> angle estimation technique is determined as a function of the effective receiver signal to noise ratio. Other sources of error are addressed and evaluated with inadequate calibration being of major concern. The impact of <span class="hlt">pointing</span> error on the computation of normalized surface scattering cross section (sigma) from <span class="hlt">radar</span> and the waveform attitude induced altitude bias is considered and quantitative results are presented. <span class="hlt">Pointing</span> angle and sigma processing algorithms are presented along with some initial data. The intensive mode clean vs. clutter AGC calibration problem is analytically resolved. The use clutter AGC data in the intensive mode is confirmed as the correct calibration set for the sigma computations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19780011403','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19780011403"><span>Broad perspectives in <span class="hlt">radar</span> for ocean measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jain, A.</p> <p>1978-01-01</p> <p>The various active <span class="hlt">radar</span> implementation options available for the measurement functions of interest for the SEASAT follow-on missions were evaluated. These functions include surface feature imaging, surface pressure and <span class="hlt">vertical</span> profile, atmospheric sounding, surface backscatter and wind speed determination, surface current location, wavelength spectra, sea surface topography, and ice/snow thickness. Some concepts for the Synthetic Aperture Imaging <span class="hlt">Radar</span> were examined that may be useful in the design and selection of the implementation options for these missions. The applicability of these instruments for the VOIR mission was also kept under consideration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405977-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-point-lay-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405977-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-point-lay-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group Civil Engineering Squadron, Elmendorf AFB, Alaska. Remedial investigation and feasibility study <span class="hlt">Point</span> Lay <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-03-04</p> <p>The United States Air Force (Air Force) has prepared this Remedial Investigation/Feasibility Study (RI/FS) report to present the results of RI/FS activities at four sites located at the <span class="hlt">Point</span> Lay <span class="hlt">radar</span> installation. The remedial investigation (RI) field activities were conducted at the <span class="hlt">Point</span> Lay <span class="hlt">radar</span> installation during the summer of 1993. The four sites at <span class="hlt">Point</span> Lay were investigated because they were suspected of being contaminated with hazardous substances. RI activities were conducted using methods and procedures specified in the RI/FS Work Plan, Sampling and Analysis Plan (SAP), and Health and Safety Plan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17..722H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17..722H"><span>Classification and correction of the <span class="hlt">radar</span> bright band with polarimetric <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hall, Will; Rico-Ramirez, Miguel; Kramer, Stefan</p> <p>2015-04-01</p> <p>The annular region of enhanced <span class="hlt">radar</span> reflectivity, known as the Bright Band (BB), occurs when the <span class="hlt">radar</span> beam intersects a layer of melting hydrometeors. <span class="hlt">Radar</span> reflectivity is related to rainfall through a power law equation and so this enhanced region can lead to overestimations of rainfall by a factor of up to 5, so it is important to correct for this. The BB region can be identified by using several techniques including hydrometeor classification and freezing level forecasts from mesoscale meteorological models. Advances in dual-polarisation <span class="hlt">radar</span> measurements and continued research in the field has led to increased accuracy in the ability to identify the melting snow region. A method proposed by Kitchen et al (1994), a form of which is currently used operationally in the UK, utilises idealised <span class="hlt">Vertical</span> Profiles of Reflectivity (VPR) to correct for the BB enhancement. A simpler and more computationally efficient method involves the formation of an average VPR from multiple elevations for correction that can still cause a significant decrease in error (Vignal 2000). The purpose of this research is to evaluate a method that relies only on analysis of measurements from an operational C-band polarimetric <span class="hlt">radar</span> without the need for computationally expensive models. Initial results show that LDR is a strong classifier of melting snow with a high Critical Success Index of 97% when compared to the other variables. An algorithm based on idealised VPRs resulted in the largest decrease in error when BB corrected scans are compared to rain gauges and to lower level scans with a reduction in RMSE of 61% for rain-rate measurements. References Kitchen, M., R. Brown, and A. G. Davies, 1994: Real-time correction of weather <span class="hlt">radar</span> data for the effects of bright band, range and orographic growth in widespread precipitation. Q.J.R. Meteorol. Soc., 120, 1231-1254. Vignal, B. et al, 2000: Three methods to determine profiles of reflectivity from volumetric <span class="hlt">radar</span> data to correct</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6779843-small-scale-wind-disturbances-observed-mu-radar-during-passage-typhoon-kelly','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6779843-small-scale-wind-disturbances-observed-mu-radar-during-passage-typhoon-kelly"><span>Small-scale wind disturbances observed by the MU <span class="hlt">radar</span> during the passage of typhoon Kelly</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>Sato, Kaoru</p> <p>1993-02-14</p> <p>This paper describes small-scale wind disturbances associated with Typhoon Kelly (October 1987) that were observed by the MU <span class="hlt">radar</span>, one of the MST (mesosphere, stratosphere, and troposphere) <span class="hlt">radars</span>, for about 60 hours with fine time and height resolution. To elucidate the background of small-scale disturbances, synoptic-scale variation in atmospheric stability related to the typhoon structure during the observation is examined. When the typhoon passed near the MU <span class="hlt">radar</span> site, the structure was no longer axisymmetric. There is deep convection only in north-northeast side of the typhoon while convection behind it is suppressed by a synoptic-scale cold air mass moving eastwardmore » to the west of the typhoon. A change in atmospheric stability over the <span class="hlt">radar</span> site as indicated by echo power profiles is likely due to the passage of the sharp transition zone of convection. Strong small-scale wind disturbances were observed around the typhoon passage. The statistical characteristics are different before (BT) and after (AT) the typhoon passage, especially in frequency spectra of <span class="hlt">vertical</span> wind fluctuations. The spectra for BT are unique compared with earlier studies of <span class="hlt">vertical</span> winds observed by VHF <span class="hlt">radars</span>. Another difference is dominance of a horizontal wind component with a <span class="hlt">vertical</span> wavelength of about 3 km, observed only in AT. Further analyses are made of characteristics and <span class="hlt">vertical</span> momentum fluxes for dominant disturbances. Some disturbances are generated to remove the momentum of cyclonic wind rotation of the typhoon. Deep convection, topographic effects in strong winds, and strong <span class="hlt">vertical</span> shear of horizontal winds around an inversion layer are possible sources of the disturbances. Two monochromatic disturbances lasting for more than 10 h in the lower stratosphere observed in BT and AT are identified as inertio-gravity waves, by obtaining wave parameters consistent with all observed quantities. Both of the inertio-gravity waves propagate energy away from the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01801&hterms=viewing+zone&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dviewing%2Bzone','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01801&hterms=viewing+zone&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dviewing%2Bzone"><span>Space <span class="hlt">Radar</span> Image of Sudan Collision Zone</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is a <span class="hlt">radar</span> image of a region in northern Sudan called the Keraf Suture that reveals newly discovered geologic features buried beneath layers of sand. This discovery is being used to guide field studies of the region and has opened up new perspectives on old problems, such as what controls the course of the Nile, a question that has perplexed geologists for centuries. The Nile is the yellowish/green line that runs from the top to the bottom of the image. A small town, Abu Dis, can be seen as the bright, white area on the east (right) bank of the Nile (about a third of the way down from the top) at the mouth of a dry stream valley or 'wadi' that drains into the river. Wadis flowing into the Nile from both east and west stand out as dark, reddish branch-like drainage patterns. The bright pink area on the west (left) side of the Nile is a region where rocks are exposed, but the area east (right) of the Nile is obscured by layers of sand, a few inches to several feet thick. Virtually everything visible on the right side of this <span class="hlt">radar</span> image is invisible when standing on the ground or when viewing photographs or satellite images such as the United States' Landsat or the French SPOT satellite. A sharp, straight fault cuts diagonally across the image, to the right of the Nile river. The area between the fault and the Nile is part of the collision zone where the ancient continents of East and West Gondwana crashed into each other to form the supercontinent Greater Gondwana more than 600 million years ago. On this image, the Nile approaches but never crosses the fault, indicating that this fault seems to be controlling the course of the Nile in this part of Sudan. The image is centered at 19.5 degrees north latitude, 33.35 degrees east longitude, and shows an area approximately 18 km by 20 km (10 miles by 12 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: Red is L-band, <span class="hlt">vertically</span> transmitted and <span class="hlt">vertically</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1375418-integrated-approach-monitoring-calibration-stability-operational-dual-polarization-radars','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1375418-integrated-approach-monitoring-calibration-stability-operational-dual-polarization-radars"><span>An integrated approach to monitoring the calibration stability of operational dual-polarization <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Vaccarono, Mattia; Bechini, Renzo; Chandrasekar, Chandra V.; ...</p> <p>2016-11-08</p> <p>The stability of weather <span class="hlt">radar</span> calibration is a mandatory aspect for quantitative applications, such as rainfall estimation, short-term weather prediction and initialization of numerical atmospheric and hydrological models. Over the years, calibration monitoring techniques based on external sources have been developed, specifically calibration using the Sun and calibration based on ground clutter returns. In this paper, these two techniques are integrated and complemented with a self-consistency procedure and an intercalibration technique. The aim of the integrated approach is to implement a robust method for online monitoring, able to detect significant changes in the <span class="hlt">radar</span> calibration. The physical consistency of polarimetricmore » <span class="hlt">radar</span> observables is exploited using the self-consistency approach, based on the expected correspondence between dual-polarization power and phase measurements in rain. This technique allows a reference absolute value to be provided for the <span class="hlt">radar</span> calibration, from which eventual deviations may be detected using the other procedures. In particular, the ground clutter calibration is implemented on both polarization channels (horizontal and <span class="hlt">vertical</span>) for each <span class="hlt">radar</span> scan, allowing the polarimetric variables to be monitored and hardware failures to promptly be recognized. The Sun calibration allows monitoring the calibration and sensitivity of the <span class="hlt">radar</span> receiver, in addition to the antenna <span class="hlt">pointing</span> accuracy. It is applied using observations collected during the standard operational scans but requires long integration times (several days) in order to accumulate a sufficient amount of useful data. Finally, an intercalibration technique is developed and performed to compare colocated measurements collected in rain by two <span class="hlt">radars</span> in overlapping regions. The integrated approach is performed on the C-band weather <span class="hlt">radar</span> network in northwestern Italy, during July–October 2014. The set of methods considered appears suitable to establish an online</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AnGeo..34..767S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AnGeo..34..767S"><span>Simultaneous observations of structure function parameter of refractive index using a high-resolution <span class="hlt">radar</span> and the DataHawk small airborne measurement system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scipión, Danny E.; Lawrence, Dale A.; Milla, Marco A.; Woodman, Ronald F.; Lume, Diego A.; Balsley, Ben B.</p> <p>2016-09-01</p> <p>The SOUSY (SOUnding SYstem) <span class="hlt">radar</span> was relocated to the Jicamarca Radio Observatory (JRO) near Lima, Peru, in 2000, where the <span class="hlt">radar</span> controller and acquisition system were upgraded with state-of-the-art parts to take full advantage of its potential for high-resolution atmospheric sounding. Due to its broad bandwidth (4 MHz), it is able to characterize clear-air backscattering with high range resolution (37.5 m). A campaign conducted at JRO in July 2014 aimed to characterize the lower troposphere with a high temporal resolution (8.1 Hz) using the DataHawk (DH) small unmanned aircraft system, which provides in situ atmospheric measurements at scales as small as 1 m in the lower troposphere and can be GPS-guided to obtain measurements within the beam of the <span class="hlt">radar</span>. This was a unique opportunity to make coincident observations by both systems and to directly compare their in situ and remotely sensed parameters. Because SOUSY only <span class="hlt">points</span> <span class="hlt">vertically</span>, it is only possible to retrieve <span class="hlt">vertical</span> <span class="hlt">radar</span> profiles caused by changes in the refractive index within the resolution volume. Turbulent variations due to scattering are described by the structure function parameter of refractive index Cn2. Profiles of Cn2 from the DH are obtained by combining pressure, temperature, and relative humidity measurements along the helical trajectory and integrated at the same scale as the <span class="hlt">radar</span> range resolution. Excellent agreement is observed between the Cn2 estimates obtained from the DH and SOUSY in the overlapping measurement regime from 1200 m up to 4200 m above sea level, and this correspondence provides the first accurate calibration of the SOUSY <span class="hlt">radar</span> for measuring Cn2.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPRS..130..385H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPRS..130..385H"><span><span class="hlt">Vertical</span> stratification of forest canopy for segmentation of understory trees within small-footprint airborne LiDAR <span class="hlt">point</span> clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hamraz, Hamid; Contreras, Marco A.; Zhang, Jun</p> <p>2017-08-01</p> <p>Airborne LiDAR <span class="hlt">point</span> cloud representing a forest contains 3D data, from which <span class="hlt">vertical</span> stand structure even of understory layers can be derived. This paper presents a tree segmentation approach for multi-story stands that stratifies the <span class="hlt">point</span> cloud to canopy layers and segments individual tree crowns within each layer using a digital surface model based tree segmentation method. The novelty of the approach is the stratification procedure that separates the <span class="hlt">point</span> cloud to an overstory and multiple understory tree canopy layers by analyzing <span class="hlt">vertical</span> distributions of LiDAR <span class="hlt">points</span> within overlapping locales. The procedure does not make a priori assumptions about the shape and size of the tree crowns and can, independent of the tree segmentation method, be utilized to <span class="hlt">vertically</span> stratify tree crowns of forest canopies. We applied the proposed approach to the University of Kentucky Robinson Forest - a natural deciduous forest with complex and highly variable terrain and vegetation structure. The segmentation results showed that using the stratification procedure strongly improved detecting understory trees (from 46% to 68%) at the cost of introducing a fair number of over-segmented understory trees (increased from 1% to 16%), while barely affecting the overall segmentation quality of overstory trees. Results of <span class="hlt">vertical</span> stratification of the canopy showed that the <span class="hlt">point</span> density of understory canopy layers were suboptimal for performing a reasonable tree segmentation, suggesting that acquiring denser LiDAR <span class="hlt">point</span> clouds would allow more improvements in segmenting understory trees. As shown by inspecting correlations of the results with forest structure, the segmentation approach is applicable to a variety of forest types.</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=20060039065&hterms=ambiguity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dambiguity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060039065&hterms=ambiguity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dambiguity"><span>SAR Ambiguity Study for the Cassini <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hensley, Scott; Im, Eastwood; Johnson, William T. K.</p> <p>1993-01-01</p> <p>The Cassini <span class="hlt">Radar</span>'s synthetic aperture <span class="hlt">radar</span> (SAR) ambiguity analysis is unique with respect to other spaceborne SAR ambiguity analyses owing to the non-orbiting spacecraft trajectory, asymmetric antenna pattern, and burst mode of data collection. By properly varying the <span class="hlt">pointing</span>, burst mode timing, and <span class="hlt">radar</span> parameters along the trajectory this study shows that the signal-to-ambiguity ratio of better than 15 dB can be achieved for all images obtained by the Cassini <span class="hlt">Radar</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01852&hterms=different+types+volcanoes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddifferent%2Btypes%2Bvolcanoes','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01852&hterms=different+types+volcanoes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Ddifferent%2Btypes%2Bvolcanoes"><span>Space <span class="hlt">Radar</span> Image of Pinacate Volcanic Field, Mexico</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This spaceborne <span class="hlt">radar</span> image shows the Pinacate Volcanic Field in the state of Sonora, Mexico, about 150 kilometers (93 miles) southeast of Yuma, Arizona. The United States/Mexico border runs across the upper right corner of the image. More than 300 volcanic vents occur in the Pinacate field, including cinder cones that experienced small eruptions as recently as 1934. The larger circular craters seen in the image are a type of volcano known as a 'maar', which erupts violently when rising magma encounters groundwater, producing highly pressurized steam that powers explosive eruptions. The highest elevations in the volcanic field, about 1200 meters (4000 feet), occur in the 'shield volcano' structure shown in bright white, occupying most of the left half of the image. Numerous cinder cones dot the flanks of the shield. The yellow patches to the right of center are newer, rough-textured lava flows that strongly reflect the long wavelength <span class="hlt">radar</span> signals. Along the left edge of the image are sand dunes of the Gran Desierto. The dark areas are smooth sand and the brighter brown and purple areas have vegetation on the surface. <span class="hlt">Radar</span> data provide a unique means to study the different types of lava flows and wind-blown sands. This image was acquired by Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour on April 18, 1994. The image is 57 kilometers by 48 kilometers (35 miles by 30 miles) and is centered at 31.7 degrees north latitude, 113.4 degrees West longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted, <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted, <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian, and United States space agencies, is part of NASA's Mission to Planet Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405974-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-oliktok-point-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405974-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-oliktok-point-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron, Elmendorf AFB, Alaska Risk Assessment, Oliktok <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-04-15</p> <p>This document contains the baseline human health risk assessment and the ecological risk assessment (ERA) for the Oliktok <span class="hlt">Point</span> Distant Early Warning (DEW) Line <span class="hlt">radar</span> installation. Eight sites at the Oliktok <span class="hlt">Point</span> <span class="hlt">radar</span> installation underwent remedial investigations (RIs) during the summer of 1993. The presence of chemical contamination in the soil, sediments, and surface water at the installation was evaluated and reported in the Oliktok <span class="hlt">Point</span> Remedial Investigation/Feasibility Study (RI/FS) (U.S. Air Force 1996). The analytical data reported in the RI/FS form the basis for the human health and ecological risk assessments. The primary chemicals of concern (COCs) at themore » eight sites are diesel and gasoline from past spills and/or leaks, chlorinated solvents, metals, and polychlorinated biphenyls (PCBs).« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110016436','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110016436"><span>Java <span class="hlt">Radar</span> Analysis Tool</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Zaczek, Mariusz P.</p> <p>2005-01-01</p> <p>Java <span class="hlt">Radar</span> Analysis Tool (JRAT) is a computer program for analyzing two-dimensional (2D) scatter plots derived from <span class="hlt">radar</span> returns showing pieces of the disintegrating Space Shuttle Columbia. JRAT can also be applied to similar plots representing <span class="hlt">radar</span> returns showing aviation accidents, and to scatter plots in general. The 2D scatter plots include overhead map views and side altitude views. The superposition of <span class="hlt">points</span> in these views makes searching difficult. JRAT enables three-dimensional (3D) viewing: by use of a mouse and keyboard, the user can rotate to any desired viewing angle. The 3D view can include overlaid trajectories and search footprints to enhance situational awareness in searching for pieces. JRAT also enables playback: time-tagged <span class="hlt">radar</span>-return data can be displayed in time order and an animated 3D model can be moved through the scene to show the locations of the Columbia (or other vehicle) at the times of the corresponding <span class="hlt">radar</span> events. The combination of overlays and playback enables the user to correlate a <span class="hlt">radar</span> return with a position of the vehicle to determine whether the return is valid. JRAT can optionally filter single <span class="hlt">radar</span> returns, enabling the user to selectively hide or highlight a desired <span class="hlt">radar</span> return.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01795.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01795.html"><span>Space <span class="hlt">Radar</span> Image of Florence, Italy</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This <span class="hlt">radar</span> image shows land use patterns in and around the city of Florence, Italy, shown here in the center of the image. Florence is situated on a plain in the Chianti Hill region of Central Italy. The Arno River flows through town and is visible as the dark line running from the upper right to the bottom center of the image. The city is home to some of the world's most famous art museums. The bridges seen crossing the Arno, shown as faint red lines in the upper right portion of the image, were all sacked during World War II with the exception of the Ponte Vecchio, which remains as Florence's only covered bridge. The large, black V-shaped feature near the center of the image is the Florence Railroad Station. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the Space Shuttle Endeavour on April 14, 1994. SIR-C/X-SAR, a joint mission of the German, Italian, and United States space agencies, is part of NASA's Mission to Planet Earth. This image is centered at 43.7 degrees north latitude and 11.15 degrees east longitude with North toward the upper left of the image. The area shown measures 20 kilometers by 17 kilometers (12.4 miles by 10.6 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, <span class="hlt">vertically</span> received; blue is C-band horizontally transmitted, <span class="hlt">vertically</span> received. http://photojournal.jpl.nasa.gov/catalog/PIA01795</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01815&hterms=image+alignment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dimage%2Balignment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01815&hterms=image+alignment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dimage%2Balignment"><span>Space <span class="hlt">Radar</span> Image of Washington D.C.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>The city of Washington, D.C., is shown is this space <span class="hlt">radar</span> image. Images like these are useful tools for urban planners and managers, who use them to map and monitor land use patterns. Downtown Washington is the bright area between the Potomac (upper center to lower left) and Anacostia (middle right) rivers. The dark cross shape that is formed by the National Mall, Tidal Basin, the White House and Ellipse is seen in the center of the image. Arlington National Cemetery is the dark blue area on the Virginia (left) side of the Potomac River near the center of the image. The Pentagon is visible in bright white and red, south of the cemetery. Due to the alignment of the <span class="hlt">radar</span> and the streets, the avenues that form the boundary between Washington and Maryland appear as bright red lines in the top, right and bottom parts of the image, parallel to the image borders. This image is centered at 38.85 degrees north latitude, 77.05 degrees west longitude. North is toward the upper right. The area shown is approximately 29 km by 26 km (18 miles by 16 miles). Colors are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: Red is the L-band horizontally transmitted, horizontally received; green is the L-band horizontally transmitted, <span class="hlt">vertically</span> received; blue is the C-band horizontally transmitted, <span class="hlt">vertically</span> received. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture (SIR-C/X-SAR) imaging <span class="hlt">radar</span> when it flew aboard the space shuttle Endeavour on April 18, 1994. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1353477','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1353477"><span>ARM Cloud <span class="hlt">Radar</span> Simulator Package for Global Climate Models Value-Added Product</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>Zhang, Yuying; Xie, Shaocheng</p> <p></p> <p>It has been challenging to directly compare U.S. Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) Climate Research Facility ground-based cloud <span class="hlt">radar</span> measurements with climate model output because of limitations or features of the observing processes and the spatial gap between model and the single-<span class="hlt">point</span> measurements. To facilitate the use of ARM <span class="hlt">radar</span> data in numerical models, an ARM cloud <span class="hlt">radar</span> simulator was developed to converts model data into pseudo-ARM cloud <span class="hlt">radar</span> observations that mimic the instrument view of a narrow atmospheric column (as compared to a large global climate model [GCM] grid-cell), thus allowing meaningful comparison between model outputmore » and ARM cloud observations. The ARM cloud <span class="hlt">radar</span> simulator value-added product (VAP) was developed based on the CloudSat simulator contained in the community satellite simulator package, the Cloud Feedback Model Intercomparison Project (CFMIP) Observation Simulator Package (COSP) (Bodas-Salcedo et al., 2011), which has been widely used in climate model evaluation with satellite data (Klein et al., 2013, Zhang et al., 2010). The essential part of the CloudSat simulator is the QuickBeam <span class="hlt">radar</span> simulator that is used to produce CloudSat-like <span class="hlt">radar</span> reflectivity, but is capable of simulating reflectivity for other <span class="hlt">radars</span> (Marchand et al., 2009; Haynes et al., 2007). Adapting QuickBeam to the ARM cloud <span class="hlt">radar</span> simulator within COSP required two primary changes: one was to set the frequency to 35 GHz for the ARM Ka-band cloud <span class="hlt">radar</span>, as opposed to 94 GHz used for the CloudSat W-band <span class="hlt">radar</span>, and the second was to invert the view from the ground to space so as to attenuate the beam correctly. In addition, the ARM cloud <span class="hlt">radar</span> simulator uses a finer <span class="hlt">vertical</span> resolution (100 m compared to 500 m for CloudSat) to resolve the more detailed structure of clouds captured by the ARM <span class="hlt">radars</span>. The ARM simulator has been developed following the COSP workflow (Figure 1) and using the capabilities available in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JGRD..12212747C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRD..12212747C"><span>GPM Satellite <span class="hlt">Radar</span> Measurements of Precipitation and Freezing Level in Atmospheric Rivers: Comparison With Ground-Based <span class="hlt">Radars</span> and Reanalyses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cannon, Forest; Ralph, F. Martin; Wilson, Anna M.; Lettenmaier, Dennis P.</p> <p>2017-12-01</p> <p>Atmospheric rivers (ARs) account for more than 90% of the total meridional water vapor flux in midlatitudes, and 25-50% of the annual precipitation in the coastal western United States. In this study, reflectivity profiles from the Global Precipitation Measurement Dual-Frequency Precipitation <span class="hlt">Radar</span> (GPM-DPR) are used to evaluate precipitation and temperature characteristics of ARs over the western coast of North America and the eastern North Pacific Ocean. Evaluation of GPM-DPR bright-band height using a network of ground-based <span class="hlt">vertically</span> <span class="hlt">pointing</span> <span class="hlt">radars</span> along the West Coast demonstrated exceptional agreement, and comparison with freezing level height from reanalyses over the eastern North Pacific Ocean also consistently agreed, indicating that GPM-DPR can be used to independently validate freezing level in models. However, precipitation comparison with gridded observations across the western United States indicated deficiencies in GPM-DPR's ability to reproduce the spatial distribution of winter precipitation, likely related to sampling frequency. Over the geographically homogeneous oceanic portion of the domain, sampling frequency was not problematic, and significant differences in the frequency and intensity of precipitation between GPM-DPR and reanalyses highlighted biases in both satellite-observed and modeled AR precipitation. Reanalyses precipitation rates below the minimum sensitivity of GPM-DPR accounted for a 20% increase in total precipitation, and 25% of <span class="hlt">radar</span>-derived precipitation rates were greater than the 99th percentile precipitation rate in reanalyses. Due to differences in the proportions of precipitation in convective, stratiform bright-band, and non-bright-band conditions, AR conditions contributed nearly 10% more to total precipitation in GPM-DPR than reanalyses.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AnGeo..36..577L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AnGeo..36..577L"><span>High-resolution <span class="hlt">vertical</span> velocities and their power spectrum observed with the MAARSY <span class="hlt">radar</span> - Part 1: frequency spectrum</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Qiang; Rapp, Markus; Stober, Gunter; Latteck, Ralph</p> <p>2018-04-01</p> <p>The Middle Atmosphere Alomar <span class="hlt">Radar</span> System (MAARSY) installed at the island of Andøya has been run for continuous probing of atmospheric winds in the upper troposphere and lower stratosphere (UTLS) region. In the current study, we present high-resolution wind measurements during the period between 2010 and 2013 with MAARSY. The spectral analysis applying the Lomb-Scargle periodogram method has been carried out to determine the frequency spectra of <span class="hlt">vertical</span> wind velocity. From a total of 522 days of observations, the statistics of the spectral slope have been derived and show a dependence on the background wind conditions. It is a general feature that the observed spectra of <span class="hlt">vertical</span> velocity during active periods (with wind velocity > 10 m s-1) are much steeper than during quiet periods (with wind velocity < 10 m s-1). The distribution of spectral slopes is roughly symmetric with a maximum at -5/3 during active periods, whereas a very asymmetric distribution with a maximum at around -1 is observed during quiet periods. The slope profiles along altitudes reveal a significant height dependence for both conditions, i.e., the spectra become shallower with increasing altitudes in the upper troposphere and maintain roughly a constant slope in the lower stratosphere. With both wind conditions considered together the general spectra are obtained and their slopes are compared with the background horizontal winds. The comparisons show that the observed spectra become steeper with increasing wind velocities under quiet conditions, approach a spectral slope of -5/3 at a wind velocity of 10 m s-1 and then roughly maintain this slope (-5/3) for even stronger winds. Our findings show an overall agreement with previous studies; furthermore, they provide a more complete climatology of frequency spectra of <span class="hlt">vertical</span> wind velocities under different wind conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150000267','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150000267"><span>Reduction of Non-uniform Beam Filling Effects by <span class="hlt">Vertical</span> Decorrelation: Theory and Simulations</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Short, David; Nakagawa, Katsuhiro; Iguchi, Toshio</p> <p>2013-01-01</p> <p>Algorithms for estimating precipitation rates from spaceborne <span class="hlt">radar</span> observations of apparent <span class="hlt">radar</span> reflectivity depend on attenuation correction procedures. The algorithm suite for the Ku-band precipitation <span class="hlt">radar</span> aboard the Tropical Rainfall Measuring Mission satellite is one such example. The well-known problem of nonuniform beam filling is a source of error in the estimates, especially in regions where intense deep convection occurs. The error is caused by unresolved horizontal variability in precipitation characteristics such as specific attenuation, rain rate, and effective reflectivity factor. This paper proposes the use of <span class="hlt">vertical</span> decorrelation for correcting the nonuniform beam filling error developed under the assumption of a perfect <span class="hlt">vertical</span> correlation. Empirical tests conducted using ground-based <span class="hlt">radar</span> observations in the current simulation study show that decorrelation effects are evident in tilted convective cells. However, the problem of obtaining reasonable estimates of a governing parameter from the satellite data remains unresolved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830005273','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830005273"><span>Aircraft and satellite measurement of ocean wave directional spectra using scanning-beam microwave <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jackson, F. C.; Walton, W. T.; Baker, P. L.</p> <p>1982-01-01</p> <p>A microwave <span class="hlt">radar</span> technique for remotely measuring the vector wave number spectrum of the ocean surface is described. The technique, which employs short-pulse, noncoherent <span class="hlt">radars</span> in a conical scan mode near <span class="hlt">vertical</span> incidence, is shown to be suitable for both aircraft and satellite application, the technique was validated at 10 km aircraft altitude, where we have found excellent agreement between buoy and <span class="hlt">radar</span>-inferred absolute wave height spectra.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/AD1001397','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/AD1001397"><span><span class="hlt">Radar</span> Detection Performance in Medium Grazing Angle X-band Sea-clutter</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2015-12-01</p> <p>polarisation HV: Horizontal transmit and <span class="hlt">Vertical</span> receive polarisation IRSG: Imagery <span class="hlt">Radar</span> Systems Group MAST06: Maritime Surveillance Trial 2006 PDF...different combinations of the polarisation, collection geometry and environmental conditions. Relevant models include the imaging <span class="hlt">radar</span> systems group (IRSG...atmospheric and system losses respectively and pulse compression adds a gain given by the pulse length - bandwidth product, TpB. The thermal noise power in the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01775&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01775&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia"><span>Space <span class="hlt">Radar</span> Image of Star City, Russia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image shows the Star City cosmonaut training center, east of Moscow, Russia. Four American astronauts are training here for future long-duration flights aboard the Russian Mir space station. These joint flights are giving NASA and the Russian Space Agency experience necessary for the construction of the international Alpha space station, beginning in late 1997. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR), on its 62nd orbit on October 3, 1994. This Star City image is centered at 55.55 degrees north latitude and 38.0 degrees east longitude. The area shown is approximately 32 kilometers by 49 kilometers (20 miles by 30 miles). North is to the top in this image. The <span class="hlt">radar</span> illumination is from the top of the image. The image was produced using three channels of SIR-C <span class="hlt">radar</span> data: red indicates L-band (23 cm wavelength, horizontally transmitted and received); green indicates L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue indicates C-band (6 cm wavelength, horizontally transmitted and <span class="hlt">vertically</span> received). In general, dark pink areas are agricultural; pink and light blue areas are urban communities; black areas represent lakes and rivers; dark blue areas are cleared forest; and light green areas are forested. The prominent black runways just right of center are Shchelkovo Airfield, about 4 km long. The textured pale blue-green area east and southeast of Shchelkovo Airfield is forest. Just east of the runways is a thin railroad line running southeast; the Star City compound lies just east of the small bend in the rail line. Star City contains the living quarters and training facilities for Russian cosmonauts and their families. Moscow's inner loop road is visible at the lower left edge of the image. The Kremlin is just off the left edge, on the banks of the meandering Moskva River. The Klyazma River snakes to the southeast from the reservoir in the upper left (shown in bright red</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01799.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01799.html"><span>Space <span class="hlt">Radar</span> Image of North Atlantic Ocean</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This is a <span class="hlt">radar</span> image showing surface features on the open ocean in the northeast Atlantic Ocean. There is no land mass in this image. The purple line in the lower left of the image is the stern wake of a ship. The ship creating the wake is the bright white spot on the middle, left side of the image. The ship's wake is about 28 kilometers (17 miles) long in this image and investigators believe that is because the ship may be discharging oil. The oil makes the wake last longer and causes it to stand out in this <span class="hlt">radar</span> image. A fairly sharp boundary or front extends from the lower left to the upper right corner of the image and separates two distinct water masses that have different temperatures. The different water temperature affects the wind patterns on the ocean. In this image, the light green area depicts rougher water with more wind, while the purple area is calmer water with less wind. The dark patches are smooth areas of low wind, probably related to clouds along the front, and the bright green patches are likely due to ice crystals in the clouds that scatter the <span class="hlt">radar</span> waves. The overall "fuzzy" look of this image is caused by long ocean waves, also called swells. Ocean <span class="hlt">radar</span> imagery allows the fine detail of ocean features and interactions to be seen, such as the wake, swell, ocean front and cloud effects, which can then be used to enhance the understanding of ocean dynamics on smaller and smaller scales. The image is centered at 42.8 degrees north latitude, 26.2 degrees west longitude and shows an area approximately 35 kilometers by 65 kilometers (22 by 40 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is C-band horizontally transmitted, horizontally received; blue is L-band <span class="hlt">vertically</span> transmitted, <span class="hlt">vertically</span> received. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01779.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01779.html"><span>Space <span class="hlt">Radar</span> Image of Teide Volcano</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This <span class="hlt">radar</span> image shows the Teide volcano on the island of Tenerife in the Canary Islands. The Canary Islands, part of Spain, are located in the eastern Atlantic Ocean off the coast of Morocco. Teide has erupted only once in the 20th Century, in 1909, but is considered a potentially threatening volcano due to its proximity to the city of Santa Cruz de Tenerife, shown in this image as the purple and white area on the lower right edge of the island. The summit crater of Teide, clearly visible in the left center of the image, contains lava flows of various ages and roughnesses that appear in shades of green and brown. Different vegetation zones, both natural and agricultural, are detected by the <span class="hlt">radar</span> as areas of purple, green and yellow on the volcano's flanks. Scientists are using images such as this to understand the evolution of the structure of Teide, especially the formation of the summit caldera and the potential for collapse of the flanks. The volcano is one of 15 identified by scientists as potentially hazardous to local populations, as part of the international The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 11, 1994. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered at 28.3 degrees North latitude and 16.6 degrees West longitude. North is toward the upper right. The area shown measures 90 kilometers by 54.5 kilometers (55.8 miles by 33.8 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, <span class="hlt">vertically</span> received; blue is C-band horizontally transmitted, <span class="hlt">vertically</span> received. http://photojournal.jpl.nasa.gov/catalog/PIA01779</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01796.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01796.html"><span>Space <span class="hlt">Radar</span> Image of Pishan, China</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This <span class="hlt">radar</span> image is centered near the small town of Pishan in northwest China, about 280 km (174 miles) southeast of the city of Kashgar along the ancient Silk Route in the Taklamakan desert of the Xinjiang Province. Geologists are using this <span class="hlt">radar</span> image as a map to study past climate changes and tectonics of the area. The irregular lavender branching patterns in the center of the image are the remains of ancient alluvial fans, gravel deposits that have accumulated at the base of the mountains during times of wetter climate. The subtle striped pattern cutting across the ancient fans are caused by thrusting of the Kun Lun Mountains north. This motion is caused by the continuing plate-tectonic collision of India with Asia. Modern fans show up as large lavender triangles above the ancient fan deposits. Yellow areas on the modern fans are vegetated oases. The gridded pattern results from the alignment of poplar trees that have been planted as wind breaks. The reservoir at the top of the image is part of a sophisticated irrigation system that supplies water to the oases. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour in April 1994. This image is centered at 37.4 degrees north latitude, 78.3 degrees east longitude and shows an area approximately 50 km by 100 km (31 miles by 62 miles). The colors are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: Red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, <span class="hlt">vertically</span> received; and blue is C-band horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian, and the United States space agencies, is part of NASA's Mission to Planet Earth. http://photojournal.jpl.nasa.gov/catalog/PIA01796</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01818&hterms=river+urban+city&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Driver%2Burban%2Bcity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01818&hterms=river+urban+city&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Driver%2Burban%2Bcity"><span>Space <span class="hlt">Radar</span> Image of Colorado River</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This space <span class="hlt">radar</span> image illustrates the recent rapid urban development occurring along the lower Colorado River at the Nevada/Arizona state line. Lake Mojave is the dark feature that occupies the river valley in the upper half of the image. The lake is actually a reservoir created behind Davis Dam, the bright white line spanning the river near the center of the image. The dam, completed in 1953, is used both for generating electric power and regulating the river's flow downstream. Straddling the river south of Davis Dam, shown in white and bright green, are the cities of Laughlin, Nevada (west of the river) and Bullhead City, Arizona (east of the river). The runway of the Laughlin, Bullhead City Airport is visible as a dark strip just east of Bullhead City. The area has experienced rapid growth associated with the gambling industry in Laughlin and on the Fort Mojave Indian Reservation to the south. The community of Riviera is the bright green area in a large bend of the river in the lower left part of the image. Complex drainage patterns and canyons are the dark lines seen throughout the image. <span class="hlt">Radar</span> is a useful tool for studying these patterns because of the instrument's sensitivity to roughness, vegetation and subtle topographic differences. This image is 50 kilometers by 35 kilometers (31 miles by 22 miles) and is centered at 35.25 degrees north latitude, 114.67 degrees west longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) on April 13, 1994, onboard the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Office of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16319123','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16319123"><span><span class="hlt">Radar</span> soundings of the ionosphere of Mars.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gurnett, D A; Kirchner, D L; Huff, R L; Morgan, D D; Persoon, A M; Averkamp, T F; Duru, F; Nielsen, E; Safaeinili, A; Plaut, J J; Picardi, G</p> <p>2005-12-23</p> <p>We report the first <span class="hlt">radar</span> soundings of the ionosphere of Mars with the MARSIS (Mars Advanced <span class="hlt">Radar</span> for Subsurface and Ionosphere Sounding) instrument on board the orbiting Mars Express spacecraft. Several types of ionospheric echoes are observed, ranging from <span class="hlt">vertical</span> echoes caused by specular reflection from the horizontally stratified ionosphere to a wide variety of oblique and diffuse echoes. The oblique echoes are believed to arise mainly from ionospheric structures associated with the complex crustal magnetic fields of Mars. Echoes at the electron plasma frequency and the cyclotron period also provide measurements of the local electron density and magnetic field strength.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405979-remedial-investigation-feasibility-study-point-lonely-radar-installation-alaska-volume-appendices-final-report-january-april','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405979-remedial-investigation-feasibility-study-point-lonely-radar-installation-alaska-volume-appendices-final-report-january-april"><span>Remedial investigation and feasibility study <span class="hlt">Point</span> Lonely <span class="hlt">Radar</span> Installation, Alaska. Volume 1. Appendices a - c. Final report, January 1995-April 1996</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>NONE</p> <p>1996-04-01</p> <p>This report presents the findings of Remedial Investigations and Feasibility Studies at sites located at the <span class="hlt">Point</span> Lonely <span class="hlt">radar</span> installation in northern Alaska. The sites were characterized based on sampling and analyses conducted during Remedial Investigation activities performed during August and September 1993.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017GeoRL..44.7519R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017GeoRL..44.7519R"><span>A case study of microphysical structures and hydrometeor phase in convection using <span class="hlt">radar</span> Doppler spectra at Darwin, Australia</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riihimaki, L. D.; Comstock, J. M.; Luke, E.; Thorsen, T. J.; Fu, Q.</p> <p>2017-07-01</p> <p>To understand the microphysical processes that impact diabatic heating and cloud lifetimes in convection, we need to characterize the spatial distribution of supercooled liquid water. To address this observational challenge, ground-based <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors at the Darwin Atmospheric Radiation Measurement site are used to classify cloud phase within a deep convective cloud. The cloud cannot be fully observed by a lidar due to signal attenuation. Therefore, we developed an objective method for identifying hydrometeor classes, including mixed-phase conditions, using k-means clustering on parameters that describe the shape of the Doppler spectra from <span class="hlt">vertically</span> <span class="hlt">pointing</span> Ka-band cloud <span class="hlt">radar</span>. This approach shows that multiple, overlapping mixed-phase layers exist within the cloud, rather than a single region of supercooled liquid. Diffusional growth calculations show that the conditions for the Wegener-Bergeron-Findeisen process exist within one of these mixed-phase microstructures.</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://www.osti.gov/pages/biblio/1402422-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1402422-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia"><span>A case study of microphysical structures and hydrometeor phase in convection using <span class="hlt">radar</span> Doppler spectra at Darwin, Australia</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Riihimaki, Laura D.; Comstock, J. M.; Luke, E.; ...</p> <p>2017-07-12</p> <p>To understand the microphysical processes that impact diabatic heating and cloud lifetimes in convection, we need to characterize the spatial distribution of supercooled liquid water. To address this observational challenge, ground-based <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors at the Darwin Atmospheric Radiation Measurement site are used to classify cloud phase within a deep convective cloud. The cloud cannot be fully observed by a lidar due to signal attenuation. Therefore, we developed an objective method for identifying hydrometeor classes, including mixed-phase conditions, using k-means clustering on parameters that describe the shape of the Doppler spectra from <span class="hlt">vertically</span> <span class="hlt">pointing</span> Ka-band cloud <span class="hlt">radar</span>. Furthermore, thismore » approach shows that multiple, overlapping mixed-phase layers exist within the cloud, rather than a single region of supercooled liquid. Diffusional growth calculations show that the conditions for the Wegener-Bergeron-Findeisen process exist within one of these mixed-phase microstructures.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1402422-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1402422-case-study-microphysical-structures-hydrometeor-phase-convection-using-radar-doppler-spectra-darwin-australia"><span>A case study of microphysical structures and hydrometeor phase in convection using <span class="hlt">radar</span> Doppler spectra at Darwin, Australia</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>Riihimaki, Laura D.; Comstock, J. M.; Luke, E.</p> <p></p> <p>To understand the microphysical processes that impact diabatic heating and cloud lifetimes in convection, we need to characterize the spatial distribution of supercooled liquid water. To address this observational challenge, ground-based <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors at the Darwin Atmospheric Radiation Measurement site are used to classify cloud phase within a deep convective cloud. The cloud cannot be fully observed by a lidar due to signal attenuation. Therefore, we developed an objective method for identifying hydrometeor classes, including mixed-phase conditions, using k-means clustering on parameters that describe the shape of the Doppler spectra from <span class="hlt">vertically</span> <span class="hlt">pointing</span> Ka-band cloud <span class="hlt">radar</span>. Furthermore, thismore » approach shows that multiple, overlapping mixed-phase layers exist within the cloud, rather than a single region of supercooled liquid. Diffusional growth calculations show that the conditions for the Wegener-Bergeron-Findeisen process exist within one of these mixed-phase microstructures.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910031150&hterms=space+mapping&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dspace%2Bmapping','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910031150&hterms=space+mapping&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dspace%2Bmapping"><span>Tropical rain mapping <span class="hlt">radar</span> on the Space Station</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Im, Eastwood; Li, Fuk</p> <p>1989-01-01</p> <p>The conceptual design for a tropical rain mapping <span class="hlt">radar</span> for flight on the manned Space Station is discussed. In this design the <span class="hlt">radar</span> utilizes a narrow, dual-frequency (9.7 GHz and 24.1 GHz) beam, electronically scanned antenna to achieve high spatial (4 km) and <span class="hlt">vertical</span> (250 m) resolutions and a relatively large (800 km) cross-track swath. An adaptive scan strategy will be used for better utilization of <span class="hlt">radar</span> energy and dwell time. Such a system can detect precipitation at rates of up to 100 mm/hr with accuracies of roughly 15 percent. With the proposed space-time sampling strategy, the monthly averaged rainfall rate can be estimated to within 8 percent, which is essential for many climatological studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20120009530&hterms=remote+viewing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dremote%2Bviewing','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20120009530&hterms=remote+viewing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dremote%2Bviewing"><span>Characteristics of Deep Tropical and Subtropical Convection from Nadir-Viewing High-Altitude Airborne Doppler <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Heymsfield, Gerald M.; Tian, Lin; Heymsfield, Andrew J.; Li, Lihua; Guimond, Stephen</p> <p>2010-01-01</p> <p>This paper presents observations of deep convection characteristics in the tropics and subtropics that have been classified into four categories: tropical cyclone, oceanic, land, and sea breeze. <span class="hlt">Vertical</span> velocities in the convection were derived from Doppler <span class="hlt">radar</span> measurements collected during several NASA field experiments from the nadir-viewing high-altitude ER-2 Doppler <span class="hlt">radar</span> (EDOP). Emphasis is placed on the <span class="hlt">vertical</span> structure of the convection from the surface to cloud top (sometimes reaching 18-km altitude). This unique look at convection is not possible from other approaches such as ground-based or lower-altitude airborne scanning <span class="hlt">radars</span>. The <span class="hlt">vertical</span> motions from the <span class="hlt">radar</span> measurements are derived using new relationships between <span class="hlt">radar</span> reflectivity and hydrometeor fall speed. Various convective properties, such as the peak updraft and downdraft velocities and their corresponding altitude, heights of reflectivity levels, and widths of reflectivity cores, are estimated. The most significant findings are the following: 1) strong updrafts that mostly exceed 15 m/s, with a few exceeding 30 m/s, are found in all the deep convection cases, whether over land or ocean; 2) peak updrafts were almost always above the 10-km level and, in the case of tropical cyclones, were closer to the 12-km level; and 3) land-based and sea-breeze convection had higher reflectivities and wider convective cores than oceanic and tropical cyclone convection. In addition, the high-resolution EDOP data were used to examine the connection between reflectivity and <span class="hlt">vertical</span> velocity, for which only weak linear relationships were found. The results are discussed in terms of dynamical and microphysical implications for numerical models and future remote sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005Icar..177...32N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005Icar..177...32N"><span><span class="hlt">Radar</span> imaging of Saturn'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>Nicholson, Philip D.; French, Richard G.; Campbell, Donald B.; Margot, Jean-Luc; Nolan, Michael C.; Black, Gregory J.; Salo, Heikki J.</p> <p>2005-09-01</p> <p>We present delay-Doppler images of Saturn's rings based on <span class="hlt">radar</span> observations made at Arecibo Observatory between 1999 and 2003, at a wavelength of 12.6 cm and at ring opening angles of 20.1°⩽|B|⩽26.7°. The average <span class="hlt">radar</span> cross-section of the A ring is ˜77% relative to that of the B ring, while a stringent upper limit of 3% is placed on the cross-section of the C ring and 9% on that of the Cassini Division. These results are consistent with those obtained by Ostro et al. [1982, Icarus 49, 367-381] from <span class="hlt">radar</span> observations at |B|=21.4°, but provide higher resolution maps of the rings' reflectivity profile. The average cross-section of the A and B rings, normalized by their projected unblocked area, is found to have decreased from 1.25±0.31 to 0.74±0.19 as the rings have opened up, while the circular polarization ratio has increased from 0.64±0.06 to 0.77±0.06. The steep decrease in cross-section is at variance with previous <span class="hlt">radar</span> measurements [Ostro et al., 1980, Icarus 41, 381-388], and neither this nor the polarization variations are easily understood within the framework of either classical, many-particle-thick or monolayer ring models. One possible explanation involves <span class="hlt">vertical</span> size segregation in the rings, whereby observations at larger elevation angles which see deeper into the rings preferentially see the larger particles concentrated near the rings' mid-plane. These larger particles may be less reflective and/or rougher and thus more depolarizing than the smaller ones. Images from all four years show a strong m=2 azimuthal asymmetry in the reflectivity of the A ring, with an amplitude of ±20% and minima at longitudes of 67±4° and 247±4° from the sub-Earth <span class="hlt">point</span>. We attribute the asymmetry to the presence of gravitational wakes in the A ring as invoked by Colombo et al. [1976, Nature 264, 344-345] to explain the similar asymmetry long seen at optical wavelengths. A simple radiative transfer model suggests that the enhancement of the azimuthal</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AtmRe.196..200Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AtmRe.196..200Z"><span>Cross-evaluation of reflectivity from the space-borne precipitation <span class="hlt">radar</span> and multi-type ground-based weather <span class="hlt">radar</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>Zhong, Lingzhi; Yang, Rongfang; Wen, Yixin; Chen, Lin; Gou, Yabin; Li, Ruiyi; Zhou, Qing; Hong, Yang</p> <p>2017-11-01</p> <p>China operational weather <span class="hlt">radar</span> network consists of more than 200 ground-based <span class="hlt">radars</span> (GR(s)). The lack of unified calibrators often result in poor mosaic products as well as its limitation in <span class="hlt">radar</span> data assimilation in numerical models. In this study, <span class="hlt">radar</span> reflectivity and precipitation <span class="hlt">vertical</span> structures observed from space-borne TRMM (Tropical Rainfall Measurement Mission) PR (precipitation <span class="hlt">radar</span>) and GRs are volumetrically matched and cross-evaluated. It is found that observation of GRs is basically consistent with that of PR. For their overlapping scanning regions, the GRs are often affected by the beam blockage for complex terrain. The statistics show the better agreement among S band A type (SA) <span class="hlt">radars</span>, S band B type (SB) <span class="hlt">radars</span> and PR, as well as poor performance of S band C type (SC) <span class="hlt">radars</span>. The reflectivity offsets between GRs and PR depend on the reflectivity magnitudes: They are positive for weak precipitation and negative for middle and heavy precipitation, respectively. Although the GRs are quite consistent with PR for large sample, an individual GR has its own fluctuated biases monthly. When the sample number is small, the bias statistics may be determined by a single bad GR in a group. Results from this study shed lights that the space-borne precipitation <span class="hlt">radars</span> could be used to quantitatively calibrate systematic bias existing in different GRs in order to improve the consistency of ground-based weather <span class="hlt">radar</span> network across China, and also bears the promise to provide a robust reference even form a space and ground constellation network for the dual-frequency precipitation <span class="hlt">radars</span> onboard the satellites anticipated in the near future.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01859&hterms=wine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dwine','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01859&hterms=wine&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dwine"><span>Space <span class="hlt">Radar</span> Image of Rhine River, France and Germany</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This spaceborne <span class="hlt">radar</span> image shows a segment of the Rhine River where it forms the border between the Alsace region of northeastern France on the left and the Black Forest region of Germany on the right. The Rhine, one of the largest and most used waterways in central Europe, winds its way through five countries from the Swiss-Austrian Alps to the North Sea coast of the Netherlands. The river valley is densely populated, as seen in this image, which shows the French city of Strasbourg, the light blue and orange area in the upper left center; and the German cities of Kehl, across the river from Strasbourg and Offenburg, the bright area in right center. The fertile valley is famous for its wine production and most of the agricultural areas in the image, shown in purple patches, are vineyards. The light green areas are forest. Scientists can use <span class="hlt">radar</span> images like this one to monitor the effects of urban and agricultural development on sensitive ecosystems such as the Rhine River valley. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour on October 2, 1994. The image is 34.2 kilometers by 33.2 kilometers (21.2 miles by 20.6 miles) and is centered at 48.5 degrees north latitude, 7.7 degrees east longitude. North is toward the upper left. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted, <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted, <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/23704','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/23704"><span>Imaging tree roots with borehole <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>John R. Butnor; Kurt H. Johnsen; Per Wikstrom; Tomas Lundmark; Sune Linder</p> <p>2006-01-01</p> <p>Ground-penetrating <span class="hlt">radar</span> has been used to de-tect and map tree roots using surface-based antennas in reflection mode. On amenable soils these methods can accurately detect lateral tree roots. In some tree species (e.g. Pinus taeda, Pinus palustris), <span class="hlt">vertically</span> orientated tap roots directly beneath the tree, comprise most of the root mass. It is...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..1611666N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..1611666N"><span><span class="hlt">Radar</span>-based rainfall estimation: Improving Z/R relations through comparison of drop size distributions, rainfall rates and <span class="hlt">radar</span> reflectivity patterns</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Neuper, Malte; Ehret, Uwe</p> <p>2014-05-01</p> <p>The relation between the measured <span class="hlt">radar</span> reflectivity factor Z and surface rainfall intensity R - the Z/R relation - is profoundly complex, so that in general one speaks about <span class="hlt">radar</span>-based quantitative precipitation estimation (QPE) rather than exact measurement. Like in Plato's Allegory of the Cave, what we observe in the end is only the 'shadow' of the true rainfall field through a very small backscatter of an electromagnetic signal emitted by the <span class="hlt">radar</span>, which we hope has been actually reflected by hydrometeors. The meteorological relevant and valuable Information is gained only indirectly by more or less justified assumptions. One of these assumptions concerns the drop size distribution, through which the rain intensity is finally associated with the measured <span class="hlt">radar</span> reflectivity factor Z. The real drop size distribution is however subject to large spatial and temporal variability, and consequently so is the true Z/R relation. Better knowledge of the true spatio-temporal Z/R structure therefore has the potential to improve <span class="hlt">radar</span>-based QPE compared to the common practice of applying a single or a few standard Z/R relations. To this end, we use observations from six laser-optic disdrometers, two <span class="hlt">vertically</span> <span class="hlt">pointing</span> micro rain <span class="hlt">radars</span>, 205 rain gauges, one rawindsonde station and two C-band Doppler <span class="hlt">radars</span> installed or operated in and near the Attert catchment (Luxembourg). The C-band <span class="hlt">radars</span> and the rawindsonde station are operated by the Belgian and German Weather Services, the rain gauge data was partly provided by the French, Dutch, Belgian, German Weather Services and the Ministry of Agriculture of Luxembourg and the other equipment was installed as part of the interdisciplinary DFG research project CAOS (Catchment as Organized Systems). With the various data sets correlation analyzes were executed. In order to get a notion on the different appearance of the reflectivity patterns in the <span class="hlt">radar</span> image, first of all various simple distribution indices (for example the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140002246','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140002246"><span>Momentum Flux Determination Using the Multi-beam Poker Flat Incoherent Scatter <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nicolls, M. J.; Fritts, D. C.; Janches, Diego; Heinselman, C. J.</p> <p>2012-01-01</p> <p>In this paper, we develop an estimator for the <span class="hlt">vertical</span> flux of horizontal momentum with arbitrary beam <span class="hlt">pointing</span>, applicable to the case of arbitrary but fixed beam <span class="hlt">pointing</span> with systems such as the Poker Flat Incoherent Scatter <span class="hlt">Radar</span> (PFISR). This method uses information from all available beams to resolve the variances of the wind field in addition to the <span class="hlt">vertical</span> flux of both meridional and zonal momentum, targeted for high-frequency wave motions. The estimator utilises the full covariance of the distributed measurements, which provides a significant reduction in errors over the direct extension of previously developed techniques and allows for the calculation of an error covariance matrix of the estimated quantities. We find that for the PFISR experiment, we can construct an unbiased and robust estimator of the momentum flux if sufficient and proper beam orientations are chosen, which can in the future be optimized for the expected frequency distribution of momentum-containing scales. However, there is a potential trade-off between biases and standard errors introduced with the new approach, which must be taken into account when assessing the momentum fluxes. We apply the estimator to PFISR measurements on 23 April 2008 and 21 December 2007, from 60-85 km altitude, and show expected results as compared to mean winds and in relation to the measured <span class="hlt">vertical</span> velocity variances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080025047&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbiomass%2Bforest','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080025047&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dbiomass%2Bforest"><span>Forest Attributes from <span class="hlt">Radar</span> Interferometric Structure and its Fusion with Optical Remote Sensing</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Treuhaft, Robert N.; Law, Beverly E.; Asner, Gregory P.</p> <p>2004-01-01</p> <p>The possibility of global, three-dimensional remote sensing of forest structure with interferometric synthetic aperture <span class="hlt">radar</span> (InSAR) bears on important forest ecological processes, particularly the carbon cycle. InSAR supplements two-dimensional remote sensing with information in the <span class="hlt">vertical</span> dimension. Its strengths in potential for global coverage complement those of lidar (light detecting and ranging), which has the potential for high-accuracy <span class="hlt">vertical</span> profiles over small areas. InSAR derives its sensitivity to forest <span class="hlt">vertical</span> structure from the differences in signals received by two, spatially separate <span class="hlt">radar</span> receivers. Estimation of parameters describing <span class="hlt">vertical</span> structure requires multiple-polarization, multiple-frequency, or multiple-baseline InSAR. Combining InSAR with complementary remote sensing techniques, such as hyperspectral optical imaging and lidar, can enhance <span class="hlt">vertical</span>-structure estimates and consequent biophysical quantities of importance to ecologists, such as biomass. Future InSAR experiments will supplement recent airborne and spaceborne demonstrations, and together with inputs from ecologists regarding structure, they will suggest designs for future spaceborne strategies for measuring global vegetation structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..556..922P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..556..922P"><span>Spatial variability of extreme rainfall at <span class="hlt">radar</span> subpixel scale</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Peleg, Nadav; Marra, Francesco; Fatichi, Simone; Paschalis, Athanasios; Molnar, Peter; Burlando, Paolo</p> <p>2018-01-01</p> <p>Extreme rainfall is quantified in engineering practice using Intensity-Duration-Frequency curves (IDF) that are traditionally derived from rain-gauges and more recently also from remote sensing instruments, such as weather <span class="hlt">radars</span>. These instruments measure rainfall at different spatial scales: rain-gauge samples rainfall at the <span class="hlt">point</span> scale while weather <span class="hlt">radar</span> averages precipitation on a relatively large area, generally around 1 km2. As such, a <span class="hlt">radar</span> derived IDF curve is representative of the mean areal rainfall over a given <span class="hlt">radar</span> pixel and neglects the within-pixel rainfall variability. In this study, we quantify subpixel variability of extreme rainfall by using a novel space-time rainfall generator (STREAP model) that downscales in space the rainfall within a given <span class="hlt">radar</span> pixel. The study was conducted using a unique <span class="hlt">radar</span> data record (23 years) and a very dense rain-gauge network in the Eastern Mediterranean area (northern Israel). <span class="hlt">Radar</span>-IDF curves, together with an ensemble of <span class="hlt">point</span>-based IDF curves representing the <span class="hlt">radar</span> subpixel extreme rainfall variability, were developed fitting Generalized Extreme Value (GEV) distributions to annual rainfall maxima. It was found that the mean areal extreme rainfall derived from the <span class="hlt">radar</span> underestimate most of the extreme values computed for <span class="hlt">point</span> locations within the <span class="hlt">radar</span> pixel (on average, ∼70%). The subpixel variability of rainfall extreme was found to increase with longer return periods and shorter durations (e.g. from a maximum variability of 10% for a return period of 2 years and a duration of 4 h to 30% for 50 years return period and 20 min duration). For the longer return periods, a considerable enhancement of extreme rainfall variability was found when stochastic (natural) climate variability was taken into account. Bounding the range of the subpixel extreme rainfall derived from <span class="hlt">radar</span>-IDF can be of major importance for different applications that require very local estimates of rainfall extremes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1256528-path-towards-uncertainty-assignment-operational-cloud-phase-algorithm-from-arm-vertically-pointing-active-sensors','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1256528-path-towards-uncertainty-assignment-operational-cloud-phase-algorithm-from-arm-vertically-pointing-active-sensors"><span>A path towards uncertainty assignment in an operational cloud-phase algorithm from ARM <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Riihimaki, Laura D.; Comstock, Jennifer M.; Anderson, Kevin K.; ...</p> <p>2016-06-10</p> <p>Knowledge of cloud phase (liquid, ice, mixed, etc.) is necessary to describe the radiative impact of clouds and their lifetimes, but is a property that is difficult to simulate correctly in climate models. One step towards improving those simulations is to make observations of cloud phase with sufficient accuracy to help constrain model representations of cloud processes. In this study, we outline a methodology using a basic Bayesian classifier to estimate the probabilities of cloud-phase class from Atmospheric Radiation Measurement (ARM) <span class="hlt">vertically</span> <span class="hlt">pointing</span> active remote sensors. The advantage of this method over previous ones is that it provides uncertainty informationmore » on the phase classification. We also test the value of including higher moments of the cloud <span class="hlt">radar</span> Doppler spectrum than are traditionally used operationally. Using training data of known phase from the Mixed-Phase Arctic Cloud Experiment (M-PACE) field campaign, we demonstrate a proof of concept for how the method can be used to train an algorithm that identifies ice, liquid, mixed phase, and snow. Over 95 % of data are identified correctly for pure ice and liquid cases used in this study. Mixed-phase and snow cases are more problematic to identify correctly. When lidar data are not available, including additional information from the Doppler spectrum provides substantial improvement to the algorithm. As a result, this is a first step towards an operational algorithm and can be expanded to include additional categories such as drizzle with additional training data.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ASCMO...2...49R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ASCMO...2...49R"><span>A path towards uncertainty assignment in an operational cloud-phase algorithm from ARM <span class="hlt">vertically</span> <span class="hlt">pointing</span> active sensors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Riihimaki, Laura D.; Comstock, Jennifer M.; Anderson, Kevin K.; Holmes, Aimee; Luke, Edward</p> <p>2016-06-01</p> <p>Knowledge of cloud phase (liquid, ice, mixed, etc.) is necessary to describe the radiative impact of clouds and their lifetimes, but is a property that is difficult to simulate correctly in climate models. One step towards improving those simulations is to make observations of cloud phase with sufficient accuracy to help constrain model representations of cloud processes. In this study, we outline a methodology using a basic Bayesian classifier to estimate the probabilities of cloud-phase class from Atmospheric Radiation Measurement (ARM) <span class="hlt">vertically</span> <span class="hlt">pointing</span> active remote sensors. The advantage of this method over previous ones is that it provides uncertainty information on the phase classification. We also test the value of including higher moments of the cloud <span class="hlt">radar</span> Doppler spectrum than are traditionally used operationally. Using training data of known phase from the Mixed-Phase Arctic Cloud Experiment (M-PACE) field campaign, we demonstrate a proof of concept for how the method can be used to train an algorithm that identifies ice, liquid, mixed phase, and snow. Over 95 % of data are identified correctly for pure ice and liquid cases used in this study. Mixed-phase and snow cases are more problematic to identify correctly. When lidar data are not available, including additional information from the Doppler spectrum provides substantial improvement to the algorithm. This is a first step towards an operational algorithm and can be expanded to include additional categories such as drizzle with additional training data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090032643&hterms=bateman&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dbateman','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090032643&hterms=bateman&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dbateman"><span>Developing Lightning Prediction Tools for the CCAFS Dual-Polarimetric <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Petersen, W. A.; Carey, L. D.; Deierling, W.; Johnson, E.; Bateman, M.</p> <p>2009-01-01</p> <p>NASA Marshall Space Flight Center and the University of Alabama Huntsville are collaborating with the 45th Weather Squadron (45WS) to develop improved lightning prediction capabilities for the new C-band dual-polarimetric weather <span class="hlt">radar</span> being acquired for use by 45WS and launch weather forecasters at Cape Canaveral Air Force Station (CCAFS). In particular, these algorithms will focus on lightning onset, cessation and combined lightning-<span class="hlt">radar</span> applications for convective winds assessment. Research using <span class="hlt">radar</span> reflectivity (Z) data for prediction of lightning onset has been extensively discussed in the literature and subsequently applied by launch weather forecasters as it pertains to lightning nowcasting. Currently the forecasters apply a relatively straight forward but effective temperature-Z threshold algorithm for assessing the likelihood of lightning onset in a given storm. In addition, a layered VIL above the freezing level product is used as automated guidance for the onset of lightning. Only limited research and field work has been conducted on lightning cessation using Z and <span class="hlt">vertically</span>-integrated Z for determining cessation. Though not used operationally <span class="hlt">vertically</span>-integrated Z (basis for VIL) has recently shown promise as a tool for use in nowcasting lightning cessation. The work discussed herein leverages and expands upon these and similar reflectivity-threshold approaches via the application/addition of over two decades of polarimetric <span class="hlt">radar</span> research focused on distinct multi-parameter <span class="hlt">radar</span> signatures of ice/mixed-phase initiation and ice-crystal orientation in highly electrified convective clouds. Specifically, our approach is based on numerous previous studies that have observed repeatable patterns in the behavior of the <span class="hlt">vertical</span> hydrometeor column as it relates to the temporal evolution of differential reflectivity and depolarization (manifested in either LDR or p(sub hv)), development of in-situ mixed and ice phase microphysics, electric fields, and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01829&hterms=parliament&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dparliament','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01829&hterms=parliament&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dparliament"><span>Space <span class="hlt">Radar</span> Image of Canberra, Australia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>Australia's capital city, Canberra, is shown in the center of this spaceborne <span class="hlt">radar</span> image. Images like this can help urban planners assess land use patterns. Heavily developed areas appear in bright patchwork patterns of orange, yellow and blue. Dense vegetation appears bright green, while cleared areas appear in dark blue or black. Located in southeastern Australia, the site of Canberra was selected as the capital in 1901 as a geographic compromise between Sydney and Melbourne. Design and construction of the city began in 1908 under the supervision of American architect Walter Burley-Griffin. Lake Burley-Griffin is located above and to the left of the center of the image. The bright pink area is the Parliament House. The city streets, lined with government buildings, radiate like spokes from the Parliament House. The bright purple cross in the lower left corner of the image is a reflection from one of the large dish-shaped radio antennas at the Tidbinbilla, Canberra Deep Space Network Communication Complex, operated jointly by NASA and the Australian Space Office. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) on April 10, 1994, onboard the space shuttle Endeavour. The image is 28 kilometers by 25 kilometers (17 miles by 15 miles) and is centered at 35.35 degrees south latitude, 149.17 degrees east longitude. North is toward the upper left. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian, and United States space agencies, is part of NASA's Office of Mission to Planet Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730017144','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730017144"><span>Venus wind-altitude <span class="hlt">radar</span> study</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Goldfischer, L. I.</p> <p>1973-01-01</p> <p>A study was made of a wind/altitude <span class="hlt">radar</span> for a Venus probe. The baseline configuration was taken to be the SKD-2100 Doppler <span class="hlt">radar</span> modified to accommodate the altimeter portion of the APN-187 and a single beam antenna. Using current models of the Venus environment, engineering studies were made to define design requirements and to estimate the operational and physical characteristics of the conceptual design. The results of the study are that: (1) the <span class="hlt">radar</span> instrument should have an altitude limit of at least 34 km for velocity and at least 17.5 km for altitude measurement, (2) <span class="hlt">vertical</span> accuracy should be better than + or - 0.9 percent and horizontal velocity accuracy should be better than + or - 3 percent over the operating altitude range, and (3) altimeter accuracy should be within + or - 3 percent up to about 2.5 km and should improve over the remainder of the altimeter operating range. The <span class="hlt">radar</span> is expected to require between 48.5 and 69.3 watts of power and to weigh between 3.86 and 5.21 kg (8.5 and 11.5 lb). In each case, if power could be supplied directly from the probe batteries the lower figures would apply; the upper figures would apply if a power conditioner must be used.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ak0486.photos.193537p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ak0486.photos.193537p/"><span>52. View from ground level showing lower <span class="hlt">radar</span> scanner switch ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>52. View from ground level showing lower <span class="hlt">radar</span> scanner switch with open port door in <span class="hlt">radar</span> scanner building 105 showing emanating waveguides from lower switch in <span class="hlt">vertical</span> run; photograph also shows catwalk to upper scanner switch in upper left side of photograph and structural supports. - Clear Air Force Station, Ballistic Missile Early Warning System Site II, One mile west of mile marker 293.5 on Parks Highway, 5 miles southwest of Anderson, Anderson, Denali Borough, AK</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01794.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01794.html"><span>Space <span class="hlt">Radar</span> Image of Great Wall of China</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>These <span class="hlt">radar</span> images show two segments of the Great Wall of China in a desert region of north-central China, about 700 kilometers (434 miles) west of Beijing. The wall appears as a thin orange band, running from the top to the bottom of the left image, and from the middle upper-left to the lower-right of the right image. These segments of the Great Wall were constructed in the 15th century, during the Ming Dynasty. The wall is between 5 and 8 meters high (16 to 26 feet) in these areas. The entire wall is about 3,000 kilometers (1,864 miles) long and about 150 kilometers (93 miles) of the wall appear in these two images. The wall is easily detected from space by <span class="hlt">radar</span> because its steep, smooth sides provide a prominent surface for reflection of the <span class="hlt">radar</span> beam. Near the center of the left image, two dry lake beds have been developed for salt extraction. Rectangular patterns in both images indicate agricultural development, primarily wheat fields. The images were acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour on April 10, 1994. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The left image is centered at 37.7 degrees North latitude and 107.5 degrees East longitude. The right image is centered at 37.5 degrees North latitude and 108.1 degrees East longitude. North is toward the upper right. Each area shown measures 25 kilometers by 75 kilometers (15.5 miles by 45.5 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is L-band horizontally transmitted, <span class="hlt">vertically</span> received; blue is C-band horizontally transmitted, <span class="hlt">vertically</span> received. http://photojournal.jpl.nasa.gov/catalog/PIA01794</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000057028','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000057028"><span>A System Concept for the Advanced Post-TRMM Rainfall Profiling <span class="hlt">Radars</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Im, Eastwood; Smith, Eric A.</p> <p>1998-01-01</p> <p>Atmospheric latent heating field is fundamental to all modes of atmospheric circulation and upper mixed layer circulations of the ocean. The key to understanding the atmospheric heating process is understanding how and where precipitation occurs. The principal atmospheric processes which link precipitation to atmospheric circulation include: (1) convective mass fluxes in the form of updrafts and downdrafts; (2) microphysical. nucleation and growth of hydrometeors; and (3) latent heating through dynamical controls on the gravitation-driven <span class="hlt">vertical</span> mass flux of precipitation. It is well-known that surface and near-surface rainfall are two of the key forcing functions on a number of geophysical parameters at the surface-air interface. Over ocean, rainfall variation contributes to the redistribution of water salinity, sea surface temperature, fresh water supply, and marine biology and eco-system. Over land, rainfall plays a significant role in rainforest ecology and chemistry, land hydrology and surface runoff. Precipitation has also been closely linked to a number of atmospheric anomalies and natural hazards that occur at various time scales, including hurricanes, cyclones, tropical depressions, flash floods, droughts, and most noticeable of all, the El Ninos. From this <span class="hlt">point</span> of view, the significance of global atmospheric precipitation has gone far beyond the science arena - it has a far-reaching impact on human's socio-economic well-being and sustenance. These and many other science applications require the knowledge of, in a global basis, the <span class="hlt">vertical</span> rain structures, including <span class="hlt">vertical</span> motion, rain intensity, differentiation of the precipitating hydrometeors' phase state, and the classification of mesoscale physical structure of the rain systems. The only direct means to obtain such information is the use of a spaceborne profiling <span class="hlt">radar</span>. It is important to mention that the Tropical Rainfall Measuring Mission (TRMM) have made a great stride forward towards this</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/1991dsp..conf...28M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1991dsp..conf...28M"><span>Hardware description ADSP-21020 40-bit floating <span class="hlt">point</span> DSP as designed in a remotely controlled digital CW Doppler <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Morrison, R. E.; Robinson, S. H.</p> <p></p> <p>A continuous wave Doppler <span class="hlt">radar</span> system has been designed which is portable, easily deployed, and remotely controlled. The heart of this system is a DSP/control board using Analog Devices ADSP-21020 40-bit floating <span class="hlt">point</span> digital signal processor (DSP) microprocessor. Two 18-bit audio A/D converters provide digital input to the DSP/controller board for near real time target detection. Program memory for the DSP is dual ported with an Intel 87C51 microcontroller allowing DSP code to be up-loaded or down-loaded from a central controlling computer. The 87C51 provides overall system control for the remote <span class="hlt">radar</span> and includes a time-of-day/day-of-year real time clock, system identification (ID) switches, and input/output (I/O) expansion by an Intel 82C55 I/O expander.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts068-s-052.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts068-s-052.html"><span>STS-68 <span class="hlt">radar</span> image: Mt. Rainier, Washington</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1994-10-01</p> <p>STS068-S-052 (3 October 1994) --- This is a <span class="hlt">radar</span> image of Mount Rainier in Washington state. The volcano last erupted about 150 years ago and numerous large floods and debris flows have originated on its slopes during the last century. Today the volcano is heavily mantled with glaciers and snow fields. More than 100,000 people live on young volcanic mud flows less than 10,000 years old and, are within the range of future, devastating mud slides. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the Space Shuttle Endeavour on its 20th orbit on October 1, 1994. The area shown in the image is approximately 59 by 60 kilometers (36.5 by 37 miles). North is toward the top left of the image, which was composed by assigning red and green colors to the L-Band, horizontally transmitted and <span class="hlt">vertically</span>, and the L-Band, horizontally transmitted and <span class="hlt">vertically</span> received. Blue indicates the C-Band, horizontally transmitted and <span class="hlt">vertically</span> received. In addition to highlighting topographic slopes facing the Space Shuttle, SIR-C records rugged areas as brighter and smooth areas as darker. The scene was illuminated by the Shuttle's <span class="hlt">radar</span> from the northwest so that northwest-facing slopes are brighter and southeast-facing slopes are dark. Forested regions are pale green in color, clear cuts and bare ground are bluish or purple; ice is dark green and white. The round cone at the center of the image is the 14,435 feet (4,399 meters) active volcano, Mount Rainier. On the lower slopes is a zone of rock ridges and rubble (purple to reddish) above coniferous forests (in yellow/green). The western boundary of Mount Rainier National Park is seen as a transition from protected, old-growth forest to heavily logged private land, a mosaic of recent clear cuts (bright purple/blue) and partially re-grown timber plantations (pale blue). The prominent river seen curving away from the mountain at the top of the image (to the northwest) is the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015OcDyn..65..679R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015OcDyn..65..679R"><span>Comparison of HF <span class="hlt">radar</span> measurements with Eulerian and Lagrangian surface currents</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Röhrs, Johannes; Sperrevik, Ann Kristin; Christensen, Kai Håkon; Broström, Göran; Breivik, Øyvind</p> <p>2015-05-01</p> <p>High-frequency (HF) <span class="hlt">radar</span>-derived ocean currents are compared with in situ measurements to conclude if the <span class="hlt">radar</span> observations include effects of surface waves that are of second order in the wave amplitude. Eulerian current measurements from a high-resolution acoustic Doppler current profiler and Lagrangian measurements from surface drifters are used as references. Directional wave spectra are obtained from a combination of pressure sensor data and a wave model. Our analysis shows that the wave-induced Stokes drift is not included in the HF <span class="hlt">radar</span>-derived currents, that is, HF <span class="hlt">radars</span> measure the Eulerian current. A disputed nonlinear correction to the phase velocity of surface gravity waves, which may affect HF <span class="hlt">radar</span> signals, has a magnitude of about half the Stokes drift at the surface. In our case, this contribution by nonlinear dispersion would be smaller than the accuracy of the HF <span class="hlt">radar</span> currents, hence no conclusion can be made. Finally, the analysis confirms that the HF <span class="hlt">radar</span> data represent an exponentially weighted <span class="hlt">vertical</span> average where the decay scale is proportional to the wavelength of the transmitted signal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405959-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-decision-document-further-response-action-planned-bullen-point-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405959-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-decision-document-further-response-action-planned-bullen-point-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron Elmendorf AFB, Alaska. Decision document for no further response action planned: Bullen <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-05-24</p> <p>This Decision Document discusses the selection of no further action as the recommended action for two sites located at the Bullen <span class="hlt">Point</span> <span class="hlt">radar</span> installation. The United States Air Force (Air Force) completed a Remedial Investigation/Feasibility Study and a Risk Assessment for the five sites located at the Bullen <span class="hlt">Point</span> installation (U.S. Air Force 1996a,b). Based on the findings of these activities, two sites are recommended for no further action. Sites at the Bullen <span class="hlt">Point</span> <span class="hlt">radar</span> installation recommended for no further action are: Old Landfill/Dump Site East (LF06) and Drum Storage Area (SS10).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995JApMe..34.1978S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995JApMe..34.1978S"><span>Climatological Characterization of Three-Dimensional Storm Structure from Operational <span class="hlt">Radar</span> and Rain Gauge Data.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Steiner, Matthias; Houze, Robert A., Jr.; Yuter, Sandra E.</p> <p>1995-09-01</p> <p>Three algorithms extract information on precipitation type, structure, and amount from operational <span class="hlt">radar</span> and rain gauge data. Tests on one month of data from one site show that the algorithms perform accurately and provide products that characterize the essential features of the precipitation climatology. Input to the algorithms are the operationally executed volume scans of a <span class="hlt">radar</span> and the data from a surrounding rain gauge network. The algorithms separate the <span class="hlt">radar</span> echoes into convective and stratiform regions, statistically summarize the <span class="hlt">vertical</span> structure of the <span class="hlt">radar</span> echoes, and determine precipitation rates and amounts on high spatial resolution.The convective and stratiform regions are separated on the basis of the intensity and sharpness of the peaks of echo intensity. The peaks indicate the centers of the convective region. Precipitation not identified as convective is stratiform. This method avoids the problem of underestimating the stratiform precipitation. The separation criteria are applied in exactly the same way throughout the observational domain and the product generated by the algorithm can be compared directly to model output. An independent test of the algorithm on data for which high-resolution dual-Doppler observations are available shows that the convective stratiform separation algorithm is consistent with the physical definitions of convective and stratiform precipitation.The <span class="hlt">vertical</span> structure algorithm presents the frequency distribution of <span class="hlt">radar</span> reflectivity as a function of height and thus summarizes in a single plot the <span class="hlt">vertical</span> structure of all the <span class="hlt">radar</span> echoes observed during a month (or any other time period). Separate plots reveal the essential differences in structure between the convective and stratiform echoes.Tests yield similar results (within less than 10%) for monthly rain statistics regardless of the technique used for estimating the precipitation, as long as the <span class="hlt">radar</span> reflectivity values are adjusted to agree with monthly</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA114708','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA114708"><span>Coordinated <span class="hlt">Radar</span> and Aircraft Observations of Turbulence.</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1981-05-26</p> <p>VELOCITY (il/) Jig. 10. Spectrum at two <span class="hlt">points</span> having excessive <span class="hlt">radar</span> c / 23 ACKNOWLEDGMENr The direction and support of Mr. 1. Goldman of the FAA...of Doppler Weather <span class="hlt">Radar</span> to Turbulence Measure- ments Which Affect Aircraft," FAA Report RD-77-145 (March 1977). 2. R. T. Strauch, "Applications of...Meteorological Doppler <span class="hlt">Radar</span> for Weather- Surveillance Near Air Terminals", IEEE Trans. Geosci. Electron., G15-17, 4 (1979). 3. P. B. MacCready</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16.8593V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16.8593V"><span>A Wing Pod-based Millimeter Wave Cloud <span class="hlt">Radar</span> on HIAPER</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vivekanandan, Jothiram; Tsai, Peisang; Ellis, Scott; Loew, Eric; Lee, Wen-Chau; Emmett, Joanthan</p> <p>2014-05-01</p> <p>One of the attractive features of a millimeter wave <span class="hlt">radar</span> system is its ability to detect micron-sized particles that constitute clouds with lower than 0.1 g m-3 liquid or ice water content. Scanning or <span class="hlt">vertically-pointing</span> ground-based millimeter wavelength <span class="hlt">radars</span> are used to study stratocumulus (Vali et al. 1998; Kollias and Albrecht 2000) and fair-weather cumulus (Kollias et al. 2001). Airborne millimeter wavelength <span class="hlt">radars</span> have been used for atmospheric remote sensing since the early 1990s (Pazmany et al. 1995). Airborne millimeter wavelength <span class="hlt">radar</span> systems, such as the University of Wyoming King Air Cloud <span class="hlt">Radar</span> (WCR) and the NASA ER-2 Cloud <span class="hlt">Radar</span> System (CRS), have added mobility to observe clouds in remote regions and over oceans. Scientific requirements of millimeter wavelength <span class="hlt">radar</span> are mainly driven by climate and cloud initiation studies. Survey results from the cloud <span class="hlt">radar</span> user community indicated a common preference for a narrow beam W-band <span class="hlt">radar</span> with polarimetric and Doppler capabilities for airborne remote sensing of clouds. For detecting small amounts of liquid and ice, it is desired to have -30 dBZ sensitivity at a 10 km range. Additional desired capabilities included a second wavelength and/or dual-Doppler winds. Modern <span class="hlt">radar</span> technology offers various options (e.g., dual-polarization and dual-wavelength). Even though a basic fixed beam Doppler <span class="hlt">radar</span> system with a sensitivity of -30 dBZ at 10 km is capable of satisfying cloud detection requirements, the above-mentioned additional options, namely dual-wavelength, and dual-polarization, significantly extend the measurement capabilities to further reduce any uncertainty in <span class="hlt">radar</span>-based retrievals of cloud properties. This paper describes a novel, airborne pod-based millimeter wave <span class="hlt">radar</span>, preliminary <span class="hlt">radar</span> measurements and corresponding derived scientific products. Since some of the primary engineering requirements of this millimeter wave <span class="hlt">radar</span> are that it should be deployable on an airborne platform</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01726.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01726.html"><span>Space <span class="hlt">Radar</span> Image of the Silk Route in Niya, Taklamak, China</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This composite image is of an area thought to contain the ruins of the ancient settlement of Niya. It is located in the southwest corner of the Taklamakan Desert in China Sinjiang Province. This region was part of some of China's earliest dynasties and from the third century BC on was traversed by the famous Silk Road. The Silk Road, passing east-west through this image, was an ancient trade route that led across Central Asia's desert to Persia, Byzantium and Rome. The multi-frequency, multi-polarized <span class="hlt">radar</span> imagery was acquired on orbit 106 of the space shuttle Endeavour on April 16, 1994 by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span>. The image is centered at 37.78 degrees north latitude and 82.41 degrees east longitude. The area shown is approximately 35 kilometers by 83 kilometers (22 miles by 51 miles). The image is a composite of an image from an Earth-orbiting satellite called Systeme Probatoire d'Observation de la Terre (SPOT) and a SIR-C multi-frequency, multi-polarized <span class="hlt">radar</span> image. The false-color <span class="hlt">radar</span> image was created by displaying the C-band (horizontally transmitted and received) return in red, the L-band (horizontally transmitted and received) return in green, and the L-band (horizontally transmitted and <span class="hlt">vertically</span> received) return in blue. The prominent east/west pink formation at the bottom of the image is most likely a ridge of loosely consolidated sedimentary rock. The Niya River -- the black feature in the lower right of the French satellite image -- meanders north-northeast until it clears the sedimentary ridge, at which <span class="hlt">point</span> it abruptly turns northwest. Sediment and evaporite deposits left by the river over millennia dominate the center and upper right of the <span class="hlt">radar</span> image (in light pink). High ground, ridges and dunes are seen among the riverbed meanderings as mottled blue. Through image enhancement and analysis, a new feature probably representing a man-made canal has been discovered and mapped. http</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996JAtS...53.1887B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996JAtS...53.1887B"><span><span class="hlt">Vertical</span> Motion Characteristics of Tropical Cyclones Determined with Airborne Doppler Radial Velocities.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Black, Micheal L.; Burpee, Robert W.; Marks, Frank D., Jr.</p> <p>1996-07-01</p> <p><span class="hlt">Vertical</span> motions in seven Atlantic hurricanes are determined from data recorded by Doppler <span class="hlt">radars</span> on research aircraft. The database consists of Doppler velocities and reflectivities from <span class="hlt">vertically</span> <span class="hlt">pointing</span> <span class="hlt">radar</span> rays collected along radial flight legs through the hurricane centers. The <span class="hlt">vertical</span> motions are estimated throughout the depth of the troposphere from the Doppler velocities and bulk estimates of particle fallspeeds.Portions of the flight tracks are subjectively divided into eyewall, rainband, stratiform, and `other' regions. Characteristics of the <span class="hlt">vertical</span> velocity and <span class="hlt">radar</span> structure are described as a function of altitude for the entire dataset and each of the four regions. In all of the regions, more than 70% of the <span class="hlt">vertical</span> velocities range from 2 to 2 m s1. The broadest distribution of <span class="hlt">vertical</span> motion is in the eyewall region where 5% of the <span class="hlt">vertical</span> motions are >5 m s1. Averaged over the entire dataset, the mean <span class="hlt">vertical</span> velocity is upward at all altitudes. Mean downward motion occurs only in the lower troposphere of the stratiform region. Significant <span class="hlt">vertical</span> variations in the mean profiles of <span class="hlt">vertical</span> velocity and reflectivity are discussed and related to microphysical processes.In the lower and middle troposphere, the characteristics of the Doppler-derived <span class="hlt">vertical</span> motions are similar to those described in an earlier study using flight-level <span class="hlt">vertical</span> velocities, even though the horizontal resolution of the Doppler data is 750 m compared to 125 m from the in situ flight-level measurements. The Doppler data are available at higher altitudes than those reached by turboprop aircraft and provide information on <span class="hlt">vertical</span> as well as horizontal variations. In a <span class="hlt">vertical</span> plane along the radial flight tracks, Doppler up- and downdrafts are defined at each 300-m altitude interval as <span class="hlt">vertical</span> velocities whose absolute values continuously exceed 1.5 m s1, with at least one speed having an absolute value greater than 3.0 m s1. The properties of the Doppler</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870001040','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870001040"><span>UHF and VHF <span class="hlt">radar</span> observations of thunderstorms</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Holden, D. N.; Ulbrich, C. W.; Larsen, M. F.; Rottger, J.; Ierkic, H. M.; Swartz, W.</p> <p>1986-01-01</p> <p>A study of thunderstorms was made in the Summer of 1985 with the 430-MHz and 50-MHz <span class="hlt">radars</span> at the Arecibo Observatory in Puerto Rico. Both <span class="hlt">radars</span> use the 300-meter dish, which gives a beam width of less than 2 degrees even at these long wavelengths. Though the <span class="hlt">radars</span> are steerable, only <span class="hlt">vertical</span> beams were used in this experiment. The height resolution was 300 and 150 meters for the UHF and VHF, respectively. Lightning echoes, as well as returns from precipitation and clear-air turbulence were detected with both wavelengths. Large increases in the returned power were found to be coincident with increasing downward <span class="hlt">vertical</span> velocities at UHF, whereas at VHF the total power returned was relatively constant during the life of a storm. This was attributed to the fact that the VHF is more sensitive to scattering from the turbulence-induced inhomogeneities in the refractive index and less sensitive to scatter from precipitation particles. On occasion, the shape of the Doppler spectra was observed to change with the occurrence of a lightning discharge in the pulse volume. Though the total power and mean reflectivity weighted Doppler velocity changed little during these events, the power is Doppler frequency bins near that corresponding to the updraft did increase substantially within a fraction of a second after a discharge was detected in the beam. This suggests some interaction between precipitation and lightning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01746.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01746.html"><span>Space <span class="hlt">Radar</span> Image of Mammoth Mountain, California</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This false-color composite <span class="hlt">radar</span> image of the Mammoth Mountain area in the Sierra Nevada Mountains, California, was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> aboard the space shuttle Endeavour on its 67th orbit on October 3, 1994. The image is centered at 37.6 degrees north latitude and 119.0 degrees west longitude. The area is about 39 kilometers by 51 kilometers (24 miles by 31 miles). North is toward the bottom, about 45 degrees to the right. In this image, red was created using L-band (horizontally transmitted/<span class="hlt">vertically</span> received) polarization data; green was created using C-band (horizontally transmitted/<span class="hlt">vertically</span> received) polarization data; and blue was created using C-band (horizontally transmitted and received) polarization data. Crawley Lake appears dark at the center left of the image, just above or south of Long Valley. The Mammoth Mountain ski area is visible at the top right of the scene. The red areas correspond to forests, the dark blue areas are bare surfaces and the green areas are short vegetation, mainly brush. The purple areas at the higher elevations in the upper part of the scene are discontinuous patches of snow cover from a September 28 storm. New, very thin snow was falling before and during the second space shuttle pass. In parallel with the operational SIR-C data processing, an experimental effort is being conducted to test SAR data processing using the Jet Propulsion Laboratory's massively parallel supercomputing facility, centered around the Cray Research T3D. These experiments will assess the abilities of large supercomputers to produce high throughput Synthetic Aperture <span class="hlt">Radar</span> processing in preparation for upcoming data-intensive SAR missions. The image released here was produced as part of this experimental effort. http://photojournal.jpl.nasa.gov/catalog/PIA01746</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2000SPIE.4052..363L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2000SPIE.4052..363L"><span><span class="hlt">Radar</span> volume reflectivity estimation using an array of ground-based rainfall drop size detectors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lane, John; Merceret, Francis; Kasparis, Takis; Roy, D.; Muller, Brad; Jones, W. Linwood</p> <p>2000-08-01</p> <p>Rainfall drop size distribution (DSD) measurements made by single disdrometers at isolated ground sites have traditionally been used to estimate the transformation between weather <span class="hlt">radar</span> reflectivity Z and rainfall rate R. Despite the immense disparity in sampling geometries, the resulting Z-R relation obtained by these single <span class="hlt">point</span> measurements has historically been important in the study of applied <span class="hlt">radar</span> meteorology. Simultaneous DSD measurements made at several ground sites within a microscale area may be used to improve the estimate of <span class="hlt">radar</span> reflectivity in the air volume surrounding the disdrometer array. By applying the equations of motion for non-interacting hydrometers, a volume estimate of Z is obtained from the array of ground based disdrometers by first calculating a 3D drop size distribution. The 3D-DSD model assumes that only gravity and terminal velocity due to atmospheric drag within the sampling volume influence hydrometer dynamics. The sampling volume is characterized by wind velocities, which are input parameters to the 3D-DSD model, composed of <span class="hlt">vertical</span> and horizontal components. Reflectivity data from four consecutive WSR-88D volume scans, acquired during a thunderstorm near Melbourne, FL on June 1, 1997, are compared to data processed using the 3D-DSD model and data form three ground based disdrometers of a microscale array.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010cosp...38.1175C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010cosp...38.1175C"><span>Validation of COSMIC radio occultation electron density profiles by incoherent scatter <span class="hlt">radar</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>Cherniak, Iurii; Zakharenkova, Irina</p> <p></p> <p> majority of the ionospheric parameters -density and kinetic temperature of electron and main ions, the plasma drift velocity and others. The comparison of RO reveals that usually COSMIC RO profiles are in a rather good agreement with ISR profiles both in the F2 layer peak electron density (NmF2) and the form of profiles. The coincidence of profiles is better in the cases when projection of the ray path of tangent <span class="hlt">points</span> is closer to the ISR location. It is necessary to note that retrieved electron density profiles should not be interpreted as actual <span class="hlt">vertical</span> profiles. The geographical location of the ray path tangent <span class="hlt">points</span> at the top and at the bottom of a profile may differ by several hundred kilometers. So the spatial smearing of data takes place and RO technique represents an image of <span class="hlt">vertical</span> and horizontal ionospheric structure. That is why the comparison with ground-based data has rather relative character. We derived quantitative parameters to char-acterize the differences of the compared profiles: the peak height difference, the relative peak density difference. Most of the compared profiles agree within error limits, depending on the accuracy of the occultation-and the <span class="hlt">radar</span>-derived profiles. In general COSMIC RO profiles are in a good agreement with incoherent <span class="hlt">radar</span> profiles both in the F2 layer peak electron density (NmF2) and the form of the profiles. The coincidence of COSMIC and incoherent <span class="hlt">radar</span> pro-files is better in the cases when projection of the ray path tangent <span class="hlt">points</span> is closer to the <span class="hlt">radar</span> location. COSMIC measurements can be efficiently used to study the topside part of the iono-spheric electron density. To validate the reliability of the COSMIC ionospheric observations it must be done the big work on the analysis and statistical generalization of the huge data array (today the total number of ionospheric occultation is more than 2.300.000), but this technique is a very promising one to retrieve accurate profiles of the ionospheric electron density</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998SPIE.3462...52P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998SPIE.3462...52P"><span><span class="hlt">Radar</span> images analysis for scattering surfaces characterization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Piazza, Enrico</p> <p>1998-10-01</p> <p>According to the different problems and techniques related to the detection and recognition of airplanes and vehicles moving on the Airport surface, the present work mainly deals with the processing of images gathered by a high-resolution <span class="hlt">radar</span> sensor. The <span class="hlt">radar</span> images used to test the investigated algorithms are relative to sequence of images obtained in some field experiments carried out by the Electronic Engineering Department of the University of Florence. The <span class="hlt">radar</span> is the Ka band <span class="hlt">radar</span> operating in the'Leonardo da Vinci' Airport in Fiumicino (Rome). The images obtained from the <span class="hlt">radar</span> scan converter are digitized and putted in x, y, (pixel) co- ordinates. For a correct matching of the images, these are corrected in true geometrical co-ordinates (meters) on the basis of fixed <span class="hlt">points</span> on an airport map. Correlating the airplane 2-D multipoint template with actual <span class="hlt">radar</span> images, the value of the signal in the <span class="hlt">points</span> involved in the template can be extracted. Results for a lot of observation show a typical response for the main section of the fuselage and the wings. For the fuselage, the back-scattered echo is low at the prow, became larger near the center on the aircraft and than it decrease again toward the tail. For the wings the signal is growing with a pretty regular slope from the fuselage to the tips, where the signal is the strongest.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.G53A0762W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.G53A0762W"><span>Integration of Space-borne SAR and Ground-Based <span class="hlt">Radar</span> for 3D Deformation Mapping of the Central Calaveras Fault at Coyote Dam</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Werner, C. L.; Baker, B.; Milillo, P.; Magnard, C.; Strozzi, T.; Wegmüller, U.</p> <p>2017-12-01</p> <p>The Central Calaveras Fault (CCF) passes directly through Coyote Dam located southeast of Morgan Hill, California. This earthen embankment dam owned and operated by the Santa Clara Valley Water District (District), has experienced over 80 cm of accumulated fault creep since its construction in 1936. The average slip rate is 10 to 15 mm/year as measured using surveying, GPS, and more recently, terrestrial <span class="hlt">radar</span> interferometry (TRI). The CCF is a right-lateral strike-slip fault that has the potential for a M7.25 earthquake resulting in meter scale displacement. In 2015, the District initiated a geological analysis of the CCF integrating past surveying, GPS data, TRI deformation mapping, paleoseismic trenching, and boreholes. The initial TRI survey included dam measurements from two locations, imaging the upstream and downstream embankments over the period from February to July 2015. The TRI data from the downstream embankment data showed a complex deformation pattern not consistent with a strike-slip fault model. A second measurement campaign was initiated utilizing multiple <span class="hlt">radar</span> viewpoints with the aim of resolving the 3D deformation field of the downstream embankment. The campaign occurred between May and November 2016 and showed an unexpected strong westward and downward movement exceeding 2 cm/year (see Figure). TRI data were acquired from 4 separate observation <span class="hlt">points</span> every 2 to 4 weeks during this campaign. <span class="hlt">Point</span> target analysis methods were used to avoid contamination of the deformation data by vegetation and <span class="hlt">radar</span> shadow. Deformation uncertainty in the downstream fault zone was relatively high due to the nearly coplanar arrangement of the TRI observation <span class="hlt">points</span>. To better constrain the <span class="hlt">vertical</span> deformation, in this report we integrate spaceborne measurements from the Cosmo-SkyMed (CS) <span class="hlt">radar</span> satellite in the 3D deformation solution. The LOS to the satellite has a large <span class="hlt">vertical</span> component not present in the TRI measurement geometry that facilitates the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992PhDT.......186H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992PhDT.......186H"><span><span class="hlt">Radar</span> Polarimetry: Theory, Analysis, and Applications</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hubbert, John Clark</p> <p></p> <p>The fields of <span class="hlt">radar</span> polarimetry and optical polarimetry are compared. The mathematics of optic polarimetry are formulated such that a local right handed coordinate system is always used to describe the polarization states. This is not done in <span class="hlt">radar</span> polarimetry. <span class="hlt">Radar</span> optimum polarization theory is redeveloped within the framework of optical polarimetry. The <span class="hlt">radar</span> optimum polarizations and optic eigenvalues of common scatterers are compared. In addition a novel definition of an eigenpolarization state is given and the accompanying mathematics is developed. The polarization response calculated using optic, <span class="hlt">radar</span> and novel definitions is presented for a variety of scatterers. Polarimetric transformation provides a means to characterize scatters in more than one polarization basis. Polarimetric transformation for an ensemble of scatters is obtained via two methods: (1) the covariance method and (2) the instantaneous scattering matrix (ISM) method. The covariance method is used to relate the mean <span class="hlt">radar</span> parameters of a +/-45^circ linear polarization basis to those of a horizontal and <span class="hlt">vertical</span> polarization basis. In contrast the ISM method transforms the individual time samples. Algorithms are developed for transforming the time series from fully polarimetric <span class="hlt">radars</span> that switch between orthogonal states. The transformed time series are then used to calculate the mean <span class="hlt">radar</span> parameters of interest. It is also shown that propagation effects do not need to be removed from the ISM's before transformation. The techniques are demonstrated using data collected by POLDIRAD, the German Aerospace Research Establishment's fully polarimetric C-band <span class="hlt">radar</span>. The differential phase observed between two copolar states, Psi_{CO}, is composed of two phases: (1) differential propagation phase, phi_{DP}, and (2) differential backscatter phase, delta. The slope of phi_{DP } with range is an estimate of the specific differential phase, K_{DP}. The process of estimating K_{DP} is complicated when</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950038689&hterms=glacier+melt&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dglacier%2Bmelt','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950038689&hterms=glacier+melt&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dglacier%2Bmelt"><span><span class="hlt">Radar</span> measurements of melt zones on the Greenland Ice Sheet</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jezek, Kenneth C.; Gogineni, Prasad; Shanableh, M.</p> <p>1994-01-01</p> <p>Surface-based microwave <span class="hlt">radar</span> measurements were performed at a location on the western flank of the Greenland Ice Sheet. Here, firn metamorphasis is dominated by seasonal melt, which leads to marked contrasts in the <span class="hlt">vertical</span> structure of winter and summer firn. This snow regime is also one of the brightest <span class="hlt">radar</span> targets on Earth with an average backscatter coefficient of 0 dB at 5.3 GHz and an incidence angle of 25 deg. By combining detailed observations of firn physical properties with ranging <span class="hlt">radar</span> measurements we find that the glaciological mechanism associated with this strong electromagnetic response is summer ice lens formation within the previous winter's snow pack. This observation has important implications for monitoring and understanding changes in ice sheet volume using spaceborne microwave sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01783.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01783.html"><span>Space <span class="hlt">Radar</span> Image of Houston, Texas</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This image of Houston, Texas, shows the amount of detail that is possible to obtain using spaceborne <span class="hlt">radar</span> imaging. Images such as this -- obtained by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) flying aboard the space shuttle Endeavor last fall -- can become an effective tool for urban planners who map and monitor land use patterns in urban, agricultural and wetland areas. Central Houston appears pink and white in the upper portion of the image, outlined and crisscrossed by freeways. The image was obtained on October 10, 1994, during the space shuttle's 167th orbit. The area shown is 100 kilometers by 60 kilometers (62 miles by 38 miles) and is centered at 29.38 degrees north latitude, 95.1 degrees west longitude. North is toward the upper left. The pink areas designate urban development while the green-and blue-patterned areas are agricultural fields. Black areas are bodies of water, including Galveston Bay along the right edge and the Gulf of Mexico at the bottom of the image. Interstate 45 runs from top to bottom through the image. The narrow island at the bottom of the image is Galveston Island, with the city of Galveston at its northeast (right) end. The dark cross in the upper center of the image is Hobby Airport. Ellington Air Force Base is visible below Hobby on the other side of Interstate 45. Clear Lake is the dark body of water in the middle right of the image. The green square just north of Clear Lake is Johnson Space Center, home of Mission Control and the astronaut training facilities. The black rectangle with a white center that appears to the left of the city center is the Houston Astrodome. The colors in this image were obtained using the follow <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted, <span class="hlt">vertically</span> received); green represents the C-band (horizontally transmitted, <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and received). http</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AMT....10.1739N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AMT....10.1739N"><span>Wind turbine impact on operational weather <span class="hlt">radar</span> I/Q data: characterisation and filtering</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Norin, Lars</p> <p>2017-05-01</p> <p>For the past 2 decades wind turbines have been growing in number all over the world as a response to the increasing demand for renewable energy. However, the rapid expansion of wind turbines presents a problem for many <span class="hlt">radar</span> systems, including weather <span class="hlt">radars</span>. Wind turbines in the line of sight of a weather <span class="hlt">radar</span> can have a negative impact on the <span class="hlt">radar</span>'s measurements. As weather <span class="hlt">radars</span> are important instruments for meteorological offices, finding a way for wind turbines and weather <span class="hlt">radars</span> to co-exist would be of great societal value.Doppler weather <span class="hlt">radars</span> base their measurements on in-phase and quadrature phase (I/Q) data. In this work a month's worth of recordings of high-resolution I/Q data from an operational Swedish C-band weather <span class="hlt">radar</span> are presented. The impact of <span class="hlt">point</span> targets, such as masts and wind turbines, on the I/Q data is analysed and characterised. It is shown that the impact of <span class="hlt">point</span> targets on single <span class="hlt">radar</span> pulses, when normalised by amplitude, is manifested as a distinct and highly repeatable signature. The shape of this signature is found to be independent of the size, shape and yaw angle of the wind turbine. It is further demonstrated how the robustness of the <span class="hlt">point</span> target signature can be used to identify and filter out the impact of wind turbines in the <span class="hlt">radar</span>'s signal processor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19..488D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19..488D"><span>Rain/snow <span class="hlt">radar</span> remote sensing with two X-band <span class="hlt">radars</span> operating over an altitude gradient in the French Alps</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Delrieu, Guy; Cazenave, Frédéric; Yu, Nan; Boudevillain, Brice; Faure, Dominique; Gaussiat, Nicolas</p> <p>2017-04-01</p> <p>Operating weather <span class="hlt">radars</span> in high-mountain regions faces the following well-known dilemma: (1) installing <span class="hlt">radar</span> on top of mountains allows for the detection of severe summer convective events over 360° but may give poor QPE performance during a very significant part of the year when the 0°C isotherm is located below or close to the <span class="hlt">radar</span> altitude; (2) installing <span class="hlt">radar</span> at lower altitudes may lead to better QPE over sensitive areas such as cities located in valleys, but at the cost of reduced visibility and detection capability in other geographical sectors. We have the opportunity to study this question in detail in the region of Grenoble (an Alpine city of 500 000 inhabitants with an average altitude of 210 m asl) with a pair of X-band polarimetric weather <span class="hlt">radars</span> operated respectively by Meteo-France on top of Mount Moucherotte (1920 m asl) and by IGE on the Grenoble Campus (213 m asl). The XPORT <span class="hlt">radar</span> (IGE) performs a combination of PPIs at elevations of 3.5, 7.5, 15 and 25° complemented by two RHIs in the <span class="hlt">vertical</span> plane passing by the two <span class="hlt">radar</span> sites, in order to document the 4D precipitation variability within the Grenoble intermountain valley. In the proposed communication, preliminary results of this experiment (started in September 2016) will be presented with highlights on (1) the calibration of the two <span class="hlt">radar</span> systems, (2) the characterization of the melting layer during significant precipitation events (>5mm/day) occurring in autumn, winter and spring; (3) the simulation of the relative effects of attenuation and non-uniform beam filling at X-band and (4) the possibility to use the mountain returns for quantifying the attenuation by the rain and the melting layer.</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/2017JGRA..122.4777S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JGRA..122.4777S"><span>Meteor <span class="hlt">radar</span> observations of <span class="hlt">vertically</span> propagating low-frequency inertia-gravity waves near the southern polar 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>Song, I.-S.; Lee, C.; Kim, J.-H.; Jee, G.; Kim, Y.-H.; Choi, H.-J.; Chun, H.-Y.; Kim, Y. H.</p> <p>2017-04-01</p> <p><span class="hlt">Vertically</span> propagating low-frequency inertia-gravity waves (IGWs) are retrieved from meteor <span class="hlt">radar</span> winds observed at King Sejong Station (KSS: 62.22°S, 58.78°W), Antarctica. IGW horizontal winds extracted from temporal band-pass filtering in regular time-height bins show the frequent occurrence of IGWs with the downward phase progression and the counterclockwise rotation of their horizontal wind vectors with time (i.e., upward energy propagation) near the mesopause region throughout the whole year of 2014. The <span class="hlt">vertical</span> wavelengths of the observed IGWs roughly range from 14 km to more than 20 km, which is consistent with previous observational studies on the mesospheric IGWs over Antarctica. Stokes parameters and rotary spectra computed from the hodographs of the IGW horizontal wind components reveal that the intrinsic frequencies of the upward propagating IGWs are |f|-3|f| with seasonal variations of the relative predominance between |f|-2|f| and 2|f|-3|f|, where f is the Coriolis parameter at KSS. The hodograph analysis also indicates that the N-S propagation is dominant in austral summer, while the NE-SW propagation is pronounced in austral winter. The propagation direction is discussed in relation to the generation of IGWs due to dynamical imbalances occurring in the tropospheric and stratospheric jet flow systems. Ray tracing results indicate that the N-S propagation in summer may be due to the jet flow systems roughly north of KSS and the NE-SW propagation in winter may be either the SW propagation from the jet flow systems northeast of KSS or the NE propagation (around the South Pole) from the south of Australia and Southern Indian and Pacific Oceans.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ACP....17.6455S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ACP....17.6455S"><span>Quasi-12 h inertia-gravity waves in the lower mesosphere observed by the PANSY <span class="hlt">radar</span> at Syowa Station (39.6° E, 69.0° S)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shibuya, Ryosuke; Sato, Kaoru; Tsutsumi, Masaki; Sato, Toru; Tomikawa, Yoshihiro; Nishimura, Koji; Kohma, Masashi</p> <p>2017-05-01</p> <p>The first observations made by a complete PANSY <span class="hlt">radar</span> system (Program of the Antarctic Syowa MST/IS <span class="hlt">Radar</span>) installed at Syowa Station (39.6° E, 69.0° S) were successfully performed from 16 to 24 March 2015. Over this period, quasi-half-day period (12 h) disturbances in the lower mesosphere at heights of 70 to 80 km were observed. Estimated <span class="hlt">vertical</span> wavelengths, wave periods and <span class="hlt">vertical</span> phase velocities of the disturbances were approximately 13.7 km, 12.3 h and -0.3 m s-1, respectively. Under the working hypothesis that such disturbances are attributable to inertia-gravity waves, wave parameters are estimated using a hodograph analysis. The estimated horizontal wavelengths are longer than 1100 km, and the wavenumber vectors tend to <span class="hlt">point</span> northeastward or southwestward. Using the nonhydrostatic numerical model with a model top of 87 km, quasi-12 h disturbances in the mesosphere were successfully simulated. We show that quasi-12 h disturbances are due to wave-like disturbances with horizontal wavelengths longer than 1400 km and are not due to semidiurnal migrating tides. Wave parameters, such as horizontal wavelengths, <span class="hlt">vertical</span> wavelengths and wave periods, simulated by the model agree well with those estimated by the PANSY <span class="hlt">radar</span> observations under the abovementioned assumption. The parameters of the simulated waves are consistent with the dispersion relationship of the inertia-gravity wave. These results indicate that the quasi-12 h disturbances observed by the PANSY <span class="hlt">radar</span> are attributable to large-scale inertia-gravity waves. By examining a residual of the nonlinear balance equation, it is inferred that the inertia-gravity waves are likely generated by the spontaneous radiation mechanism of two different jet streams. One is the midlatitude tropospheric jet around the tropopause while the other is the polar night jet. Large <span class="hlt">vertical</span> fluxes of zonal and meridional momentum associated with large-scale inertia-gravity waves are distributed across a slanted region</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800021136','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800021136"><span>Wallops waveform analysis of SEASAT-1 <span class="hlt">radar</span> altimeter data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hayne, G. S.</p> <p>1980-01-01</p> <p>Fitting a six parameter model waveform to over ocean experimental data from the waveform samplers in the SEASAT-1 <span class="hlt">radar</span> altimeter is described. The fitted parameters include a waveform risetime, skewness, and track <span class="hlt">point</span>; from these can be obtained estimates of the ocean surface significant waveheight, the surface skewness, and a correction to the altimeter's on board altitude measurement, respectively. Among the difficulties encountered are waveform sampler gains differing from calibration mode data, and incorporating the actual SEASAT-1 sampled <span class="hlt">point</span> target response in the fitted wave form. There are problems in using the spacecraft derived attitude angle estimates, and a different attitude estimator is developed. <span class="hlt">Points</span> raised in this report have consequences for the SEASAT-1 <span class="hlt">radar</span> altimeter's ocean surface measurements are for the design and calibration of <span class="hlt">radar</span> altimeters in future oceanographic satellites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23099859','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23099859"><span>Application of wind-profiling <span class="hlt">radar</span> data to the analysis of dust weather in the Taklimakan Desert.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Minzhong; Wei, Wenshou; Ruan, Zheng; He, Qing; Ge, Runsheng</p> <p>2013-06-01</p> <p>The Urumqi Institute of Desert Meteorology of the China Meteorological Administration carried out an atmospheric scientific experiment to detect dust weather using a wind-profiling <span class="hlt">radar</span> in the hinterland of the Taklimakan Desert in April 2010. Based on the wind-profiling data obtained from this experiment, this paper seeks to (a) analyze the characteristics of the horizontal wind field and <span class="hlt">vertical</span> velocity of a breaking dust weather in a desert hinterland; (b) calculate and give the <span class="hlt">radar</span> echo intensity and <span class="hlt">vertical</span> distribution of a dust storm, blowing sand, and floating dust weather; and (c) discuss the atmosphere dust counts/concentration derived from the wind-profiling <span class="hlt">radar</span> data. Studies show that: (a) A wind-profiling <span class="hlt">radar</span> is an upper-air atmospheric remote sensing system that effectively detects and monitors dust. It captures the beginning and ending of a dust weather process as well as monitors the sand and dust being transported in the air in terms of height, thickness, and <span class="hlt">vertical</span> intensity. (b) The echo intensity of a blowing sand and dust storm weather episode in Taklimakan is about -1~10 dBZ while that of floating dust -1~-15 dBZ, indicating that the dust echo intensity is significantly weaker than that of precipitation but stronger than that of clear air. (c) The <span class="hlt">vertical</span> shear of horizontal wind and the maintenance of low-level east wind are usually dynamic factors causing a dust weather process in Taklimakan. The moment that the low-level horizontal wind field finds a shear over time, it often coincides with the onset of a sand blowing and dust storm weather process. (d) When a blowing sand or dust storm weather event occurs, the atmospheric <span class="hlt">vertical</span> velocity tends to be of upward motion. This <span class="hlt">vertical</span> upward movement of the atmosphere supported with a fast horizontal wind and a dry underlying surface carries dust particles from the ground up to the air to form blown sand or a dust storm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110012243','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110012243"><span>Airborne <span class="hlt">Radar</span> Interferometric Repeat-Pass Processing</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hensley, Scott; Michel, Thierry R.; Jones, Cathleen E.; Muellerschoen, Ronald J.; Chapman, Bruce D.; Fore, Alexander; Simard, Marc; Zebker, Howard A.</p> <p>2011-01-01</p> <p>Earth science research often requires crustal deformation measurements at a variety of time scales, from seconds to decades. Although satellites have been used for repeat-track interferometric (RTI) synthetic-aperture-<span class="hlt">radar</span> (SAR) mapping for close to 20 years, RTI is much more difficult to implement from an airborne platform owing to the irregular trajectory of the aircraft compared with microwave imaging <span class="hlt">radar</span> wavelengths. Two basic requirements for robust airborne repeat-pass <span class="hlt">radar</span> interferometry include the ability to fly the platform to a desired trajectory within a narrow tube and the ability to have the <span class="hlt">radar</span> beam <span class="hlt">pointed</span> in a desired direction to a fraction of a beam width. Uninhabited Aerial Vehicle Synthetic Aperture <span class="hlt">Radar</span> (UAVSAR) is equipped with a precision auto pilot developed by NASA Dryden that allows the platform, a Gulfstream III, to nominally fly within a 5 m diameter tube and with an electronically scanned antenna to position the <span class="hlt">radar</span> beam to a fraction of a beam width based on INU (inertial navigation unit) attitude angle measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.9248E..0IM','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.9248E..0IM"><span>All-digital <span class="hlt">radar</span> architecture</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Molchanov, Pavlo A.</p> <p>2014-10-01</p> <p>All digital <span class="hlt">radar</span> architecture requires exclude mechanical scan system. The phase antenna array is necessarily large because the array elements must be co-located with very precise dimensions and will need high accuracy phase processing system for aggregate and distribute T/R modules data to/from antenna elements. Even phase array cannot provide wide field of view. New nature inspired all digital <span class="hlt">radar</span> architecture proposed. The fly's eye consists of multiple angularly spaced sensors giving the fly simultaneously thee wide-area visual coverage it needs to detect and avoid the threats around him. Fly eye <span class="hlt">radar</span> antenna array consist multiple directional antennas loose distributed along perimeter of ground vehicle or aircraft and coupled with receiving/transmitting front end modules connected by digital interface to central processor. Non-steering antenna array allows creating all-digital <span class="hlt">radar</span> with extreme flexible architecture. Fly eye <span class="hlt">radar</span> architecture provides wide possibility of digital modulation and different waveform generation. Simultaneous correlation and integration of thousands signals per second from each <span class="hlt">point</span> of surveillance area allows not only detecting of low level signals ((low profile targets), but help to recognize and classify signals (targets) by using diversity signals, polarization modulation and intelligent processing. Proposed all digital <span class="hlt">radar</span> architecture with distributed directional antenna array can provide a 3D space vector to the jammer by verification direction of arrival for signals sources and as result jam/spoof protection not only for <span class="hlt">radar</span> systems, but for communication systems and any navigation constellation system, for both encrypted or unencrypted signals, for not limited number or close positioned jammers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20070035996&hterms=LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DLDR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20070035996&hterms=LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DLDR"><span>Processing of High Resolution, Multiparametric <span class="hlt">Radar</span> Data for the Airborne Dual-Frequency Precipitation <span class="hlt">Radar</span> APR-2</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tanelli, Simone; Meagher, Jonathan P.; Durden, Stephen L.; Im, Eastwood</p> <p>2004-01-01</p> <p>Following the successful Precipitation <span class="hlt">Radar</span> (PR) of the Tropical Rainfall Measuring Mission, a new airborne, 14/35 GHz rain profiling <span class="hlt">radar</span>, known as Airborne Precipitation <span class="hlt">Radar</span> - 2 (APR-2), has been developed as a prototype for an advanced, dual-frequency spaceborne <span class="hlt">radar</span> for a future spaceborne precipitation measurement mission. . This airborne instrument is capable of making simultaneous measurements of rainfall parameters, including co-pol and cross-pol rain reflectivities and <span class="hlt">vertical</span> Doppler velocities, at 14 and 35 GHz. furthermore, it also features several advanced technologies for performance improvement, including real-time data processing, low-sidelobe dual-frequency pulse compression, and dual-frequency scanning antenna. Since August 2001, APR-2 has been deployed on the NASA P3 and DC8 aircrafts in four experiments including CAMEX-4 and the Wakasa Bay Experiment. Raw <span class="hlt">radar</span> data are first processed to obtain reflectivity, LDR (linear depolarization ratio), and Doppler velocity measurements. The dataset is then processed iteratively to accurately estimate the true aircraft navigation parameters and to classify the surface return. These intermediate products are then used to refine reflectivity and LDR calibrations (by analyzing clear air ocean surface returns), and to correct Doppler measurements for the aircraft motion. Finally, the the melting layer of precipitation is detected and its boundaries and characteristics are identifIed at the APR-2 range resolution of 30m. The resulting 3D dataset will be used for validation of other airborne and spaceborne instruments, development of multiparametric rain/snow retrieval algorithms and melting layer characterization and statistics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405980-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-point-lonely-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405980-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-risk-assessment-point-lonely-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron, Elmendorf AFB, Alaska. Risk assessment <span class="hlt">Point</span> Lonely <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-04-01</p> <p>This document contains the baseline human health risk assessment and the ecological risk assessment (ERA) for the <span class="hlt">Point</span> Lonely Distant Early Warning (DEW) Line <span class="hlt">radar</span> installation. Twelve sites at the <span class="hlt">Point</span> Lonely <span class="hlt">radar</span> installation underwent remedial investigations (RIs) during the summer of 1993. The Vehicle Storage Area (SS14) was combined with the Inactive Landfill because the two sites were essentially co-located and were sampled during the RI as a single unit. Therefore, 11 sites are discussed in this risk assessment. The presence of chemical contamination in the soil, sediments, and surface water at the installation was evaluated and reported inmore » the <span class="hlt">Point</span> Lonely Remedial Investigation/Feasibility Study (RI/FS). The analytical data reported in the RI/FS form the basis for the human health and ecological risk assessments. The primary chemicals of concern (COCs) at the 11 sites are diesel and gasoline from past spills and/or leaks, chlorinated solvents, and manganese. The 11 sites investigated and the types of samples collected at each site are presented.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850024225','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850024225"><span>MENTOR: Adding an outlying receiver to an ST <span class="hlt">radar</span> for meteor-wind measurement</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Roper, R. G.</p> <p>1984-01-01</p> <p><span class="hlt">Radar</span> scattering from ionized meteor trails has been used for many years as a way to determine mesopause-level winds. Scattering occurs perpendicular to the trails, and since the ionizing efficiency of the incoming meteoroids depends on the cosine of the zenith angle of the radiant, echoes directly overhead are rare. Stratosphere-troposphere (ST) <span class="hlt">radars</span> normally sample within 15 deg of the <span class="hlt">vertical</span>, and thus receive few meteor echoes. Even the higher powdered mesosphere-stratosphere-troposphere (MST) <span class="hlt">radars</span> are not good meteor <span class="hlt">radars</span>, although they were used to successfully retrieved meteor winds from the Poker Flat, Alaska MST <span class="hlt">radar</span> by averaging long data intervals. It has been suggested that a receiving station some distance from an ST <span class="hlt">radar</span> could receive pulses being scattered from meteor trails, determine the particular ST beam in which the scattering occurred, measure the radial Doppler velocity, and thus determine the wind field. This concept has been named MENTOR (Meteor Echoes; No Transmitter, Only Receivers).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01727.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01727.html"><span>Space <span class="hlt">Radar</span> Image of Mt. Rainer, Washington</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This is a <span class="hlt">radar</span> image of Mount Rainier in Washington state. The volcano last erupted about 150 years ago and numerous large floods and debris flows have originated on its slopes during the last century. Today the volcano is heavily mantled with glaciers and snowfields. More than 100,000 people live on young volcanic mudflows less than 10,000 years old and, consequently, are within the range of future, devastating mudslides. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 20th orbit on October 1, 1994. The area shown in the image is approximately 59 kilometers by 60 kilometers (36.5 miles by 37 miles). North is toward the top left of the image, which was composed by assigning red and green colors to the L-band, horizontally transmitted and <span class="hlt">vertically</span>, and the L-band, horizontally transmitted and <span class="hlt">vertically</span> received. Blue indicates the C-band, horizontally transmitted and <span class="hlt">vertically</span> received. In addition to highlighting topographic slopes facing the space shuttle, SIR-C records rugged areas as brighter and smooth areas as darker. The scene was illuminated by the shuttle's <span class="hlt">radar</span> from the northwest so that northwest-facing slopes are brighter and southeast-facing slopes are dark. Forested regions are pale green in color; clear cuts and bare ground are bluish or purple; ice is dark green and white. The round cone at the center of the image is the 14,435-foot (4,399-meter) active volcano, Mount Rainier. On the lower slopes is a zone of rock ridges and rubble (purple to reddish) above coniferous forests (in yellow/green). The western boundary of Mount Rainier National Park is seen as a transition from protected, old-growth forest to heavily logged private land, a mosaic of recent clear cuts (bright purple/blue) and partially regrown timber plantations (pale blue). The prominent river seen curving away from the mountain at the top of the image (to the northwest) is the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140002254','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140002254"><span>Longitudinal Differences of Ionospheric <span class="hlt">Vertical</span> Density Distribution and Equatorial Electrodynamics</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yizengaw, E.; Zesta, E.; Moldwin, M. B.; Damtie, B.; Mebrahtu, A.; Valledares, C.E.; Pfaff, R. F.</p> <p>2012-01-01</p> <p>Accurate estimation of global <span class="hlt">vertical</span> distribution of ionospheric and plasmaspheric density as a function of local time, season, and magnetic activity is required to improve the operation of space-based navigation and communication systems. The <span class="hlt">vertical</span> density distribution, especially at low and equatorial latitudes, is governed by the equatorial electrodynamics that produces a <span class="hlt">vertical</span> driving force. The <span class="hlt">vertical</span> structure of the equatorial density distribution can be observed by using tomographic reconstruction techniques on ground-based global positioning system (GPS) total electron content (TEC). Similarly, the <span class="hlt">vertical</span> drift, which is one of the driving mechanisms that govern equatorial electrodynamics and strongly affect the structure and dynamics of the ionosphere in the low/midlatitude region, can be estimated using ground magnetometer observations. We present tomographically reconstructed density distribution and the corresponding <span class="hlt">vertical</span> drifts at two different longitudes: the East African and west South American sectors. Chains of GPS stations in the east African and west South American longitudinal sectors, covering the equatorial anomaly region of meridian approx. 37 deg and 290 deg E, respectively, are used to reconstruct the <span class="hlt">vertical</span> density distribution. Similarly, magnetometer sites of African Meridian B-field Education and Research (AMBER) and INTERMAGNET for the east African sector and South American Meridional B-field Array (SAMBA) and Low Latitude Ionospheric Sensor Network (LISN) are used to estimate the <span class="hlt">vertical</span> drift velocity at two distinct longitudes. The comparison between the reconstructed and Jicamarca Incoherent Scatter <span class="hlt">Radar</span> (ISR) measured density profiles shows excellent agreement, demonstrating the usefulness of tomographic reconstruction technique in providing the <span class="hlt">vertical</span> density distribution at different longitudes. Similarly, the comparison between magnetometer estimated <span class="hlt">vertical</span> drift and other independent drift observation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1169516','SCIGOV-DOEDE'); return false;" href="https://www.osti.gov/servlets/purl/1169516"><span>SGP and TWP (Manus) Ice Cloud <span class="hlt">Vertical</span> Velocities</span></a></p> <p><a target="_blank" href="http://www.osti.gov/dataexplorer">DOE Data Explorer</a></p> <p>Kalesse, Heike</p> <p>2013-06-27</p> <p>Daily netcdf-files of ice-cloud dynamics observed at the ARM sites at SGP (Jan1997-Dec2010) and Manus (Jul1999-Dec2010). The files include variables at different time resolution (10s, 20min, 1hr). Profiles of <span class="hlt">radar</span> reflectivity factor (dbz), Doppler velocity (vel) as well as retrieved <span class="hlt">vertical</span> air motion (V_air) and reflectivity-weighted particle terminal fall velocity (V_ter) are given at 10s, 20min and 1hr resolution. Retrieved V_air and V_ter follow <span class="hlt">radar</span> notation, so positive values indicate downward motion. Lower level clouds are removed, however a multi-layer flag is included.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090032029','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090032029"><span>Evaluation of TRMM Ground-Validation <span class="hlt">Radar</span>-Rain Errors Using Rain Gauge Measurements</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wang, Jianxin; Wolff, David B.</p> <p>2009-01-01</p> <p>Ground-validation (GV) <span class="hlt">radar</span>-rain products are often utilized for validation of the Tropical Rainfall Measuring Mission (TRMM) spaced-based rain estimates, and hence, quantitative evaluation of the GV <span class="hlt">radar</span>-rain product error characteristics is vital. This study uses quality-controlled gauge data to compare with TRMM GV <span class="hlt">radar</span> rain rates in an effort to provide such error characteristics. The results show that significant differences of concurrent <span class="hlt">radar</span>-gauge rain rates exist at various time scales ranging from 5 min to 1 day, despite lower overall long-term bias. However, the differences between the <span class="hlt">radar</span> area-averaged rain rates and gauge <span class="hlt">point</span> rain rates cannot be explained as due to <span class="hlt">radar</span> error only. The error variance separation method is adapted to partition the variance of <span class="hlt">radar</span>-gauge differences into the gauge area-<span class="hlt">point</span> error variance and <span class="hlt">radar</span> rain estimation error variance. The results provide relatively reliable quantitative uncertainty evaluation of TRMM GV <span class="hlt">radar</span> rain estimates at various times scales, and are helpful to better understand the differences between measured <span class="hlt">radar</span> and gauge rain rates. It is envisaged that this study will contribute to better utilization of GV <span class="hlt">radar</span> rain products to validate versatile spaced-based rain estimates from TRMM, as well as the proposed Global Precipitation Measurement, and other satellites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950031653&hterms=work+shift&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwork%2Bshift','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950031653&hterms=work+shift&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dwork%2Bshift"><span>Measuring rainwater content by <span class="hlt">radar</span> using propagation differential phase shift</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Jameson, A. R.</p> <p>1994-01-01</p> <p>While <span class="hlt">radars</span> measure several quantities closely coupled to the rainfall rate, for frequencies less than 15 GHz, estimates of the rainwater content W are traditionally computed from the <span class="hlt">radar</span> reflectivity factor Z or the rate of attenuation A--quantities only weakly related to W. Consequently, instantaneous <span class="hlt">point</span> estimates of W using Z and A are often erroneous. A more natural, alternative parameter for estimating W at these frequencies is the specific polarization propagation differential phase shift phi(sub DP), which is a measure of the change in the difference between phases of <span class="hlt">vertically</span> (V) and horizontally (H) polarized waves with increasing distance from a <span class="hlt">radar</span>. It is now well known that W is nearly linearly related to phi(sub DP) divided by (1 - reversed R), where reversed R is the mass-weighted mean axis ratio of the raindrops. Unfortunately, such relations are not widely used in part because measurements of phi(sub DP) are scarce but also because one must determine reversed R. In this work it is shown that this parameter can be estimated using the differential reflectivity (Z(sub H)/Z(sub V) at 3 GHz. An alternative technique is suggested for higher frequencies when the differential reflectivity becomes degraded by attenuation. While theory indicates that it should be possible using phi(sub DP) to estimate W quite accurately, measurement errors increase the uncertainty to +/- 18%-35% depending on reversed R. While far from ideal, it appears that these estimates are likely to be considerably more accurate than those deduced using currently available methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012icha.book...70K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012icha.book...70K"><span>Kharkiv Meteor <span class="hlt">Radar</span> System (the XX Age)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kolomiyets, S. V.</p> <p>2012-09-01</p> <p>Kharkiv meteor <span class="hlt">radar</span> research are of historic value (Kolomiyets and Sidorov 2007). Kharkiv <span class="hlt">radar</span> observations of meteors proved internationally as the best in the world, it was noted at the IAU General Assembly in 1958. In the 1970s Kharkiv meteor automated <span class="hlt">radar</span> system (MARS) was recommended at the international level as a successful prototype for wide distribution. Until now, this <span class="hlt">radar</span> system is one of the most sensitive instruments of meteor <span class="hlt">radars</span> in the world for astronomical observations. In 2004 Kharkiv meteor <span class="hlt">radar</span> system is included in the list of objects which compose the national property of Ukraine. Kharkiv meteor <span class="hlt">radar</span> system has acquired the status of the important historical astronomical instrument in world history. Meteor Centre for researching meteors in Kharkiv is a analogue of the observatory and performs the same functions of a generator and a battery of special knowledge and skills (the world-famous studio). Kharkiv and the location of the instrument were brand <span class="hlt">points</span> on the globe, as the place where the world-class meteor <span class="hlt">radar</span> studies were carried out. They are inscribed in the history of meteor astronomy, in large letters and should be immortalized on a world-wide level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A42C..02H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A42C..02H"><span>Toward Improving Ice Water Content and Snow Rate Retrievals from Spaceborne <span class="hlt">Radars</span>, Emphasizing Ku and Ka-Bands</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Heymsfield, A.; Bansemer, A.; Tanelli, S.; Poellot, M.</p> <p>2015-12-01</p> <p>This study uses a data set from either overflying aircraft or ground-based <span class="hlt">radars</span> operating at Ku and Ka bands, combined with in-situ microphysical measurements to develop <span class="hlt">radar</span> reflectivity (Ze)-ice water content (IWC) and Ze-snowfall rate (S) relationships that are suited for retrieval of snowfall rate from the GPM <span class="hlt">radars</span>. During GCPEX, the NASA DC-8 aircraft, carrying the JPL APR-2 KU and KA band <span class="hlt">radars</span> overflew the UND Citation aircraft, making microphysical measurements in the ice clouds below. On two days, 19 and 28 January 2011, there are a total of almost 7000 1-sec colocations of the aircraft, where a collocation was defined as having a combination of a spatial separation of less than 3 km and a time separation of less than 10 minutes. During the NASA GPM Mid-latitude Continental Convective Cloud Experiment (MC3E), the Citation aircraft made in-situ observations over Oklahoma in 2011. We evaluated the data from two types of collocations. First, there were two Citation spirals on 27 April 2011, over the NPOL <span class="hlt">radar</span>. At the same time, the UHF-band KUZR <span class="hlt">radar</span> was collecting data in a <span class="hlt">vertically-pointing</span> mode. Also, the Ka band KAZR Doppler <span class="hlt">radar</span> was operating in a zenith orientation. Reflectivities and Doppler velocities, without and with appreciable Mie-scattering effects of the hydrometers (for KUZR and KAZR, respectively), are thus available during the spirals. Also during MC3E, six deep convective clouds with a total of more than 5000 5-sec samples and a range of temperatures from -40 to 0C were sampled by the Citation at the same time that NEXRAD reflectivities were measured at about the same position. These data allows us to evaluate various backscatter models and to develop multi-wavelength Z-IWC and Z-S relationships. We will present the results of this study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930043226&hterms=rain+storm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Drain%2Bstorm','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930043226&hterms=rain+storm&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Drain%2Bstorm"><span>Preliminary results from multiparameter airborne rain <span class="hlt">radar</span> measurement in the western Pacific</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kumagai, Hiroshi; Meneghini, Robert; Kozu, Toshiaki</p> <p>1993-01-01</p> <p>Preliminary results are presented from multiparameter airborne <span class="hlt">radar</span> measurements of tropical storms. The experiment was conducted in the western Pacific in September 1990 with the NASA DC-8 aircraft that was equipped with a dual-wavelength <span class="hlt">radar</span> at X and Ka bands and several microwave radiometers. The modification to dual-polarization at X-band <span class="hlt">radar</span> enabled measurements of the linear depolarization ratio (LDR). <span class="hlt">Vertical</span> profiles of dual-polarization and dual-frequency observables for an example of stratiform rain and three examples of convective rain cells are examined. It is shown that at nadir incidence the LDR measurement often can be used to distinguish the phase states of the hydrometeors and to identify the melting layer. In addition to the information concerning particle shape and orientation from LDR, the ratio of the <span class="hlt">radar</span> reflectivity factors in two frequency bands (X and Ka bands) provides insight into particle size. The capabilities of dual-wavelength and dual-polarization <span class="hlt">radar</span> in the identification of particle size and phase will be important considerations in the design of future spaceborne weather <span class="hlt">radars</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040012672&hterms=deep+neural+network&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Ddeep%2Bneural%2Bnetwork','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040012672&hterms=deep+neural+network&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Ddeep%2Bneural%2Bnetwork"><span>Objective Classification of <span class="hlt">Radar</span> Profile Types, and Their Relationship to Lightning Occurrence</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Boccippio, Dennis</p> <p>2003-01-01</p> <p>A cluster analysis technique is used to identify 16 "archetypal" <span class="hlt">vertical</span> <span class="hlt">radar</span> profile types from a large, globally representative sample of profiles from the TRMM Precipitation <span class="hlt">Radar</span>. These include nine convective types (7 of these deep convective) and seven stratiform types (5 of these clearly glaciated). <span class="hlt">Radar</span> profile classification provides an alternative to conventional deep convective storm metrics, such as 30 dBZ echo height, maximum reflectivity or VIL. As expected, the global frequency of occurrence of deep convective profile types matches satellite-observed total lightning production, including to very small scall local features. Each location's "mix" of profile types provides an objective description of the local convective spectrum, and in turn, is a first step in objectively classifying convective regimes. These classifiers are tested as inputs to a neural network which attempts to predict lightning occurrence based on <span class="hlt">radar</span>-only storm observations, and performance is compared with networks using traditional <span class="hlt">radar</span> metrics as inputs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/870851','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/870851"><span>Imaging synthetic aperture <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Burns, Bryan L.; Cordaro, J. Thomas</p> <p>1997-01-01</p> <p>A linear-FM SAR imaging <span class="hlt">radar</span> method and apparatus to produce a real-time image by first arranging the returned signals into a plurality of subaperture arrays, the columns of each subaperture array having samples of dechirped baseband pulses, and further including a processing of each subaperture array to obtain coarse-resolution in azimuth, then fine-resolution in range, and lastly, to combine the processed subapertures to obtain the final fine-resolution in azimuth. Greater efficiency is achieved because both the transmitted signal and a local oscillator signal mixed with the returned signal can be varied on a pulse-to-pulse basis as a function of <span class="hlt">radar</span> motion. Moreover, a novel circuit can adjust the sampling location and the A/D sample rate of the combined dechirped baseband signal which greatly reduces processing time and hardware. The processing steps include implementing a window function, stabilizing either a central reference <span class="hlt">point</span> and/or all other <span class="hlt">points</span> of a subaperture with respect to doppler frequency and/or range as a function of <span class="hlt">radar</span> motion, sorting and compressing the signals using a standard fourier transforms. The stabilization of each processing part is accomplished with vector multiplication using waveforms generated as a function of <span class="hlt">radar</span> motion wherein these waveforms may be synthesized in integrated circuits. Stabilization of range migration as a function of doppler frequency by simple vector multiplication is a particularly useful feature of the invention; as is stabilization of azimuth migration by correcting for spatially varying phase errors prior to the application of an autofocus process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950005961','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950005961"><span>The design and development of signal-processing algorithms for an airborne x-band Doppler weather <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nicholson, Shaun R.</p> <p>1994-01-01</p> <p>Improved measurements of precipitation will aid our understanding of the role of latent heating on global circulations. Spaceborne meteorological sensors such as the planned precipitation <span class="hlt">radar</span> and microwave radiometers on the Tropical Rainfall Measurement Mission (TRMM) provide for the first time a comprehensive means of making these global measurements. Pre-TRMM activities include development of precipitation algorithms using existing satellite data, computer simulations, and measurements from limited aircraft campaigns. Since the TRMM <span class="hlt">radar</span> will be the first spaceborne precipitation <span class="hlt">radar</span>, there is limited experience with such measurements, and only recently have airborne <span class="hlt">radars</span> become available that can attempt to address the issue of the limitations of a spaceborne <span class="hlt">radar</span>. There are many questions regarding how much attenuation occurs in various cloud types and the effect of cloud <span class="hlt">vertical</span> motions on the estimation of precipitation rates. The EDOP program being developed by NASA GSFC will provide data useful for testing both rain-retrieval algorithms and the importance of <span class="hlt">vertical</span> motions on the rain measurements. The purpose of this report is to describe the design and development of real-time embedded parallel algorithms used by EDOP to extract reflectivity and Doppler products (velocity, spectrum width, and signal-to-noise ratio) as the first step in the aforementioned goals.</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('https://ntrs.nasa.gov/search.jsp?R=19840008350&hterms=camouflage&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcamouflage','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840008350&hterms=camouflage&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dcamouflage"><span><span class="hlt">Radar</span> activities of the DFVLR Institute for Radio Frequency Technology</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Keydel, W.</p> <p>1983-01-01</p> <p>Aerospace research and the respective applications microwave tasks with respect to remote sensing, position finding and communication are discussed. The <span class="hlt">radar</span> activities are directed at <span class="hlt">point</span> targets, area targets and volume targets; they center around signature research for earth and ocean remote sensing, target recognition, reconnaissance and camouflage and imaging and area observation <span class="hlt">radar</span> techniques (SAR and SLAR). The <span class="hlt">radar</span> activities cover a frequency range from 1 GHz up to 94 GHz. The <span class="hlt">radar</span> program is oriented to four possible application levels: ground, air, shuttle orbits and satellite orbits. Ground based studies and measurements, airborne scatterometers and imaging <span class="hlt">radars</span>, a space shuttle <span class="hlt">radar</span>, the MRSE, and follow on experiments are considered.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840000263&hterms=imitation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dimitation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840000263&hterms=imitation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dimitation"><span>Synchronized <span class="hlt">Radar</span>-Target Simulator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chin, B. C.</p> <p>1985-01-01</p> <p>Apparatus for testing <span class="hlt">radar</span> system generates signals that simulate amplitude and phase characteristics of target returns and their variation with antenna-<span class="hlt">pointing</span> direction. Antenna movement causes equipment to alter test signal in imitation of behavior of real signal received during tracking.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016E%26ES...46a2013L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016E%26ES...46a2013L"><span>Airborne LIDAR <span class="hlt">point</span> cloud tower inclination judgment</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>liang, Chen; zhengjun, Liu; jianguo, Qian</p> <p>2016-11-01</p> <p>Inclined transmission line towers for the safe operation of the line caused a great threat, how to effectively, quickly and accurately perform inclined judgment tower of power supply company safety and security of supply has played a key role. In recent years, with the development of unmanned aerial vehicles, unmanned aerial vehicles equipped with a laser scanner, GPS, inertial navigation is one of the high-precision 3D Remote Sensing System in the electricity sector more and more. By airborne <span class="hlt">radar</span> scan <span class="hlt">point</span> cloud to visually show the whole picture of the three-dimensional spatial information of the power line corridors, such as the line facilities and equipment, terrain and trees. Currently, LIDAR <span class="hlt">point</span> cloud research in the field has not yet formed an algorithm to determine tower inclination, the paper through the existing power line corridor on the tower base extraction, through their own tower shape characteristic analysis, a <span class="hlt">vertical</span> stratification the method of combining convex hull algorithm for <span class="hlt">point</span> cloud tower scarce two cases using two different methods for the tower was Inclined to judge, and the results with high reliability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.V43D3172A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.V43D3172A"><span>Automatic Real-Time Estimation of Plume Height and Mass Eruption Rate Using <span class="hlt">Radar</span> Data During Explosive Volcanism</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arason, P.; Barsotti, S.; De'Michieli Vitturi, M.; Jónsson, S.; Arngrímsson, H.; Bergsson, B.; Pfeffer, M. A.; Petersen, G. N.; Bjornsson, H.</p> <p>2016-12-01</p> <p>Plume height and mass eruption rate are the principal scale parameters of explosive volcanic eruptions. Weather <span class="hlt">radars</span> are important instruments in estimating plume height, due to their independence of daylight, weather and visibility. The Icelandic Meteorological Office (IMO) operates two fixed position C-band weather <span class="hlt">radars</span> and two mobile X-band <span class="hlt">radars</span>. All volcanoes in Iceland can be monitored by IMO's <span class="hlt">radar</span> network, and during initial phases of an eruption all available <span class="hlt">radars</span> will be set to a more detailed volcano scan. When the <span class="hlt">radar</span> volume data is retrived at IMO-headquarters in Reykjavík, an automatic analysis is performed on the <span class="hlt">radar</span> data above the proximity of the volcano. The plume height is automatically estimated taking into account the <span class="hlt">radar</span> scanning strategy, beam width, and a likely reflectivity gradient at the plume top. This analysis provides a distribution of the likely plume height. The automatically determined plume height estimates from the <span class="hlt">radar</span> data are used as input to a numerical suite that calculates the eruptive source parameters through an inversion algorithm. This is done by using the coupled system DAKOTA-PlumeMoM which solves the 1D plume model equations iteratively by varying the input values of vent radius and <span class="hlt">vertical</span> velocity. The model accounts for the effect of wind on the plume dynamics, using atmospheric <span class="hlt">vertical</span> profiles extracted from the ECMWF numerical weather prediction model. Finally, the resulting estimates of mass eruption rate are used to initialize the dispersal model VOL-CALPUFF to assess hazard due to tephra fallout, and communicated to London VAAC to support their modelling activity for aviation safety purposes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8361E..0KS','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8361E..0KS"><span>Indoor imagery with a 3D through-wall synthetic aperture <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sévigny, Pascale; DiFilippo, David J.; Laneve, Tony; Fournier, Jonathan</p> <p>2012-06-01</p> <p>Through-wall <span class="hlt">radar</span> imaging is an emerging technology with great interest to military and police forces operating in an urban environment. A through-wall imaging <span class="hlt">radar</span> can potentially provide interior room layouts as well as detection and localization of targets of interest within a building. In this paper, we present our through-wall <span class="hlt">radar</span> system mounted on the side of a vehicle and driven along a path in front of a building of interest. The vehicle is equipped with a LIDAR (Light Detection and Ranging) and motion sensors that provide auxiliary information. The <span class="hlt">radar</span> uses an ultra wideband frequency-modulated continuous wave (FMCW) waveform to obtain high range resolution. Our system is composed of a <span class="hlt">vertical</span> linear receive array to discriminate targets in elevation, and two transmit elements operated in a slow multiple-input multiple output (MIMO) configuration to increase the achievable elevation resolution. High resolution in the along-track direction is obtained through synthetic aperture <span class="hlt">radar</span> (SAR) techniques. We present experimental results that demonstrate the 3-D capability of the <span class="hlt">radar</span>. We further demonstrate target detection behind challenging walls, and imagery of internal wall features. Finally, we discuss future work.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1922l0003P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1922l0003P"><span>Analytical approach to determine <span class="hlt">vertical</span> dynamics of a semi-trailer truck from the <span class="hlt">point</span> of view of goods protection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pidl, Renáta</p> <p>2018-01-01</p> <p>The overwhelming majority of intercontinental long-haul transportations of goods are usually carried out on road by semi-trailer trucks. Vibration has a major effect regarding the safety of the transport, the load and the transported goods. This paper deals with the logistics goals from the <span class="hlt">point</span> of view of vibration and summarizes the methods to predict or measure the vibration load in order to design a proper system. From these methods, the focus of this paper is on the computer simulation of the vibration. An analytical method is presented to calculate the <span class="hlt">vertical</span> dynamics of a semi-trailer truck containing general viscous damping and exposed to harmonic base excitation. For the purpose of a better understanding, the method will be presented through a simplified four degrees-of-freedom (DOF) half-vehicle model, which neglects the stiffness and damping of the tires, thus the four degrees-of-freedom are the <span class="hlt">vertical</span> and angular displacements of the truck and the trailer. From the <span class="hlt">vertical</span> and angular accelerations of the trailer, the <span class="hlt">vertical</span> acceleration of each <span class="hlt">point</span> of the platform of the trailer can easily be determined, from which the forces acting on the transported goods are given. As a result of this paper the response of the full platform-load-packaging system to any kind of vehicle, any kind of load and any kind of road condition can be analyzed. The peak acceleration of any <span class="hlt">point</span> on the platform can be determined by the presented analytical method.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFM.A42C..04T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFM.A42C..04T"><span>Retrievals of Ice Cloud Microphysical Properties of Deep Convective Systems using <span class="hlt">Radar</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>Tian, J.; Dong, X.; Xi, B.; Wang, J.; Homeyer, C. R.</p> <p>2015-12-01</p> <p>This study presents innovative algorithms for retrieving ice cloud microphysical properties of Deep Convective Systems (DCSs) using Next-Generation <span class="hlt">Radar</span> (NEXRAD) reflectivity and newly derived empirical relationships from aircraft in situ measurements in Wang et al. (2015) during the Midlatitude Continental Convective Clouds Experiment (MC3E). With composite gridded NEXRAD <span class="hlt">radar</span> reflectivity, four-dimensional (space-time) ice cloud microphysical properties of DCSs are retrieved, which is not possible from either in situ sampling at a single altitude or from <span class="hlt">vertical</span> <span class="hlt">pointing</span> <span class="hlt">radar</span> measurements. For this study, aircraft in situ measurements provide the best-estimated ice cloud microphysical properties for validating the <span class="hlt">radar</span> retrievals. Two statistical comparisons between retrieved and aircraft in situ measured ice microphysical properties are conducted from six selected cases during MC3E. For the temporal-averaged method, the averaged ice water content (IWC) and median mass diameter (Dm) from aircraft in situ measurements are 0.50 g m-3 and 1.51 mm, while the retrievals from <span class="hlt">radar</span> reflectivity have negative biases of 0.12 g m-3 (24%) and 0.02 mm (1.3%) with correlations of 0.71 and 0.48, respectively. For the spatial-averaged method, the IWC retrievals are closer to the aircraft results (0.51 vs. 0.47 g m-3) with a positive bias of 8.5%, whereas the Dm retrievals are larger than the aircraft results (1.65 mm vs. 1.51 mm) with a positive bias of 9.3%. The retrieved IWCs decrease from ~0.6 g m-3 at 5 km to ~0.15 g m-3 at 13 km, and Dm values decrease from ~2 mm to ~0.7 mm at the same levels. In general, the aircraft in situ measured IWC and Dm values at each level are within one standard derivation of retrieved properties. Good agreements between microphysical properties measured from aircraft and retrieved from <span class="hlt">radar</span> reflectivity measurements indicate the reasonable accuracy of our retrievals.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1253785','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1253785"><span><span class="hlt">Radar</span> Wind Profiler for Cloud Forecasting at Brookhaven National Laboratory (BNL) Field Campaign Report</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>Jensen, Michael P; Giangrande, Scott E; Bartholomew, Mary Jane</p> <p></p> <p>The <span class="hlt">Radar</span> Wind Profiler for Cloud Forecasting at Brookhaven National Laboratory (BNL) [http://www.arm.gov/campaigns/osc2013rwpcf] campaign was scheduled to take place from 15 July 2013 through 15 July 2015 (or until shipped for the next U.S. Department of Energy Atmospheric Radiation Measurement [ARM] Climate Research Facility first Mobile Facility [AMF1] deployment). The campaign involved the deployment of the AMF1 Scintec 915 MHz <span class="hlt">Radar</span> Wind Profiler (RWP) at BNL, in conjunction with several other ARM, BNL and National Weather Service (NWS) instruments. The two main scientific foci of the campaign were: 1) To provide profiles of the horizontal wind to be used tomore » test and validate short-term cloud advection forecasts for solar-energy applications and 2) to provide <span class="hlt">vertical</span> profiling capabilities for the study of dynamics (i.e., <span class="hlt">vertical</span> velocity) and hydrometeors in winter storms. This campaign was a serendipitous opportunity that arose following the deployment of the RWP at the Two-Column Aerosol Project (TCAP) campaign in Cape Cod, Massachusetts and restriction from participation in the Green Ocean Amazon 2014/15 (GoAmazon 2014/15) campaign due to radio-frequency allocation restriction for international deployments. The RWP arrived at BNL in the fall of 2013, but deployment was delayed until fall of 2014 as work/safety planning and site preparation were completed. The RWP further encountered multiple electrical failures, which eventually required several shipments of instrument power supplies and the final amplifier to the vendor to complete repairs. Data collection began in late January 2015. The operational modes of the RWP were changed such that in addition to collecting traditional profiles of the horizontal wind, a <span class="hlt">vertically</span> <span class="hlt">pointing</span> mode was also included for the purpose of precipitation sensing and estimation of <span class="hlt">vertical</span> velocities. The RWP operated well until the end of the campaign in July 2015 and collected observations for more than 20</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AMT.....9.3837V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AMT.....9.3837V"><span>Close-range <span class="hlt">radar</span> rainfall estimation and error analysis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>van de Beek, C. Z.; Leijnse, H.; Hazenberg, P.; Uijlenhoet, R.</p> <p>2016-08-01</p> <p>Quantitative precipitation estimation (QPE) using ground-based weather <span class="hlt">radar</span> is affected by many sources of error. The most important of these are (1) <span class="hlt">radar</span> calibration, (2) ground clutter, (3) wet-radome attenuation, (4) rain-induced attenuation, (5) <span class="hlt">vertical</span> variability in rain drop size distribution (DSD), (6) non-uniform beam filling and (7) variations in DSD. This study presents an attempt to separate and quantify these sources of error in flat terrain very close to the <span class="hlt">radar</span> (1-2 km), where (4), (5) and (6) only play a minor role. Other important errors exist, like beam blockage, WLAN interferences and hail contamination and are briefly mentioned, but not considered in the analysis. A 3-day rainfall event (25-27 August 2010) that produced more than 50 mm of precipitation in De Bilt, the Netherlands, is analyzed using <span class="hlt">radar</span>, rain gauge and disdrometer data. Without any correction, it is found that the <span class="hlt">radar</span> severely underestimates the total rain amount (by more than 50 %). The calibration of the <span class="hlt">radar</span> receiver is operationally monitored by analyzing the received power from the sun. This turns out to cause a 1 dB underestimation. The operational clutter filter applied by KNMI is found to incorrectly identify precipitation as clutter, especially at near-zero Doppler velocities. An alternative simple clutter removal scheme using a clear sky clutter map improves the rainfall estimation slightly. To investigate the effect of wet-radome attenuation, stable returns from buildings close to the <span class="hlt">radar</span> are analyzed. It is shown that this may have caused an underestimation of up to 4 dB. Finally, a disdrometer is used to derive event and intra-event specific Z-R relations due to variations in the observed DSDs. Such variations may result in errors when applying the operational Marshall-Palmer Z-R relation. Correcting for all of these effects has a large positive impact on the <span class="hlt">radar</span>-derived precipitation estimates and yields a good match between <span class="hlt">radar</span> QPE and gauge</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840019008','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840019008"><span>Middle Atmosphere Program. Handbook for MAP, volume 9</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bowhill, S. A. (Editor); Edwards, B. (Editor)</p> <p>1983-01-01</p> <p>The term Mesosphere-Stratosphere-Troposphere <span class="hlt">radar</span> (MST) was invented to describe the use of a high power <span class="hlt">radar</span> transmitter together with a large <span class="hlt">vertically</span>, or near <span class="hlt">vertically</span>, <span class="hlt">pointing</span> antenna to study the dynamics and structure of the atmosphere from about 10 to 100 km, using the very weak coherently scattered radiation returned from small scale irregularities in refractive index. Nine topics were addressed including: meteorological and dynamic requirements for MST <span class="hlt">radar</span> networks; interpretation of <span class="hlt">radar</span> returns for clear air; techniques for the measurement of horizontal and <span class="hlt">vertical</span> velocities; techniques for studying gravity waves and turbulence; capabilities and limitations of existing MST <span class="hlt">radar</span>; design considerations for high power VHF <span class="hlt">radar</span> transceivers; optimum <span class="hlt">radar</span> antenna configurations; and data analysis techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT.......303S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT.......303S"><span>Fpga based L-band pulse doppler <span class="hlt">radar</span> design and implementation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Savci, Kubilay</p> <p></p> <p> <span class="hlt">point</span> arithmetic operations as it is fast and facilitates source requirement as it consumes less hardware than floating <span class="hlt">point</span> arithmetic operations. The software uses floating <span class="hlt">point</span> arithmetic operations, which ensure precision in processing at the expense of speed. The functionality of the <span class="hlt">radar</span> system has been tested for experimental validation in the field with a moving car and the validation of submodules are tested with synthetic data simulated on MATLAB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts068-s-053.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts068-s-053.html"><span>STS-68 <span class="hlt">radar</span> image: Mt. Pinatubo, Philippines</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1994-10-07</p> <p>STS068-S-053 (7 October 1994) --- These are color composite <span class="hlt">radar</span> images showing the area around Mount Pinatubo in the Philippines. The images were acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the Space Shuttle Endeavour on April 14, 1994 (left image) and October 5, 1994 (right image). The images are centered at about 15 degrees north latitude and 120.5 degrees east longitude. Both images were obtained with the same viewing geometry. The color composites were made by displaying the L-Band (horizontally transmitted and received) in red; the L-Band (horizontally transmitted and <span class="hlt">vertically</span> received) in green; and the C-Band (horizontally transmitted and <span class="hlt">vertically</span> received) in blue. The area shown is approximately 40 by 65 kilometers (25 by 40 miles). The main volcanic crater on Mount Pinatubo produced by the June 1991 eruptions and the steep slopes on the upper flanks of the volcano are easily seen in these images. Red on the high slopes shows the distribution of the ash deposited during the 1991 eruption, which appears red because of the low cross-polarized <span class="hlt">radar</span> returns at C and L Bands. The dark drainage's radiating away from the summit are smooth mud flows, which even three years after the eruption continue to flood the river valleys after heavy rain. Comparing the two images shows that significant changes have occurred in the intervening five months along the Pasig-Potero rivers (the dark area in the lower right of the images). Mud flows, called "lahars", that occurred during the 1994 monsoon season filled the river valleys, allowing the lahars to spread over the surrounding countryside. Three weeks before the second image was obtained, devastating lahars more than doubled the area affected in the Pasig-Potero rivers, which is clearly visible as the increase in dark area on the lower right of the images. Migration of deposition to the east (right) has affected many communities. Newly affected areas included the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01708&hterms=Charles+Darwin&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DCharles%2BDarwin','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01708&hterms=Charles+Darwin&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DCharles%2BDarwin"><span>Space <span class="hlt">Radar</span> Image Isla Isabela in 3-D</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1999-01-01</p> <p>This is a three-dimensional view of Isabela, one of the Galapagos Islands located off the western coast of Ecuador, South America. This view was constructed by overlaying a Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) image on a digital elevation map produced by TOPSAR, a prototype airborne interferometric <span class="hlt">radar</span> which produces simultaneous image and elevation data. The <span class="hlt">vertical</span> scale in this image is exaggerated by a factor of 1.87. The SIR-C/X-SAR image was taken on the 40th orbit of space shuttle Endeavour. The image is centered at about 0.5 degree south latitude and 91 degrees west longitude and covers an area of 75 by 60 kilometers (47 by 37 miles). The <span class="hlt">radar</span> incidence angle at the center of the image is about 20 degrees. The western Galapagos Islands, which lie about 1,200 kilometers (750 miles)west of Ecuador in the eastern Pacific, have six active volcanoes similar to the volcanoes found in Hawaii and reflect the volcanic processes that occur where the ocean floor is created. Since the time of Charles Darwin's visit to the area in 1835, there have been more than 60 recorded eruptions on these volcanoes. This SIR-C/X-SAR image of Alcedo and Sierra Negra volcanoes shows the rougher lava flows as bright features, while ash deposits and smooth pahoehoe lava flows appear dark. <span class="hlt">Vertical</span> exaggeration of relief is a common tool scientists use to detect relationships between structure (for example, faults, and fractures) and topography. Spaceborne Imaging <span class="hlt">Radar</span>-C and X-Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The <span class="hlt">radars</span> illuminate Earth with microwaves allowing detailed observations at any time, regardless of weather or sunlight conditions. SIR-C/X-SAR uses three microwave wavelengths: L-band (24 cm), C-band (6 cm) and X-band (3 cm). The multi-frequency data will be used by the international scientific community to better understand the global environment and how it is changing. The SIR-C/X-SAR data</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01825&hterms=dry+eyes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Ddry%2Beyes','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01825&hterms=dry+eyes&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Ddry%2Beyes"><span>Space <span class="hlt">Radar</span> Image of Safsaf Oasis, Egypt</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This three-frequency space <span class="hlt">radar</span> image of south-central Egypt demonstrates the unique capability of imaging <span class="hlt">radar</span> to penetrate thin sand cover in arid regions to reveal hidden details below the surface. Nearly all of the structures seen in this image are invisible to the naked eye and to conventional optical satellite sensors. Features appear in various colors because the three separate <span class="hlt">radar</span> wavelengths are able to penetrate the sand to different depths. Areas that appear red or orange are places that can be seen only by the longest wavelength, L-band, and they are the deepest of the buried structures. Field studies in this area indicate L-band can penetrate as much as 2 meters (6.5 feet) of very dry sand to image buried rock structures. Ancient drainage channels at the bottom of the image are filled with sand more than 2 meters (6.5 feet) thick and therefore appear dark because the <span class="hlt">radar</span> waves cannot penetrate them. The fractured orange areas at the top of the image and the blue circular structures in the center of the image are granitic areas that may contain mineral ore deposits. Scientists are using the penetrating capabilities of <span class="hlt">radar</span> imaging in desert areas in studies of structural geology, mineral exploration, ancient climates, water resources and archaeology. This image is 51.9 kilometers by 30.2 kilometers (32.2 miles by 18.7 miles) and is centered at 22.7 degrees north latitude, 29.3degrees east longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is C-band, horizontally transmitted and received; and blue is X-band, <span class="hlt">vertically</span> transmitted and received. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) on April 16, 1994, on board the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19740024146','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19740024146"><span><span class="hlt">Radar</span> studies of the planets. [<span class="hlt">radar</span> measurements of lunar surface, Mars, Mercury, and Venus</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ingalls, R. P.; Pettengill, G. H.; Rogers, A. E. E.; Sebring, P. B. (Editor); Shapiro, I. I.</p> <p>1974-01-01</p> <p>The <span class="hlt">radar</span> measurements phase of the lunar studies involving reflectivity and topographic mapping of the visible lunar surface was ended in December 1972, but studies of the data and production of maps have continued. This work was supported by Manned Spacecraft Center, Houston. Topographic mapping of the equatorial regions of Mars has been carried out during the period of each opposition since that of 1967. The method comprised extended precise traveling time measurements to a small area centered on the subradar <span class="hlt">point</span>. As measurements continued, planetary motions caused this <span class="hlt">point</span> to sweep out extensive areas in both latitude and longitude permitting the development of a fairly extensive topographical map in the equatorial region. <span class="hlt">Radar</span> observations of Mercury and Venus have also been made over the past few years. Refinements of planetary motions, reflectivity maps and determinations of rotation rates have resulted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19980045337&hterms=InSAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DInSAR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19980045337&hterms=InSAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DInSAR"><span>The Information Content of Interferometric Synthetic Aperture <span class="hlt">Radar</span>: Vegetation and Underlying Surface Topography</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Treuhaft, Robert N.</p> <p>1996-01-01</p> <p>This paper first gives a heuristic description of the sensitivity of Interferometric Synthetic Aperture <span class="hlt">Radar</span> to <span class="hlt">vertical</span> vegetation distributions and underlying surface topography. A parameter estimation scenario is then described in which the Interferometric Synthetic Aperture <span class="hlt">Radar</span> cross-correlation amplitude and phase are the observations from which vegetation and surface topographic parameters are estimated. It is shown that, even in the homogeneous-layer model of the vegetation, the number of parameters needed to describe the vegetation and underlying topography exceeds the number of Interferometric Synthetic Aperture <span class="hlt">Radar</span> observations for single-baseline, single-frequency, single-incidence-angle, single-polarization Interferometric Synthetic Aperture <span class="hlt">Radar</span>. Using ancillary ground-truth data to compensate for the underdetermination of the parameters, forest depths are estimated from the INSAR data. A recently-analyzed multibaseline data set is also discussed and the potential for stand-alone Interferometric Synthetic Aperture <span class="hlt">Radar</span> parameter estimation is assessed. The potential of combining the information content of Interferometric Synthetic Aperture <span class="hlt">Radar</span> with that of infrared/optical remote sensing data is briefly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20080039627&hterms=Ultra+wideband&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DUltra%2Bwideband','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20080039627&hterms=Ultra+wideband&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DUltra%2Bwideband"><span>Ultra-Wideband <span class="hlt">Radar</span> Measurements of Thickness of Snow Over Sea Ice</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Kanagaratnam, P.; Markus, T.; Lytle, V.; Heavey, B.; Jansen, P.; Prescott, G.; Gogineni, S.</p> <p>2007-01-01</p> <p>An accurate knowledge of snow thickness and its variability over sea ice is crucial for determining the overall polar heat and freshwater budget, which influences the global climate. Recently, algorithms have been developed to extract snow thicknesses from passive microwave satellite data. However, validation of these data over the large footprint of the passive microwave sensor has been a challenge. The only method used thus far has been with meter sticks during ship cruises. To address this problem, we developed an ultra wideband frequency-modulated continuous-wave (FM-CW) <span class="hlt">radar</span> to measure snow thickness over sea ice. We made snow-thickness measurements over Antarctic sea ice by operating the <span class="hlt">radar</span> from a sled during September and October, 2003. We performed <span class="hlt">radar</span> measurements over 11 stations with varying snow thickness between 4 and 85 cm. We observed excellent agreement between <span class="hlt">radar</span> estimates of snow thickness with physical measurements, achieving a correlation coefficient of 0.95 and a <span class="hlt">vertical</span> resolution of about 3 cm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70185076','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70185076"><span>Evaluation of landslide hazards with ground-penetrating <span class="hlt">radar</span>, Lake Michigan coast</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Barnhardt, Walter A.; Jaffe, Bruce E.; Kayen, Robert</p> <p>1999-01-01</p> <p>Ground-penetrating <span class="hlt">radar</span> (GPR) and boreholes were used to investigate a landslide-prone bluff at Sleeping Bear Dunes National Lakeshore on the northeastern coast of Lake Michigan. Based on borehole observations, sediment underlying the area is homogeneous, consisting of well-sorted, medium to coarse sand. GPR penetrated up to 20 m deep in these sediments, revealing the late Quaternary stratigraphy in great detail. We define four units, or <span class="hlt">radar</span> facies, based on criteria similar to those used in seismic stratigraphy. Directly beneath a landslide at Sleeping Bear <span class="hlt">Point</span> (and nowhere else in this survey) is a deeply incised, channel-fill deposit that intersects the shoreline at a high angle. The buried channel is at least 10 m deep and 400 m wide, and it might be a subglacially carved feature of Pleistocene age. A prominent, planar unconformity marks the upper surface of the channel deposit, which is overlain by stratified beach and dune material. Several crosshole GPR surveys were performed in the vicinity of the landslide: 1) a constant offset profile (COP), 2) a multiple offset gather (MOG), and 3) a <span class="hlt">vertical</span> <span class="hlt">radar</span> profile (VRP). Tomographic analysis of these data determined the velocity structure of sandy sediment that underlie the failed bluff. Because GPR velocity is dependent on electrical properties, we use it as a proxy for geotechnical properties of the soils. Our working hypothesis is that the hidden channel may act as a conduit for pore water flow between upland regions and Lake Michigan, and thereby locally reduce soil strength and promote slope failure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20150006554&hterms=TYPES+RADAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DTYPES%2BOF%2BRADAR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150006554&hterms=TYPES+RADAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3DTYPES%2BOF%2BRADAR"><span><span class="hlt">Radar</span> and Lidar <span class="hlt">Radar</span> DEM</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Liskovich, Diana; Simard, Marc</p> <p>2011-01-01</p> <p>Using <span class="hlt">radar</span> and lidar data, the aim is to improve 3D rendering of terrain, including digital elevation models (DEM) and estimates of vegetation height and biomass in a variety of forest types and terrains. The 3D mapping of vegetation structure and the analysis are useful to determine the role of forest in climate change (carbon cycle), in providing habitat and as a provider of socio-economic services. This in turn will lead to potential for development of more effective land-use management. The first part of the project was to characterize the Shuttle <span class="hlt">Radar</span> Topography Mission DEM error with respect to ICESat/GLAS <span class="hlt">point</span> estimates of elevation. We investigated potential trends with latitude, canopy height, signal to noise ratio (SNR), number of LiDAR waveform peaks, and maximum peak width. Scatter plots were produced for each variable and were fitted with 1st and 2nd degree polynomials. Higher order trends were visually inspected through filtering with a mean and median filter. We also assessed trends in the DEM error variance. Finally, a map showing how DEM error was geographically distributed globally was created.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20070038285&hterms=UAV&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DUAV','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20070038285&hterms=UAV&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DUAV"><span>X-Band <span class="hlt">Radar</span> for Studies of Tropical Storms from High Altitude UAV Platform</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rodriquez, Shannon; Heymsfield, Gerald; Li, Lihua; Bradley, Damon</p> <p>2007-01-01</p> <p>The increased role of unmanned aerial vehicles (UAV) in NASA's suborbital program has created a strong interest in the development of instruments with new capabilities, more compact sizes and reduced weights than the instruments currently operated on manned aircrafts. There is a strong demand and tremendous potential for using high altitude UAV (HUAV) to carry weather <span class="hlt">radars</span> for measurements of reflectivity and wind fields from tropical storms. Tropical storm genesis frequently occurs in ocean regions that are inaccessible to piloted aircraft due to the long off shore range and the required periods of time to gather significant data. Important factors of interest for the study of hurricane genesis include surface winds, profiled winds, sea surface temperatures, precipitation, and boundary layer conditions. Current satellite precipitation and surface wind sensors have resolutions that are too large and revisit times that are too infrequent to study this problem. Furthermore, none of the spaceborne sensors measure winds within the storm itself. A dual beam X-band Doppler <span class="hlt">radar</span>, UAV <span class="hlt">Radar</span> (URAD), is under development at the NASA Goddard Space Flight Center for the study of tropical storms from HUAV platforms, such as a Global Hawk. X-band is the most desirable frequency for airborne weather <span class="hlt">radars</span> since these can be built in a relatively compact size using off-the-shelf components which cost significantly less than other higher frequency <span class="hlt">radars</span>. Furthermore, X-band <span class="hlt">radars</span> provide good sensitivity with tolerable attenuation in storms. The low-cost and light-weight URAD will provide new capabilities for studying hurricane genesis by analyzing the <span class="hlt">vertical</span> structure of tropical cyclones as well as 3D reflectivity and wind fields in clouds. It will enable us to measure both the 3D precipitation structure and surface winds by using two antenna beams: fixed nadir and conical scanning each produced by its associated subsystem. The nadir subsystem is a magnetron based <span class="hlt">radar</span></p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990radr.conf...96S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990radr.conf...96S"><span>Software development for airborne <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sundstrom, Ingvar G.</p> <p></p> <p>Some aspects for development of software in a modern multimode airborne nose <span class="hlt">radar</span> are described. First, an overview of where software is used in the <span class="hlt">radar</span> units is presented. The development phases-system design, functional design, detailed design, function verification, and system verification-are then used as the starting <span class="hlt">point</span> for the discussion. Methods, tools, and the most important documents are described. The importance of video flight recording in the early stages and use of a digital signal generators for performance verification is emphasized. Some future trends are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100042295','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100042295"><span>High-resolution three-dimensional imaging <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cooper, Ken B. (Inventor); Chattopadhyay, Goutam (Inventor); Siegel, Peter H. (Inventor); Dengler, Robert J. (Inventor); Schlecht, Erich T. (Inventor); Mehdi, Imran (Inventor); Skalare, Anders J. (Inventor)</p> <p>2010-01-01</p> <p>A three-dimensional imaging <span class="hlt">radar</span> operating at high frequency e.g., 670 GHz, is disclosed. The active target illumination inherent in <span class="hlt">radar</span> solves the problem of low signal power and narrow-band detection by using submillimeter heterodyne mixer receivers. A submillimeter imaging <span class="hlt">radar</span> may use low phase-noise synthesizers and a fast chirper to generate a frequency-modulated continuous-wave (FMCW) waveform. Three-dimensional images are generated through range information derived for each pixel scanned over a target. A peak finding algorithm may be used in processing for each pixel to differentiate material layers of the target. Improved focusing is achieved through a compensation signal sampled from a <span class="hlt">point</span> source calibration target and applied to received signals from active targets prior to FFT-based range compression to extract and display high-resolution target images. Such an imaging <span class="hlt">radar</span> has particular application in detecting concealed weapons or contraband.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01304&hterms=Hawaii+Kilauea+volcano&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DHawaii%2BKilauea%2Bvolcano','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01304&hterms=Hawaii+Kilauea+volcano&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DHawaii%2BKilauea%2Bvolcano"><span>Space <span class="hlt">radar</span> image of Mauna Loa, Hawaii</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1995-01-01</p> <p>This image of the Mauna Loa volcano on the Big Island of Hawaii shows the capability of imaging <span class="hlt">radar</span> to map lava flows and other volcanic structures. Mauna Loa has erupted more than 35 times since the island was first visited by westerners in the early 1800s. The large summit crater, called Mokuaweoweo Caldera, is clearly visible near the center of the image. Leading away from the caldera (towards top right and lower center) are the two main rift zones shown here in orange. Rift zones are areas of weakness within the upper part of the volcano that are often ripped open as new magma (molten rock) approaches the surface at the start of an eruption. The most recent eruption of Mauna Loa was in March and April 1984, when segments of the northeast rift zones were active. If the height of the volcano was measured from its base on the ocean floor instead of from sea level, Mauna Loa would be the tallest mountain on Earth. Its peak (center of the image) rises more than 8 kilometers (5 miles) above the ocean floor. The South Kona District, known for cultivation of macadamia nuts and coffee, can be seen in the lower left as white and blue areas along the coast. North is toward the upper left. The area shown is 41.5 by 75 kilometers (25.7 by 46.5 miles), centered at 19.5 degrees north latitude and 155.6 degrees west longitude. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/ X-SAR) aboard the space shuttle Endeavour on its 36th orbit on October 2, 1994. The <span class="hlt">radar</span> illumination is from the left of the image. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted, <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted, <span class="hlt">vertically</span> received). The resulting color combinations in this <span class="hlt">radar</span> image are caused by differences in surface roughness of the lava flows. Smoother flows</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA007765','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA007765"><span>On the Statistical Analysis of the <span class="hlt">Radar</span> Signature of the MQM-34D</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1975-01-31</p> <p>target drone for aspect angles near normal to the roll axis for a <span class="hlt">vertically</span> polarized measurements system. The <span class="hlt">radar</span> cross section and glint are... drone . The raw data from RATSCAT are reported in graphical form in an AFSWC three-volume report.. The results reported here are a statistical analysis of...Ta1get Drones , AFSWC-rR.74-0l, January 1974. 2James W. Wright, On the Statistical Analysis of the <span class="hlt">Radar</span> Signature of the MQM-34D, Interim Report</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01781.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01781.html"><span>Space <span class="hlt">Radar</span> Image of San Rafael Glacier, Chile</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>A NASA <span class="hlt">radar</span> instrument has been successfully used to measure some of the fastest moving and most inaccessible glaciers in the world -- in Chile's huge, remote Patagonia ice fields -- demonstrating a technique that could produce more accurate predictions of glacial response to climate change and corresponding sea level changes. This image, produced with interferometric measurements made by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) flown on the Space Shuttle last fall, has provided the first detailed measurements of the mass and motion of the San Rafael Glacier. Very few measurements have been made of the Patagonian ice fields, which are the world's largest mid-latitude ice masses and account for more than 60 percent of the Southern Hemisphere's glacial area outside of Antarctica. These features make the area essential for climatologists attempting to understand the response of glaciers on a global scale to changes in climate, but the region's inaccessibility and inhospitable climate have made it nearly impossible for scientists to study its glacial topography, meteorology and changes over time. Currently, topographic data exist for only a few glaciers while no data exist for the vast interior of the ice fields. Velocity has been measured on only five of the more than 100 glaciers, and the data consist of only a few single-<span class="hlt">point</span> measurements. The interferometry performed by the SIR-C/X-SAR was used to generate both a digital elevation model of the glaciers and a map of their ice motion on a pixel-per-pixel basis at very high resolution for the first time. The data were acquired from nearly the same position in space on October 9, 10 and 11, 1994, at L-band frequency (24-cm wavelength), <span class="hlt">vertically</span> transmitted and received polarization, as the Space Shuttle Endeavor flew over several Patagonian outlet glaciers of the San Rafael Laguna. The area shown in these two images is 50 kilometers by 30 kilometers (30 miles by 18 miles) in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890005108','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890005108"><span>Investigation of <span class="hlt">radar</span> backscattering from second-year sea ice</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lei, Guang-Tsai; Moore, Richard K.; Gogineni, S. P.</p> <p>1988-01-01</p> <p>The scattering properties of second-year ice were studied in an experiment at Mould Bay in April 1983. <span class="hlt">Radar</span> backscattering measurements were made at frequencies of 5.2, 9.6, 13.6, and 16.6 GHz for <span class="hlt">vertical</span> polarization, horizontal polarization and cross polarizations, with incidence angles ranging from 15 to 70 deg. The results indicate that the second-year ice scattering characteristics were different from first-year ice and also different from multiyear ice. The fading properties of <span class="hlt">radar</span> signals were studied and compared with experimental data. The influence of snow cover on sea ice can be evaluated by accounting for the increase in the number of independent samples from snow volume with respect to that for bare ice surface. A technique for calculating the snow depth was established by this principle and a reasonable agreement has been observed. It appears that this is a usable way to measure depth in snow or other snow-like media using <span class="hlt">radar</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405976-from-utilization-point-view-two-approaches-seem-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-point-barrow-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405976-from-utilization-point-view-two-approaches-seem-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-point-barrow-radar-installation-alaska-final-report"><span>From the utilization <span class="hlt">point</span> of view, the two approaches seem to United States Air Force 611th Air Support Group/Civil Engineering Squadron, Elmendorf AFB, Alaska. Remedial investigation and feasibility study <span class="hlt">Point</span> Barrow <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-02-19</p> <p>This report presents the findings of Remedial Investigations and Feasibility Studies at sites located at the <span class="hlt">Point</span> Barrow <span class="hlt">radar</span> installation in northern Alaska. The sites were characterized based on sampling and analyses conducted during Remedial Investigation activities performed during August and September 1993.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990064210&hterms=physical+activity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dphysical%2Bactivity','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990064210&hterms=physical+activity&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dphysical%2Bactivity"><span>Comparisons of the <span class="hlt">Vertical</span> Development of Deep Tropical Convection and Associated Lightning Activity on a Global Basis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Williams, E.; Lin, S.; Labrada, C.; Christian, H.; Goodman, S.; Boccippio, D.; Driscoll, K.</p> <p>1999-01-01</p> <p>Simultaneous <span class="hlt">radar</span> (13.8 Ghz) and lightning (Lightning Imaging Sensor) observations from the NASA TRMM (Tropical Rainfall Measuring Mission) spacecraft afford a new opportunity to examine differences in tropical continental and oceanic convection on a global basis, The 250 meter <span class="hlt">vertical</span> resolution of the <span class="hlt">radar</span> data and the approximately 17 dBZ sensitivity are well suited to providing <span class="hlt">vertical</span> profiles of <span class="hlt">radar</span> reflectivity over the entire tropical belt. The reflectivity profile has been shown in numerous local ground-based studies to be a good indicator of both updraft velocity and electrical activity. The <span class="hlt">radar</span> and lightning observations for multiple satellite orbits have been integrated to produce global CAPPI's for various altitudes. At 7 km altitude, where mixed phase microphysics is known to be active, the mean reflectivity in continental convection is 10-15 dB greater than the value in oceanic convection. These results provide a sound physical basis for the order-of-magnitude contrast in lightning counts between continental and oceanic convection. These observations still beg the question, however, about the contrast in updraft velocity in these distinct convective regimes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRD..123.2797O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRD..123.2797O"><span>Toward Exploring the Synergy Between Cloud <span class="hlt">Radar</span> Polarimetry and Doppler Spectral Analysis in Deep Cold Precipitating Systems in the Arctic</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Oue, Mariko; Kollias, Pavlos; Ryzhkov, Alexander; Luke, Edward P.</p> <p>2018-03-01</p> <p>The study of Arctic ice and mixed-phase clouds, which are characterized by a variety of ice particle types in the same cloudy volume, is challenging research. This study illustrates a new approach to qualitative and quantitative analysis of the complexity of ice and mixed-phase microphysical processes in Arctic deep precipitating systems using the combination of Ka-band zenith-<span class="hlt">pointing</span> <span class="hlt">radar</span> Doppler spectra and quasi-<span class="hlt">vertical</span> profiles of polarimetric <span class="hlt">radar</span> variables measured by a Ka/W-band scanning <span class="hlt">radar</span>. The results illustrate the frequent occurrence of multimodal Doppler spectra in the dendritic/planar growth layer, where locally generated, slower-falling particle populations are well separated from faster-falling populations in terms of Doppler velocity. The slower-falling particle populations contribute to an increase of differential reflectivity (ZDR), while an enhanced specific differential phase (KDP) in this dendritic growth temperature range is caused by both the slower and faster-falling particle populations. Another area with frequent occurrence of multimodal Doppler spectra is in mixed-phase layers, where both populations produce ZDR and KDP values close to 0, suggesting the occurrence of a riming process. Joint analysis of the Doppler spectra and the polarimetric <span class="hlt">radar</span> variables provides important insight into the microphysics of snow formation and allows the separation of the contributions of ice of different habits to the values of reflectivity and ZDR.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3867359','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3867359"><span>Characteristics and Drivers of High-Altitude Ladybird Flight: Insights from <span class="hlt">Vertical</span>-Looking Entomological <span class="hlt">Radar</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>Jeffries, Daniel L.; Chapman, Jason; Roy, Helen E.; Humphries, Stuart; Harrington, Richard; Brown, Peter M. J.; Handley, Lori-J. Lawson</p> <p>2013-01-01</p> <p>Understanding the characteristics and drivers of dispersal is crucial for predicting population dynamics, particularly in range-shifting species. Studying long-distance dispersal in insects is challenging, but recent advances in entomological <span class="hlt">radar</span> offer unique insights. We analysed 10 years of <span class="hlt">radar</span> data collected at Rothamsted Research, U.K., to investigate characteristics (altitude, speed, seasonal and annual trends) and drivers (aphid abundance, air temperature, wind speed and rainfall) of high-altitude flight of the two most abundant U.K. ladybird species (native Coccinella septempunctata and invasive Harmonia axyridis). These species cannot be distinguished in the <span class="hlt">radar</span> data since their reflectivity signals overlap, and they were therefore analysed together. However, their signals do not overlap with other, abundant insects so we are confident they constitute the overwhelming majority of the analysed data. The target species were detected up to ∼1100 m above ground level, where displacement speeds of up to ∼60 km/h were recorded, however most ladybirds were found between ∼150 and 500 m, and had a mean displacement of 30 km/h. Average flight time was estimated, using tethered flight experiments, to be 36.5 minutes, but flights of up to two hours were observed. Ladybirds are therefore potentially able to travel 18 km in a “typical” high-altitude flight, but up to 120 km if flying at higher altitudes, indicating a high capacity for long-distance dispersal. There were strong seasonal trends in ladybird abundance, with peaks corresponding to the highest temperatures of mid-summer, and warm air temperature was the key driver of ladybird flight. Climatic warming may therefore increase the potential for long-distance dispersal in these species. Low aphid abundance was a second significant factor, highlighting the important role of aphid population dynamics in ladybird dispersal. This research illustrates the utility of <span class="hlt">radar</span> for studying high</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01727&hterms=Mount+Rainier&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DMount%2BRainier','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01727&hterms=Mount+Rainier&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DMount%2BRainier"><span>Space <span class="hlt">Radar</span> Image of Mt. Rainer, Washington</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is a <span class="hlt">radar</span> image of Mount Rainier in Washington state. The volcano last erupted about 150 years ago and numerous large floods and debris flows have originated on its slopes during the last century. Today the volcano is heavily mantled with glaciers and snowfields. More than 100,000 people live on young volcanic mudflows less than 10,000 years old and, consequently, are within the range of future, devastating mudslides. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 20th orbit on October 1, 1994. The area shown in the image is approximately 59 kilometers by 60 kilometers (36.5 miles by 37 miles). North is toward the top left of the image, which was composed by assigning red and green colors to the L-band, horizontally transmitted and <span class="hlt">vertically</span>, and the L-band, horizontally transmitted and <span class="hlt">vertically</span> received. Blue indicates the C-band, horizontally transmitted and <span class="hlt">vertically</span> received. In addition to highlighting topographic slopes facing the space shuttle, SIR-C records rugged areas as brighter and smooth areas as darker. The scene was illuminated by the shuttle's <span class="hlt">radar</span> from the northwest so that northwest-facing slopes are brighter and southeast-facing slopes are dark. Forested regions are pale green in color; clear cuts and bare ground are bluish or purple; ice is dark green and white. The round cone at the center of the image is the 14,435-foot (4,399-meter) active volcano, Mount Rainier. On the lower slopes is a zone of rock ridges and rubble (purple to reddish) above coniferous forests (in yellow/green). The western boundary of Mount Rainier National Park is seen as a transition from protected, old-growth forest to heavily logged private land, a mosaic of recent clear cuts (bright purple/blue) and partially regrown timber plantations (pale blue). The prominent river seen curving away from the mountain at the top of the image (to the northwest) is the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225462p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225462p/"><span>Detail view of southeast corner of Signal Corps <span class="hlt">Radar</span> (S.C.R.) ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Detail view of southeast corner of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Transmitter Building foundation, showing Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Tower concrete pier in background, camera facing north - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1808427','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1808427"><span>Bats Avoid <span class="hlt">Radar</span> Installations: Could Electromagnetic Fields Deter Bats from Colliding with Wind Turbines?</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nicholls, Barry; Racey, Paul A.</p> <p>2007-01-01</p> <p>Large numbers of bats are killed by collisions with wind turbines, and there is at present no direct method of reducing or preventing this mortality. We therefore determine whether the electromagnetic radiation associated with <span class="hlt">radar</span> installations can elicit an aversive behavioural response in foraging bats. Four civil air traffic control (ATC) <span class="hlt">radar</span> stations, three military ATC <span class="hlt">radars</span> and three weather <span class="hlt">radars</span> were selected, each surrounded by heterogeneous habitat. Three sampling <span class="hlt">points</span> matched for habitat type and structure, dominant vegetation species, altitude and surrounding land class were located at increasing distances from each station. A portable electromagnetic field meter measured the field strength of the <span class="hlt">radar</span> at three distances from the source: in close proximity (<200 m) with a high electromagnetic field (EMF) strength >2 volts/metre, an intermediate <span class="hlt">point</span> within line of sight of the <span class="hlt">radar</span> (200–400 m) and with an EMF strength <2 v/m, and a control site out of sight of the <span class="hlt">radar</span> (>400 m) and registering an EMF of zero v/m. At each <span class="hlt">radar</span> station bat activity was recorded three times with three independent sampling <span class="hlt">points</span> monitored on each occasion, resulting in a total of 90 samples, 30 of which were obtained within each field strength category. At these sampling <span class="hlt">points</span>, bat activity was recorded using an automatic bat recording station, operated from sunset to sunrise. Bat activity was significantly reduced in habitats exposed to an EMF strength of greater than 2 v/m when compared to matched sites registering EMF levels of zero. The reduction in bat activity was not significantly different at lower levels of EMF strength within 400 m of the <span class="hlt">radar</span>. We predict that the reduction in bat activity within habitats exposed to electromagnetic radiation may be a result of thermal induction and an increased risk of hyperthermia. PMID:17372629</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005SPIE.5659...61I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005SPIE.5659...61I"><span>Instrument concepts and technologies for future spaceborne atmospheric <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Im, Eastwood; Durden, Stephen L.</p> <p>2005-01-01</p> <p>In conjunction with the implementation of spaceborne atmospheric <span class="hlt">radar</span> flight missions, NASA is developing advanced instrument concepts and technologies for future spaceborne atmospheric <span class="hlt">radars</span>, with the over-arching objectives of making such instruments more capable in supporting future science needs, and more cost effective. Two such examples are the Second-Generation Precipitation <span class="hlt">Radar</span> (PR-2) and the Nexrad-In-Space (NIS). PR-2 is a 14/35-GHz dual-frequency rain <span class="hlt">radar</span> with a deployable 5-meter, wide-swath scanned membrane antenna, a dual-polarized/dual-frequency receiver, and a real-time digital signal processor. It is intended for Low Earth Orbit (LEO) operations to provide greatly enhanced rainfall profile retrieval accuracy while using only a fraction of the mass of the current TRMM PR. NIS is designed to be a 35-GHz Geostationary Earth Orbiting (GEO) <span class="hlt">radar</span> with the intent of providing hourly monitoring of the life cycle of hurricanes and tropical storms. It uses a 35-m, spherical, lightweight membrane antenna and Doppler processing to acquire 3-dimensional information on the intensity and <span class="hlt">vertical</span> motion of hurricane rainfall. Technologies for NIS are synergistic with those for PR-2. During the last two years, several of the technology items associated with these notional instruments have also been prototyped. This paper will give an overview of these instrument design concepts and their associated technologies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19770009481','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19770009481"><span><span class="hlt">Radar</span> systems for the water resources mission. Volume 4: Appendices E-I</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Moore, R. K.; Claassen, J. P.; Erickson, R. L.; Fong, R. K. T.; Hanson, B. C.; Komen, M. J.; Mcmillan, S. B.; Parashar, S. K.</p> <p>1976-01-01</p> <p>The use of a scanning antenna beam for a synthetic aperture system was examined. When the resolution required was modest, the <span class="hlt">radar</span> did not use all the time the beam was passing a given <span class="hlt">point</span> on the ground to build a synthetic aperture, so time was available to scan the beam to other positions and build several images at different ranges. The scanning synthetic-aperture <span class="hlt">radar</span> (SCANSAR) could achieve swathwidths of well over 100 km with modest antenna size. Design considerations for a SCANSAR for hydrologic parameter observation are presented. Because of the high sensitivity to soil moisture at angles of incidence near <span class="hlt">vertical</span>, a 7 to 22 deg swath was considered for that application. For snow and ice monitoring, a 22 to 37 deg scan was used. Frequencies from X-band to L-band were used in the design studies, but the proposed system operated in C-band at 4.75 GHz. It achieved an azimuth resolution of about 50 meters at all angles, with a range resolution varying from 150 meters at 7 deg to 31 meters at 37 deg. The antenna required an aperture of 3 x 4.16 meters, and the average transmitter power was under 2 watts.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01844&hterms=cultural+different&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcultural%2Bdifferent','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01844&hterms=cultural+different&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dcultural%2Bdifferent"><span>Space <span class="hlt">Radar</span> Image of Calcutta, West Bengal, India</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image of Calcutta, India, illustrates different urban land use patterns. Calcutta, the largest city in India, is located on the banks of the Hugli River, shown as the thick, dark line in the upper portion of the image. The surrounding area is a flat swampy region with a subtropical climate. As a result of this marshy environment, Calcutta is a compact city, concentrated along the fringes of the river. The average elevation is approximately 9 meters (30 feet) above sea level. Calcutta is located 154 kilometers (96 miles) upstream from the Bay of Bengal. Central Calcutta is the light blue and orange area below the river in the center of the image. The bridge spanning the river at the city center is the Howrah Bridge which links central Calcutta to Howrah. The dark region just below the river and to the left of the city center is Maidan, a large city park housing numerous cultural and recreational facilities. The international airport is in the lower right of the image. The bridge in the upper right is the Bally Bridge which links the suburbs of Bally and Baranagar. This image is 30 kilometers by 10 kilometers (19 miles by 6 miles)and is centered at 22.3 degrees north latitude, 88.2 degrees east longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) on October 5, 1994, onboard the Space Shuttle Endeavour. SIR-C/X SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01783&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01783&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities"><span>Space <span class="hlt">Radar</span> Image of Houston, Texas</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This image of Houston, Texas, shows the amount of detail that is possible to obtain using spaceborne <span class="hlt">radar</span> imaging. Images such as this -- obtained by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) flying aboard the space shuttle Endeavor last fall -- can become an effective tool for urban planners who map and monitor land use patterns in urban, agricultural and wetland areas. Central Houston appears pink and white in the upper portion of the image, outlined and crisscrossed by freeways. The image was obtained on October 10, 1994, during the space shuttle's 167th orbit. The area shown is 100 kilometers by 60 kilometers (62 miles by 38 miles) and is centered at 29.38 degrees north latitude, 95.1 degrees west longitude. North is toward the upper left. The pink areas designate urban development while the green-and blue-patterned areas are agricultural fields. Black areas are bodies of water, including Galveston Bay along the right edge and the Gulf of Mexico at the bottom of the image. Interstate 45 runs from top to bottom through the image. The narrow island at the bottom of the image is Galveston Island, with the city of Galveston at its northeast (right) end. The dark cross in the upper center of the image is Hobby Airport. Ellington Air Force Base is visible below Hobby on the other side of Interstate 45. Clear Lake is the dark body of water in the middle right of the image. The green square just north of Clear Lake is Johnson Space Center, home of Mission Control and the astronaut training facilities. The black rectangle with a white center that appears to the left of the city center is the Houston Astrodome. The colors in this image were obtained using the follow <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted, <span class="hlt">vertically</span> received); green represents the C-band (horizontally transmitted, <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and received). Spaceborne Imaging <span class="hlt">Radar</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950005179','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950005179"><span>Report on the comparison of the scan strategies employed by the Patrick Air Force Base WSR-74C/McGill <span class="hlt">radar</span> and the NWS Melbourne WSR-88D <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Taylor, Gregory; Evans, Randolph; Manobianco, John; Schumann, Robin; Wheeler, Mark; Yersavich, Ann</p> <p>1994-01-01</p> <p>The objective of this investigation is to determine whether the current standard WSR-88D <span class="hlt">radar</span> (NEXRAD) scan strategies permit the use of the Melbourne WSR-88D to perform the essential functions now performed by the Patrick Air Force Base (PAFB) WSR-74C/McGill <span class="hlt">radar</span> for evaluating shuttle weather flight rules (FR) and launch commit criteria (LCC). To meet this objective, the investigation compared the beam coverage patterns of the WSR-74C/McGill <span class="hlt">radar</span> located at PAFB and the WSR-88D <span class="hlt">radar</span> located at the Melbourne National Weather Service (NWS) Office over the area of concern for weather FR and LCC evaluations. The analysis focused on beam coverage within four <span class="hlt">vertical</span> 74 km radius cylinders (1 to 4 km above ground level (AGL), 4 to 8 km AGL, 8 to 12 km AGL, and 1 to 12 km AGL) centered on Kennedy Space Center (KSC) Launch Complex 39A. The PAFB WSR-74C/McGill <span class="hlt">radar</span> is approximately 17 km north-northeast of the Melbourne WSR-88D <span class="hlt">radar</span>. The beam coverage of the WSR-88D using VCP 11 located at the Melbourne NWS Office is comparable (difference in percent of the atmosphere sampled between the two <span class="hlt">radars</span> is 10 percent or less) within the area of concern to the beam coverage of the WSR-74C/McGill <span class="hlt">radar</span> located at PAFB. Both <span class="hlt">radars</span> provide good beam coverage over much of the atmospheric region of concern. In addition, both <span class="hlt">radars</span> provide poor beam coverage (coverage less than 50 percent) over limited regions near the <span class="hlt">radars</span> due to the <span class="hlt">radars</span>' cone of silence and gaps in coverage within the higher elevation scans. Based on scan strategy alone, the WSR-88D <span class="hlt">radar</span> could be used to perform the essential functions now performed by the PAFB WSR-74C/McGill <span class="hlt">radar</span> for evaluating shuttle weather FR and LCC. Other <span class="hlt">radar</span> characteristics may, however, affect the decision as to which <span class="hlt">radar</span> to use in a given case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20180000191','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20180000191"><span>Remote Sensing of Precipitation from Airborne and Spaceborne <span class="hlt">Radar</span>. Chapter 13</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Munchak, S. Joseph</p> <p>2017-01-01</p> <p>Weather <span class="hlt">radar</span> measurements from airborne or satellite platforms can be an effective remote sensing tool for examining the three-dimensional structures of clouds and precipitation. This chapter describes some fundamental properties of <span class="hlt">radar</span> measurements and their dependence on the particle size distribution (PSD) and <span class="hlt">radar</span> frequency. The inverse problem of solving for the <span class="hlt">vertical</span> profile of PSD from a profile of measured reflectivity is stated as an optimal estimation problem for single- and multi-frequency measurements. Phenomena that can change the measured reflectivity Z(sub m) from its intrinsic value Z(sub e), namely attenuation, non-uniform beam filling, and multiple scattering, are described and mitigation of these effects in the context of the optimal estimation framework is discussed. Finally, some techniques involving the use of passive microwave measurements to further constrain the retrieval of the PSD are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980021318','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980021318"><span>Three-Centimeter Doppler <span class="hlt">Radar</span> Observations of Wingtip-Generated Wake Vortices in Clear Air</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Marshall, Robert E.; Mudukutore, Ashok; Wissel, Vicki L. H.; Myers, Theodore</p> <p>1997-01-01</p> <p>This report documents a high risk, high pay-off experiment with the objective of detecting, for the first time, the presence of aircraft wake vortices in clear air using X-band Doppler <span class="hlt">radar</span>. Field experiments were conducted in January 1995 at the Wallops Flight Facility (WFF) to demonstrate the capability of the 9.33 GHz (I=3 cm) <span class="hlt">radar</span>, which was assembled using an existing nine-meter parabolic antenna reflector at VVTT and the receiver/transmitter from the NASA Airborne Windshear <span class="hlt">Radar</span>-Program. A C-130-aircraft, equipped with wingtip smoke generators, created visually marked wake vortices, which were recorded by video cameras. A C-band <span class="hlt">radar</span> also observed the wake vortices during detection attempts with the X-band <span class="hlt">radar</span>. Rawinsonde data was used to calculate <span class="hlt">vertical</span> soundings of wake vortex decay time, cross aircraft bearing wind speed, and water vapor mixing ratio for aircraft passes over the <span class="hlt">radar</span> measurement range. This experiment was a pathfinder in predicting, in real time, the location and persistence of C-130 vortices, and in setting the flight path of the aircraft to optimize X-band <span class="hlt">radar</span> measurement of the wake vortex core in real time. This experiment was conducted in support of the NASA Aircraft Vortex Spacing System (AVOSS).</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/2010JGRD..11519116L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JGRD..11519116L"><span>Observations of Kelvin-Helmholtz instability at a cloud base with the middle and upper atmosphere (MU) and weather <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Luce, Hubert; Mega, Tomoaki; Yamamoto, Masayuki K.; Yamamoto, Mamoru; Hashiguchi, Hiroyuki; Fukao, Shoichiro; Nishi, Noriyuki; Tajiri, Takuya; Nakazato, Masahisa</p> <p>2010-10-01</p> <p>Using the very high frequency (46.5 MHz) middle and upper atmosphere <span class="hlt">radar</span> (MUR), Ka band (35 GHz) and X band (9.8 GHz) weather <span class="hlt">radars</span>, a Kelvin-Helmholtz (KH) instability occurring at a cloud base and its impact on modulating cloud bottom altitudes are described by a case study on 8 October 2008 at the Shigaraki MU Observatory, Japan (34.85°N, 136.10°E). KH braids were monitored by the MUR along the slope of a cloud base gradually rising with time around an altitude of ˜5.0 km. The KH braids had a horizontal wavelength of about 3.6 km and maximum crest-to-trough amplitude of about 1.6 km. Nearly monochromatic and out of phase <span class="hlt">vertical</span> air motion oscillations exceeding ±3 m s-1 with a period of ˜3 min 20 s were measured by the MUR above and below the cloud base. The axes of the billows were at right angles of the wind and wind shear both oriented east-north-east at their altitude. The isotropy of the <span class="hlt">radar</span> echoes and the large variance of Doppler velocity in the KH billows (including the braids) indicate the presence of strong turbulence at the Bragg (˜3.2 m) scale. After the passage of the cloud system, the KH waves rapidly damped and the <span class="hlt">vertical</span> scale of the KH braids progressively decreased down to about 100 m before their disappearance. The <span class="hlt">radar</span> observations suggest that the interface between clear air and cloud was conducive to the presence of the dynamical shear instability by reducing static stability (and then the Richardson number) near the cloud base. Downward cloudy protuberances detected by the Ka band <span class="hlt">radar</span> had <span class="hlt">vertical</span> and horizontal scales of about 0.6-1.1 and 3.2 km, respectively, and were clearly associated with the downward air motions. Observed oscillations of the reflectivity-weighted Doppler velocity measured by the X band <span class="hlt">radar</span> indicate that falling ice particles underwent the <span class="hlt">vertical</span> wind motions generated by the KH instability to form the protuberances. The protuberances at the cloud base might be either KH billow clouds or perhaps</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020013938','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020013938"><span>Measurement of Attenuation with Airborne and Ground-Based <span class="hlt">Radar</span> in Convective Storms Over Land Its Microphysical Implications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tian, Lin; Heymsfield, G. M.; Srivastava, R. C.; O'C.Starr, D. (Technical Monitor)</p> <p>2001-01-01</p> <p>Observations by the airborne X-band Doppler <span class="hlt">radar</span> (EDOP) and the NCAR S-band polarimetric (S-Pol) <span class="hlt">radar</span> from two field experiments are used to evaluate the surface reference technique (SRT) for measuring the path integrated attenuation (PIA) and to study attenuation in deep convective storms. The EDOP, flying at an altitude of 20 km, uses a nadir beam and a forward <span class="hlt">pointing</span> beam. It is found that over land, the surface scattering cross-section is highly variable at nadir incidence but relatively stable at forward incidence. It is concluded that measurement by the forward beam provides a viable technique for measuring PIA using the SRT. <span class="hlt">Vertical</span> profiles of peak attenuation coefficient are derived in two deep convective storms by the dual-wavelength method. Using the measured Doppler velocity, the reflectivities at the two wavelengths, the differential reflectivity and the estimated attenuation coefficients, it is shown that: supercooled drops and (dry) ice particles probably co-existed above the melting level in regions of updraft, that water-coated partially melted ice particles probably contributed to high attenuation below the melting level.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225464p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225464p/"><span>Location plan for Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Location plan for Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5, October 8, 1943 - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012GeoRL..3917606M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012GeoRL..3917606M"><span>Synergistic surface current mapping by spaceborne stereo imaging and coastal HF <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Matthews, John Philip; Yoshikawa, Yutaka</p> <p>2012-09-01</p> <p>Well validated optical and <span class="hlt">radar</span> methods of surface current measurement at high spatial resolution (nominally <100 m) from space can greatly advance our ability to monitor earth's oceans, coastal zones, lakes and rivers. With interest growing in optical along-track stereo techniques for surface current and wave motion determinations, questions of how to interpret such data and how to relate them to measurements made by better validated techniques arise. Here we make the first systematic appraisal of surface currents derived from along-track stereo Sun glitter (ATSSG) imagery through comparisons with simultaneous synoptic flows observed by coastal HF <span class="hlt">radars</span> working at frequencies of 13.9 and 24.5 MHz, which return averaged currents within surface layers of roughly 1 m and 2 m depth respectively. At our Tsushima Strait (Japan) test site, we found that these two techniques provided largely compatible surface current patterns, with the main difference apparent in current strength. Within the northwest (southern) comparison region, the magnitudes of the ATSSG current vectors derived for 13 August 2006 were on average 22% (40%) higher than the corresponding vectors for the 1-m (2-m) depth <span class="hlt">radar</span>. These results reflect near-surface <span class="hlt">vertical</span> current structure, differences in the flow components sensed by the two techniques and disparities in instrumental performance. The <span class="hlt">vertical</span> profile constructed here from ATSSG, HF <span class="hlt">radar</span> and ADCP data is the first to resolve downwind drift in the upper 2 m of the open ocean. The profile e-folding depth suggests Stokes drift from waves of 10-m wavelength visible in the images.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001PhDT.........3I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001PhDT.........3I"><span>Clear-air <span class="hlt">radar</span> observations of the atmospheric boundary layer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ince, Turker</p> <p>2001-10-01</p> <p>This dissertation presents the design and operation of a high-resolution frequency-modulated continuous-wave (FM- CW) <span class="hlt">radar</span> system to study the structure and dynamics of clear-air turbulence in the atmospheric boundary layer (ABL). This sensitive <span class="hlt">radar</span> can image the <span class="hlt">vertical</span> structure of the ABL with both high spatial and temporal resolutions, and provide both qualitative information about the morphology of clear-air structures and quantitative information on the intensity of fluctuations in refractive-index of air. The principles of operation and the hardware and data acquisition characteristics of the <span class="hlt">radar</span> are described in the dissertation. In October 1999, the <span class="hlt">radar</span> participated in the Cooperative Atmosphere-Surface Exchange Study (CASES'99) Experiment to characterize the temporal structure and evolution of the boundary-layer features in both convective and stable conditions. The observed structures include clear-air convection, boundary layer evolution, gravity waves, Kelvin-Helmholtz instabilities, stably stratified layers, and clear-air turbulence. Many of the S-band <span class="hlt">radar</span> images also show high- reflectivity returns from Rayleigh scatterers such as insects. An adaptive median filtering technique based on local statistics has, therefore, been developed to discriminate between Bragg and Rayleigh scattering in clear-air <span class="hlt">radar</span> observations. The filter is tested on <span class="hlt">radar</span> observations of clear air convection with comparison to two commonly used image processing techniques. The dissertation also examines the statistical mean of the <span class="hlt">radar</span>-measured C2n for clear-air convection, and compares it with the theoretical predictions. The study also shows that the inversion height, local thickness of the inversion layer, and the height of the elevated atmospheric layers can be estimated from the <span class="hlt">radar</span> reflectivity measurements. In addition, comparisons to the radiosonde-based height estimates are made. To examine the temporal and spatial structure of C2n , the dissertation</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.B51C0040L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.B51C0040L"><span>Beyond <span class="hlt">Radar</span> Backscatter: Estimating Forest Structure and Biomass with <span class="hlt">Radar</span> Interferometry and Lidar Remote Sensing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lavalle, M.; Ahmed, R.</p> <p>2014-12-01</p> <p>Mapping forest structure and aboveground biomass globally is a major challenge that the remote sensing community has been facing for decades. <span class="hlt">Radar</span> backscatter is sensitive to biomass only up to a certain amount (about 150 tons/ha at L-band and 300 tons/ha at P-band), whereas lidar remote sensing is strongly limited by poor spatial coverage. In recent years <span class="hlt">radar</span> interferometry, including its extension to polarimetric <span class="hlt">radar</span> interferometry (PolInSAR), has emerged as a new technique to overcome the limitations of <span class="hlt">radar</span> backscatter. The idea of PolInSAR is to use jointly interferometric and polarimetric <span class="hlt">radar</span> techniques to separate different scattering mechanisms and retrieve the <span class="hlt">vertical</span> structure of forests. The advantage is to map ecosystem structure continuously over large areas and independently of cloud coverage. Experiments have shown that forest height - an important proxy for biomass - can be estimated using PolInSAR with accuracy between 15% and 20% at plot level. At AGU we will review the state-of-art of repeat-pass PolInSAR for biomass mapping, including its potential and limitations, and discuss how merging lidar data with PolInSAR data can be beneficial not only for product cross-validation but also for achieving better estimation of ecosystem properties over large areas. In particular, lidar data are expected to aid the inversion of PolInSAR models by providing (1) better identification of ground under the canopy, (2) approximate information of canopy structure in limited areas, and (3) maximum tree height useful for mapping PolInSAR temporal decorrelation. We will show our tree height and biomass maps using PolInSAR L-band JPL/UAVSAR data collected in tropical and temperate forests, and P-band ONERA/TROPISAR data acquired in French Guiana. LVIS lidar data will be used, as well as SRTM data, field measurements and inventory data to support our study. The use of two different <span class="hlt">radar</span> frequencies and repeat-pass JPL UAVSAR data will offer also the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940011418','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940011418"><span>X-SAR: The X-band synthetic aperture <span class="hlt">radar</span> on board the Space Shuttle</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Werner, Marian U.</p> <p>1993-01-01</p> <p>The X-band synthetic aperture <span class="hlt">radar</span> (X-SAR) is the German/Italian contribution to the NASA/JPL Shuttle <span class="hlt">Radar</span> Lab missions as part of the preparation for the Earth Observation System (EOS) program. The Shuttle <span class="hlt">Radar</span> Lab is a combination of several <span class="hlt">radars</span>: an L-band (1.2 GHz) and a C-band (5.3 GHz) multipolarization SAR known as SIR-C (Shuttle Imaging <span class="hlt">Radar</span>); and an X-band (9.6 GHz) <span class="hlt">vertically</span> polarized SAR which will be operated synchronously over the same target areas to deliver calibrated multifrequency and multipolarization SAR data at multiple incidence angles from space. A joint German/Italian project office at DARA (German Space Agency) is responsible for the management of the X-SAR project. The space hardware has been developed and manufactured under industrial contract by Dornier and Alenia Spazio. Besides supporting all the technical and scientific tasks, DLR, in cooperation with ASI (Agencia Spaziale Italiano) is responsible for mission operation, calibration, and high precision SAR processing. In addition, DLR developed an airborne X-band SAR to support the experimenters with campaigns to prepare for the missions. The main advantage of adding a shorter wavelength (3 cm) <span class="hlt">radar</span> to the SIR-C <span class="hlt">radars</span> is the X-band <span class="hlt">radar</span>'s weaker penetration into vegetation and soil and its high sensitivity to surface roughness and associated phenomena. The performance of each of the three <span class="hlt">radars</span> is comparable with respect to radiometric and geometric resolution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01807&hterms=monsanto&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dmonsanto','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01807&hterms=monsanto&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dmonsanto"><span>Space <span class="hlt">Radar</span> Image of Lisbon, Portugal</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image of Lisbon, Portugal illustrates the different land use patterns that are present in coastal Portugal. Lisbon, the national capital, lies on the north bank of the Rio Tejo where the river enters the Atlantic Ocean. The city center appears as the bright area in the center of the image. The green area west of the city center is a large city park called the Parque Florestal de Monsanto. The Lisbon Airport is visible east of the city. The Rio Tejo forms a large bay just east of the city. Many agricultural fields are visible as a patchwork pattern east of the bay. Suburban housing can be seen on the southern bank of the river. Spanning the river is the Ponte 25 de Abril, a large suspension bridge similar in architecture to San Francisco's Golden Gate Bridge. The image was acquired on April 19, 1994 and is centered at 38.8 degrees north latitude, 9.2 degrees west longitude. North is towards the upper right. The image is 50 kilometers by 30 kilometers (31 miles by 19 miles). The colors in this image represent the following <span class="hlt">radar</span> channels and polarizations: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian, and the United States space agencies, is part of NASA's Mission to Planet Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/AD1004163','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/AD1004163"><span>Coordinated <span class="hlt">Radar</span> Resource Management for Networked Phased Array <span class="hlt">Radars</span></span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2014-12-01</p> <p>Coordinated <span class="hlt">radar</span> resource management for networked phased array <span class="hlt">radars</span> Peter W. Moo and Zhen Ding <span class="hlt">Radar</span> Sensing & Exploitation Section Defence...15] P.W. Moo . Scheduling for multifunction <span class="hlt">radar</span> via two-slope benefit functions. <span class="hlt">Radar</span>, Sonar Navigation, IET, 5(8):884 –894, Oct. 2011. [16] M.I</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840015461&hterms=duricrust&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dduricrust','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840015461&hterms=duricrust&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dduricrust"><span>Is there <span class="hlt">radar</span> evidence for liquid water on Mars?</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Roth, L. E.</p> <p>1984-01-01</p> <p>The hypothesis that an extraordinary <span class="hlt">radar</span> smoothness of a lunar target suggests that ground moisture is rest on the assumption that on the penetration-depth scale, the dielectric constant be an isotropic quantity. In other words, the planet's surface should have no <span class="hlt">vertical</span> structure. Results of modeling exercises (based on the early lunar two-layer models) conducted to simulate the behavior of <span class="hlt">radar</span> reflectivity, at S-band, over Solis Lacus, without manipulating the dielectric constant of the base layer (i.e., without adding moisture) are summarized. More sophisticated, explicit, rather than iterative multi-layer models involving dust, duricrust, mollisol, and permafrost are under study. It is anticipated that a paradoxical situation will be reached when each improvement in the model introduces additional ambiguities into the data interpretation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2837276','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2837276"><span>Latitude and longitude <span class="hlt">vertical</span> disparity</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Read, Jenny C. A.; Phillipson, Graeme P.; Glennerster, Andrew</p> <p>2010-01-01</p> <p>The literature on <span class="hlt">vertical</span> disparity is complicated by the fact that several different definitions of the term “<span class="hlt">vertical</span> disparity” are in common use, often without a clear statement about which is intended or a widespread appreciation of the properties of the different definitions. Here, we examine two definitions of retinal <span class="hlt">vertical</span> disparity: elevation-latitude and elevation-longitude disparity. Near the fixation <span class="hlt">point</span>, these definitions become equivalent, but in general, they have quite different dependences on object distance and binocular eye posture, which have not previously been spelt out. We present analytical approximations for each type of <span class="hlt">vertical</span> disparity, valid for more general conditions than previous derivations in the literature: we do not restrict ourselves to objects near the fixation <span class="hlt">point</span> or near the plane of regard, and we allow for non-zero torsion, cyclovergence and <span class="hlt">vertical</span> misalignments of the eyes. We use these expressions to derive estimates of the latitude and longitude <span class="hlt">vertical</span> disparity expected at each <span class="hlt">point</span> in the visual field, averaged over all natural viewing. Finally, we present analytical expressions showing how binocular eye position – gaze direction, convergence, torsion, cyclovergence, and <span class="hlt">vertical</span> misalignment – can be derived from the <span class="hlt">vertical</span> disparity field and its derivatives at the fovea. PMID:20055544</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA21626.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA21626.html"><span>Cassini's Final Titan <span class="hlt">Radar</span> Swath</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>2017-08-11</p> <p>During its final targeted flyby of Titan on April 22, 2017, Cassini's <span class="hlt">radar</span> mapper got the mission's last close look at the moon's surface. On this 127th targeted pass by Titan (unintuitively named "T-126"), the <span class="hlt">radar</span> was used to take two images of the surface, shown at left and right. Both images are about 200 miles (300 kilometers) in width, from top to bottom. Objects appear bright when they are tilted toward the spacecraft or have rough surfaces; smooth areas appear dark. At left are the same bright, hilly terrains and darker plains that Cassini imaged during its first <span class="hlt">radar</span> pass of Titan, in 2004. Scientists do not see obvious evidence of changes in this terrain over the 13 years since the original observation. At right, the <span class="hlt">radar</span> looked once more for Titan's mysterious "magic island" (PIA20021) in a portion of one of the large hydrocarbon seas, Ligeia Mare. No "island" feature was observed during this pass. Scientists continue to work on what the transient feature might have been, with waves and bubbles being two possibilities. In between the two parts of its imaging observation, the <span class="hlt">radar</span> instrument switched to altimetry mode, in order to make a first-ever (and last-ever) measurement of the depths of some of the lakes that dot the north polar region. For the measurements, the spacecraft <span class="hlt">pointed</span> its antenna straight down at the surface and the <span class="hlt">radar</span> measured the time delay between echoes from the lakes' surface and bottom. A graph is available at https://photojournal.jpl.nasa.gov/catalog/PIA21626</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA594976','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA594976"><span>Compressive Sensing for <span class="hlt">Radar</span> and <span class="hlt">Radar</span> Sensor Networks</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2013-12-02</p> <p>Zero Correlation Zone Sequence Pair Sets for MIMO <span class="hlt">Radar</span> Inspired by recent advances in MIMO <span class="hlt">radar</span>, we apply orthogonal phase coded waveforms to MIMO ...<span class="hlt">radar</span> system in order to gain better range resolution and target direction finding performance [2]. We provide and investigate a generalized MIMO <span class="hlt">radar</span>...ZCZ) sequence-Pair Set (ZCZPS). We also study the MIMO <span class="hlt">radar</span> ambiguity function of the system using phase coded waveforms, based on which we analyze</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A31G2263R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A31G2263R"><span>Multiple Convective Cell Identification and Tracking Algorithm for documenting time-height evolution of measured polarimetric <span class="hlt">radar</span> and lightning properties</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rosenfeld, D.; Hu, J.; Zhang, P.; Snyder, J.; Orville, R. E.; Ryzhkov, A.; Zrnic, D.; Williams, E.; Zhang, R.</p> <p>2017-12-01</p> <p>A methodology to track the evolution of the hydrometeors and electrification of convective cells is presented and applied to various convective clouds from warm showers to super-cells. The input <span class="hlt">radar</span> data are obtained from the polarimetric NEXRAD weather <span class="hlt">radars</span>, The information on cloud electrification is obtained from Lightning Mapping Arrays (LMA). The development time and height of the hydrometeors and electrification requires tracking the evolution and lifecycle of convective cells. A new methodology for Multi-Cell Identification and Tracking (MCIT) is presented in this study. This new algorithm is applied to time series of <span class="hlt">radar</span> volume scans. A cell is defined as a local maximum in the <span class="hlt">Vertical</span> Integrated Liquid (VIL), and the echo area is divided between cells using a watershed algorithm. The tracking of the cells between <span class="hlt">radar</span> volume scans is done by identifying the two cells in consecutive <span class="hlt">radar</span> scans that have maximum common VIL. The <span class="hlt">vertical</span> profile of the polarimetric <span class="hlt">radar</span> properties are used for constructing the time-height cross section of the cell properties around the peak reflectivity as a function of height. The LMA sources that occur within the cell area are integrated as a function of height as well for each time step, as determined by the <span class="hlt">radar</span> volume scans. The result of the tracking can provide insights to the evolution of storms, hydrometer types, precipitation initiation and cloud electrification under different thermodynamic, aerosol and geographic conditions. The details of the MCIT algorithm, its products and their performance for different types of storm are described in this poster.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01765.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01765.html"><span>Space <span class="hlt">Radar</span> Image of Kiluchevskoi, Volcano, Russia</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This is an image of the area of Kliuchevskoi volcano, Kamchatka, Russia, which began to erupt on September 30, 1994. Kliuchevskoi is the blue triangular peak in the center of the image, towards the left edge of the bright red area that delineates bare snow cover. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 88th orbit on October 5, 1994. The image shows an area approximately 75 kilometers by 100 kilometers (46 miles by 62 miles) that is centered at 56.07 degrees north latitude and 160.84 degrees east longitude. North is toward the bottom of the image. The <span class="hlt">radar</span> illumination is from the top of the image. The Kamchatka volcanoes are among the most active volcanoes in the world. The volcanic zone sits above a tectonic plate boundary, where the Pacific plate is sinking beneath the northeast edge of the Eurasian plate. The Endeavour crew obtained dramatic video and photographic images of this region during the eruption, which will assist scientists in analyzing the dynamics of the recent activity. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). In addition to Kliuchevskoi, two other active volcanoes are visible in the image. Bezymianny, the circular crater above and to the right of Kliuchevskoi, contains a slowly growing lava dome. Tolbachik is the large volcano with a dark summit crater near the upper right edge of the red snow covered area. The Kamchatka River runs from right to left across the bottom of the image. The current eruption of Kliuchevskoi included massive ejections of gas, vapor and ash, which reached altitudes of 15,000 meters (50,000 feet). Melting snow mixed with volcanic ash triggered mud flows on the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMNH33E..06D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMNH33E..06D"><span>Microphysical Structures of Hurricane Irma Observed by Polarimetric <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Didlake, A. C.; Kumjian, M. R.</p> <p>2017-12-01</p> <p>This study examines dual-polarization <span class="hlt">radar</span> observations of Hurricane Irma as its center passed near the WSR-88D <span class="hlt">radar</span> in Puerto Rico, capturing needed microphysical information of a mature tropical cyclone. Twenty hours of observations continuously sampled the inner core precipitation features. These data were analyzed by annuli and azimuth, providing a bulk characterization of the primary eyewall, secondary eyewall, and rainbands as they varied around the storm. Polarimetric <span class="hlt">radar</span> variables displayed distinct signatures of convective and stratiform precipitation in the primary eyewall and rainbands that were organized in a manner consistent with the expected kinematic asymmetry of a storm in weak environmental wind shear but with moderate low-level storm-relative flow. In the front quadrants of the primary eyewall, <span class="hlt">vertical</span> profiles of differential reflectivity (ZDR) exhibit increasing values with decreasing height consistent with convective precipitation processes. In particular, the front-right quadrant exhibits a signature in reflectivity (ZH) and ZDR indicating larger, sparser drops, which is consistent with a stronger updraft present in this quadrant. In the rear quadrants, a sharply peaked ZDR maximum occurs within the melting layer, which is attributed of stratiform processes. In the rainbands, the convective to stratiform transition can be seen traveling from the front-right to the front-left quadrant. The front-right quadrant exhibits lower co-polar correlation coefficient (ρHV) values in the 3-8 km altitude layer, suggesting larger <span class="hlt">vertical</span> spreading of various hydrometeors that occurs in convective <span class="hlt">vertical</span> motions. The front-left quadrant exhibits larger ρHV values, suggesting less diversity of hydrometeor shapes, consistent with stratiform processes. The secondary eyewall did not exhibit a clear signature of processes preferred in a specific quadrant, and a temporal analysis of the secondary eyewall revealed a complex evolution of its structure</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29059967','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29059967"><span>Comparison of the impedance cardiogram with continuous wave <span class="hlt">radar</span> using body-contact antennas.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Buxi, Dilpreet; Dugar, Rahul; Redoute, Jean-Michel; Yuce, Mehmet Rasit</p> <p>2017-07-01</p> <p>This paper describes a continuous wave (CW) <span class="hlt">radar</span> system with body-contact antennas and basic signal processing. The goal is to assess the signals' reproducibility across different subjects as well as a respiration cycle. <span class="hlt">Radar</span> signals using body-contact antennas with a carrier frequency of 868 MHz are used to acquire the cardiac activity at the sternum. The <span class="hlt">radar</span> I and Q channel signals are combined to form their magnitude. Signals are collected from six healthy males during paced breathing conditions. The electrocardiogram (ECG) and impedance cardiogram (ICG) signals are acquired simultaneously as reference. The chosen feature in the <span class="hlt">radar</span> signal is the maximum of its second derivative, which is closest to the ICG B-<span class="hlt">point</span>. The median and mean absolute errors in pre-ejection period (PEP) in milliseconds between the ICG's B-<span class="hlt">point</span> and chosen feature in the <span class="hlt">radar</span> signal range from -6-119.7 ms and 7.8-62.3 ms for all subjects. The results indicate that a reproducible <span class="hlt">radar</span> signal is obtained from all six subjects. More work is needed on understanding the origin of the <span class="hlt">radar</span> signals using ultrasound as a comparison.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts068-s-055.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts068-s-055.html"><span>STS-68 <span class="hlt">radar</span> image: Glasgow, Missouri</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1994-10-07</p> <p>STS068-S-055 (7 October 1994) --- This is a false-color L-Band image of an area near Glasgow, Missouri, centered at about 39.2 degrees north latitude and 92.8 degrees west longitude. The image was acquired using the L-Band <span class="hlt">radar</span> channel (horizontally transmitted and received and horizontally transmitted and <span class="hlt">vertically</span> received) polarization's combined. The data were acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the Space Shuttle Endeavour on orbit 50 on October 3, 1994. The area shown is approximately 37 by 25 kilometers (23 by 16 miles). The <span class="hlt">radar</span> data, coupled with pre-flood aerial photography and satellite data and post-flood topographic and field data, are being used to evaluate changes associated with levee breaks in land forms, where deposits formed during the widespread flooding in 1993 along the Missouri and Mississippi Rivers. The distinct <span class="hlt">radar</span> scattering properties of farmland, sand fields and scoured areas will be used to inventory flood plains along the Missouri River and determine the processes by which these areas return to preflood conditions. The image shows one such levee break near Glasgow, Missouri. In the upper center of the <span class="hlt">radar</span> image, below the bend of the river, is a region covered by several meters of sand, shown as dark regions. West (left) of the dark areas, a gap in the levee tree canopy shows the area where the levee failed. <span class="hlt">Radar</span> data such as these can help scientists more accurately assess the potential for future flooding in this region and how that might impact surrounding communities. Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The <span class="hlt">radars</span> illuminate Earth with microwaves, allowing detailed observations at any time, regardless of weather or sunlight conditions. SIR-C/X-SAR uses the three microwave wavelengths: the L-Band (24 centimeters), C-Band (6 centimeters) and X-Band (3 centimeters). The multi</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01725.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01725.html"><span>Space <span class="hlt">Radar</span> Image of Niya Ruins, Taklamakan Desert</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This <span class="hlt">radar</span> image is of an area thought to contain the ruins of the ancient settlement of Niya. It is located in the southwestern corner of the Taklamakan Desert in China Sinjiang Province. This oasis was part of the famous Silk Road, an ancient trade route from one of China's earliest capitols, Xian, to the West. The image shows a white linear feature trending diagonally from the upper left to the lower right. Scientists believe this newly [sic] discovered feature is a man-made canal which presumably diverted river waters toward the settlement of Niya for irrigation purposes. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 106th orbit on April 16, 1994, and is centered at 37.78 degrees north latitude and 82.41 degrees east longitude. The false-color <span class="hlt">radar</span> image was created by displaying the C-band (horizontally transmitted and received) return in red, the L-band (horizontally transmitted and received) return in green, and the L-band (horizontally transmitted and <span class="hlt">vertically</span> received) return in blue. Areas in mottled white and purple are low-lying floodplains of the Niya River. Dark green and black areas between river courses are higher ridges or dunes confining the water flow. http://photojournal.jpl.nasa.gov/catalog/PIA01725</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870001022','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870001022"><span>Climatology of tropospheric <span class="hlt">vertical</span> velocity spectra</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ecklund, W. L.; Gage, K. S.; Balsley, B. B.; Carter, D. A.</p> <p>1986-01-01</p> <p><span class="hlt">Vertical</span> velocity power spectra obtained from Poker Flat, Alaska; Platteville, Colorado; Rhone Delta, France; and Ponape, East Caroline Islands using 50-MHz clear-air <span class="hlt">radars</span> with <span class="hlt">vertical</span> beams are given. The spectra were obtained by analyzing the quietest periods from the one-minute-resolution time series for each site. The lengths of available <span class="hlt">vertical</span> records ranged from as long as 6 months at Poker Flat to about 1 month at Platteville. The quiet-time <span class="hlt">vertical</span> velocity spectra are shown. Spectral period ranging from 2 minutes to 4 hours is shown on the abscissa and power spectral density is given on the ordinate. The Brunt-Vaisala (B-V) periods (determined from nearby sounding balloons) are indicated. All spectra (except the one from Platteville) exhibit a peak at periods slightly longer than the B-V period, are flat at longer periods, and fall rapidly at periods less than the B-V period. This behavior is expected for a spectrum of internal waves and is very similar to what is observed in the ocean (Eriksen, 1978). The spectral amplitudes vary by only a factor of 2 or 3 about the mean, and show that under quiet conditions <span class="hlt">vertical</span> velocity spectra from the troposphere are very similar at widely different locations.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ISPAr42W6..399W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ISPAr42W6..399W"><span><span class="hlt">Point</span> Cloud Analysis for Uav-Borne Laser Scanning with Horizontally and <span class="hlt">Vertically</span> Oriented Line Scanners - Concept and First Results</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weinmann, M.; Müller, M. S.; Hillemann, M.; Reydel, N.; Hinz, S.; Jutzi, B.</p> <p>2017-08-01</p> <p>In this paper, we focus on UAV-borne laser scanning with the objective of densely sampling object surfaces in the local surrounding of the UAV. In this regard, using a line scanner which scans along the <span class="hlt">vertical</span> direction and perpendicular to the flight direction results in a <span class="hlt">point</span> cloud with low <span class="hlt">point</span> density if the UAV moves fast. Using a line scanner which scans along the horizontal direction only delivers data corresponding to the altitude of the UAV and thus a low scene coverage. For these reasons, we present a concept and a system for UAV-borne laser scanning using multiple line scanners. Our system consists of a quadcopter equipped with horizontally and <span class="hlt">vertically</span> oriented line scanners. We demonstrate the capabilities of our system by presenting first results obtained for a flight within an outdoor scene. Thereby, we use a downsampling of the original <span class="hlt">point</span> cloud and different neighborhood types to extract fundamental geometric features which in turn can be used for scene interpretation with respect to linear, planar or volumetric structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003RaSc...38.8056B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003RaSc...38.8056B"><span>Quantitative estimation of Tropical Rainfall Mapping Mission precipitation <span class="hlt">radar</span> signals from ground-based polarimetric <span class="hlt">radar</span> observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bolen, Steven M.; Chandrasekar, V.</p> <p>2003-06-01</p> <p>The Tropical Rainfall Mapping Mission (TRMM) is the first mission dedicated to measuring rainfall from space using <span class="hlt">radar</span>. The precipitation <span class="hlt">radar</span> (PR) is one of several instruments aboard the TRMM satellite that is operating in a nearly circular orbit with nominal altitude of 350 km, inclination of 35°, and period of 91.5 min. The PR is a single-frequency Ku-band instrument that is designed to yield information about the <span class="hlt">vertical</span> storm structure so as to gain insight into the intensity and distribution of rainfall. Attenuation effects on PR measurements, however, can be significant and as high as 10-15 dB. This can seriously impair the accuracy of rain rate retrieval algorithms derived from PR signal returns. Quantitative estimation of PR attenuation is made along the PR beam via ground-based polarimetric observations to validate attenuation correction procedures used by the PR. The reflectivity (Zh) at horizontal polarization and specific differential phase (Kdp) are found along the beam from S-band ground <span class="hlt">radar</span> measurements, and theoretical modeling is used to determine the expected specific attenuation (k) along the space-Earth path at Ku-band frequency from these measurements. A theoretical k-Kdp relationship is determined for rain when Kdp ≥ 0.5°/km, and a power law relationship, k = a Zhb, is determined for light rain and other types of hydrometers encountered along the path. After alignment and resolution volume matching is made between ground and PR measurements, the two-way path-integrated attenuation (PIA) is calculated along the PR propagation path by integrating the specific attenuation along the path. The PR reflectivity derived after removing the PIA is also compared against ground <span class="hlt">radar</span> observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10189E..07D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10189E..07D"><span>Improved characterization of scenes with a combination of MMW <span class="hlt">radar</span> and radiometer information</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dill, Stephan; Peichl, Markus; Schreiber, Eric; Anglberger, Harald</p> <p>2017-05-01</p> <p>For security related applications MMW <span class="hlt">radar</span> and radiometer systems in remote sensing or stand-off configurations are well established techniques. The range of development stages extends from experimental to commercial systems on the civil and military market. Typical examples are systems for personnel screening at airports for concealed object detection under clothing, enhanced vision or landing aid for helicopter and vehicle based systems for suspicious object or IED detection along roads. Due to the physical principle of active (<span class="hlt">radar</span>) and passive (radiometer) MMW measurement techniques the appearance of single objects and thus the complete scenario is rather different for <span class="hlt">radar</span> and radiometer images. A reasonable combination of both measurement techniques could lead to enhanced object information. However, some technical requirements should be taken into account. The imaging geometry for both sensors should be nearly identical, the geometrical resolution and the wavelength should be similar and at best the imaging process should be carried out simultaneously. Therefore theoretical and experimental investigations on a suitable combination of MMW <span class="hlt">radar</span> and radiometer information have been conducted. First experiments in 2016 have been done with an imaging linescanner based on a cylindrical imaging geometry [1]. It combines a horizontal line scan in azimuth with a linear motion in <span class="hlt">vertical</span> direction for the second image dimension. The main drawback of the system is the limited number of pixel in <span class="hlt">vertical</span> dimension at a certain distance. Nevertheless the near range imaging results where promising. Therefore the combination of <span class="hlt">radar</span> and radiometer sensor was assembled on the DLR wide-field-of-view linescanner ABOSCA which is based on a spherical imaging geometry [2]. A comparison of both imaging systems is discussed. The investigations concentrate on rather basic scenarios with canonical targets like flat plates, spheres, corner reflectors and cylinders. First</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JGRD..11810056M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JGRD..11810056M"><span>A study on the use of <span class="hlt">radar</span> and lidar for characterizing ultragiant aerosol</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Madonna, F.; Amodeo, A.; D'Amico, G.; Pappalardo, G.</p> <p>2013-09-01</p> <p>19 April to 19 May 2010, volcanic aerosol layers originating from the Eyjafjallajökull volcano were observed at the Institute of Methodologies for Environmental Analysis of the National Research Council of Italy Atmospheric Observatory, named CIAO (40.60°N, 15.72°E, 760 m above sea level), in Southern Italy with a multiwavelength Raman lidar. During this period, ultragiant aerosols were also observed at CIAO using a colocated 8.45 mm wavelength Doppler <span class="hlt">radar</span>. The Ka-band <span class="hlt">radar</span> signatures observed in four separate days (19 April and 7, 10, and 13 May) are consistent with the observation of nonspherical ultragiant aerosols characterized by values of linear depolarization ratio (LDR) higher than -4 dB. Air mass back trajectory analysis suggests a volcanic origin of the ultragiant aerosols observed by the <span class="hlt">radar</span>. The observed values of the <span class="hlt">radar</span> reflectivity (Ze) are consistent with a particle effective radius (r) larger than 50-75 µm. Scattering simulations based on the T-matrix approach show that the high LDR values can be explained if the observed particles have an absolute aspect ratio larger than 3.0 and consist of an internal aerosol core and external ice shell, with a variable radius ratio ranging between 0.2 and 0.7 depending on the shape and aspect ratio. Comparisons between daytime <span class="hlt">vertical</span> profiles of aerosol backscatter coefficient (β) as measured by lidar and <span class="hlt">radar</span> LDR reveal a decrease of β where ultragiant particles are observed. Scattering simulations based on Mie theory show how the lidar capability in typing ultragiant aerosols could be limited by low number concentrations or by the presence of an external ice shell covering the aerosol particles. Preferential <span class="hlt">vertical</span> alignment of the particles is discussed as another possible reason for the decrease of β.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225465p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225465p/"><span>Topography adjacent to Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Topography adjacent to Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5, showing conditions before construction, May 28, 1943, this drawing shows the Bonita Ridge access road retaining wall and general conditions at Bonita Ridge before the construction of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100002840','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100002840"><span><span class="hlt">Radar</span> Interferometer for Topographic Mapping of Glaciers and Ice Sheets</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Moller, Delwyn K.; Sadowy, Gregory A.; Rignot, Eric J.; Madsen, Soren N.</p> <p>2007-01-01</p> <p>A report discusses Ka-band (35-GHz) <span class="hlt">radar</span> for mapping the surface topography of glaciers and ice sheets at high spatial resolution and high <span class="hlt">vertical</span> accuracy, independent of cloud cover, with a swath-width of 70 km. The system is a single- pass, single-platform interferometric synthetic aperture <span class="hlt">radar</span> (InSAR) with an 8-mm wavelength, which minimizes snow penetration while remaining relatively impervious to atmospheric attenuation. As exhibited by the lower frequency SRTM (Shuttle <span class="hlt">Radar</span> Topography Mission) AirSAR and GeoSAR systems, an InSAR measures topography using two antennas separated by a baseline in the cross-track direction, to view the same region on the ground. The interferometric combination of data received allows the system to resolve the pathlength difference from the illuminated area to the antennas to a fraction of a wavelength. From the interferometric phase, the height of the target area can be estimated. This means an InSAR system is capable of providing not only the position of each image <span class="hlt">point</span> in along-track and slant range as with a traditional SAR but also the height of that <span class="hlt">point</span> through interferometry. Although the evolution of InSAR to a millimeter-wave center frequency maximizes the interferometric accuracy from a given baseline length, the high frequency also creates a fundamental problem of swath coverage versus signal-to-noise ratio. While the length of SAR antennas is typically fixed by mass and stowage or deployment constraints, the width is constrained by the desired illuminated swath width. As the across-track beam width which sets the swath size is proportional to the wavelength, a fixed swath size equates to a smaller antenna as the frequency is increased. This loss of antenna size reduces the two-way antenna gain to the second power, drastically reducing the signal-to-noise ratio of the SAR system. This fundamental constraint of high-frequency SAR systems is addressed by applying digital beam-forming (DBF) techniques to</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.A11H0116P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.A11H0116P"><span>Assessment of C-band Polarimetric <span class="hlt">Radar</span> Rainfall Measurements During Strong Attenuation.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Paredes-Victoria, P. N.; Rico-Ramirez, M. A.; Pedrozo-Acuña, A.</p> <p>2016-12-01</p> <p>In the modern hydrological modelling and their applications on flood forecasting systems and climate modelling, reliable spatiotemporal rainfall measurements are the keystone. Raingauges are the foundation in hydrology to collect rainfall data, however they are prone to errors (e.g. systematic, malfunctioning, and instrumental errors). Moreover rainfall data from gauges is often used to calibrate and validate weather <span class="hlt">radar</span> rainfall, which is distributed in space. Therefore, it is important to apply techniques to control the quality of the raingauge data in order to guarantee a high level of confidence in rainfall measurements for <span class="hlt">radar</span> calibration and numerical weather modelling. Also, the reliability of <span class="hlt">radar</span> data is often limited because of the errors in the <span class="hlt">radar</span> signal (e.g. clutter, variation of the <span class="hlt">vertical</span> reflectivity profile, beam blockage, attenuation, etc) which need to be corrected in order to increase the accuracy of the <span class="hlt">radar</span> rainfall estimation. This paper presents a method for raingauge-measurement quality-control correction based on the inverse distance weighted as a function of correlated climatology (i.e. performed by using the reflectivity from weather <span class="hlt">radar</span>). Also a Clutter Mitigation Decision (CMD) algorithm is applied for clutter filtering process, finally three algorithms based on differential phase measurements are applied for <span class="hlt">radar</span> signal attenuation correction. The quality-control method proves that correlated climatology is very sensitive in the first 100 kilometres for this area. The results also showed that ground clutter affects slightly the <span class="hlt">radar</span> measurements due to the low gradient of the terrain in the area. However, strong <span class="hlt">radar</span> signal attenuation is often found in this data set due to the heavy storms that take place in this region and the differential phase measurements are crucial to correct for attenuation at C-band frequencies. The study area is located in Sabancuy-Campeche, Mexico (Latitude 18.97 N, Longitude 91.17º W) and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70047050','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70047050"><span>Phase and amplitude inversion of crosswell <span class="hlt">radar</span> data</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ellefsen, Karl J.; Mazzella, Aldo T.; Horton, Robert J.; McKenna, Jason R.</p> <p>2011-01-01</p> <p>Phase and amplitude inversion of crosswell <span class="hlt">radar</span> data estimates the logarithm of complex slowness for a 2.5D heterogeneous model. The inversion is formulated in the frequency domain using the vector Helmholtz equation. The objective function is minimized using a back-propagation method that is suitable for a 2.5D model and that accounts for the near-, intermediate-, and far-field regions of the antennas. The inversion is tested with crosswell <span class="hlt">radar</span> data collected in a laboratory tank. The model anomalies are consistent with the known heterogeneity in the tank; the model’s relative dielectric permittivity, which is calculated from the real part of the estimated complex slowness, is consistent with independent laboratory measurements. The methodologies developed for this inversion can be adapted readily to inversions of seismic data (e.g., crosswell seismic and <span class="hlt">vertical</span> seismic profiling data).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19730029740&hterms=DRONE+QUADRONE&qs=Ntx%3Dmode%2Bmatchany%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DDRONE%2BQUADRONE','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19730029740&hterms=DRONE+QUADRONE&qs=Ntx%3Dmode%2Bmatchany%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DDRONE%2BQUADRONE"><span>Intensive probing of clear air convective fields by <span class="hlt">radar</span> and instrumented drone aircraft.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rowland, J. R.</p> <p>1972-01-01</p> <p>Clear air convective fields were probed in three summer experiments (1969, 1970, and 1971) on an S-band monopulse tracking <span class="hlt">radar</span> at Wallops Island, Virginia, and a drone aircraft with a takeoff weight of 5.2 kg, wingspan of 2.5 m, and cruising glide speed of 10.3 m/sec. The drone was flown 23.2 km north of the <span class="hlt">radar</span> and carried temperature, pressure/altitude, humidity, and <span class="hlt">vertical</span> and airspeed velocity sensors. Extensive time-space convective field data were obtained by taking a large number of RHI and PPI pictures at short intervals of time. The rapidly changing overall convective field data obtained from the <span class="hlt">radar</span> could be related to the meteorological information telemetered from the drone at a reasonably low cost by this combined technique.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01747.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01747.html"><span>Space <span class="hlt">Radar</span> Image of Kennedy Space Center, Florida</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-06-25</p> <p>This image was produced during <span class="hlt">radar</span> observations taken by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> as it flew over the Gulf Stream, Florida, and past the Atlantic Ocean on October 7, 1994. The data were produced using the X-band <span class="hlt">radar</span> frequency. Knowing ahead of time that this region would be included in a regularly scheduled <span class="hlt">radar</span> pass, the Kennedy Space Center team, who assembled and integrated the SIR-C/X-SAR equipment with the Spacelab pallet system, designed a set of <span class="hlt">radar</span> reflectors from common construction materials and formed the letters "KSC" on the ground adjacent to the main headquarters building at the entrance to the Cape Canaveral launch facility. The <span class="hlt">point</span> of light formed by the bright return from these reflectors are visible in the image. Other more diffuse bright spots are reflections from building faces, roofs and other large structures at the Kennedy Space Center complex. This frame covers an area of approximately 6 kilometers by 8 kilometers (4 miles by 5 miles), which was just a small portion of the data taken on this particular pass. http://photojournal.jpl.nasa.gov/catalog/PIA01747</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340240p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340240p/"><span>2. VIEW SOUTHWEST, prime search <span class="hlt">radar</span> tower, height finder <span class="hlt">radar</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>2. VIEW SOUTHWEST, prime search <span class="hlt">radar</span> tower, height finder <span class="hlt">radar</span> towards, height finder <span class="hlt">radar</span> towers, and <span class="hlt">radar</span> tower (unknown function) - Fort Custer Military Reservation, P-67 <span class="hlt">Radar</span> Station, .25 mile north of Dickman Road, east of Clark Road, Battle Creek, Calhoun County, MI</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JHyd..357....1F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JHyd..357....1F"><span>Measuring real-time streamflow using emerging technologies: <span class="hlt">Radar</span>, hydroacoustics, and the probability concept</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fulton, John; Ostrowski, Joseph</p> <p>2008-07-01</p> <p>SummaryForecasting streamflow during extreme hydrologic events such as floods can be problematic. This is particularly true when flow is unsteady, and river forecasts rely on models that require uniform-flow rating curves to route water from one forecast <span class="hlt">point</span> to another. As a result, alternative methods for measuring streamflow are needed to properly route flood waves and account for inertial and pressure forces in natural channels dominated by nonuniform-flow conditions such as mild water surface slopes, backwater, tributary inflows, and reservoir operations. The objective of the demonstration was to use emerging technologies to measure instantaneous streamflow in open channels at two existing US Geological Survey streamflow-gaging stations in Pennsylvania. Surface-water and instream-<span class="hlt">point</span> velocities were measured using hand-held <span class="hlt">radar</span> and hydroacoustics. Streamflow was computed using the probability concept, which requires velocity data from a single <span class="hlt">vertical</span> containing the maximum instream velocity. The percent difference in streamflow at the Susquehanna River at Bloomsburg, PA ranged from 0% to 8% with an average difference of 4% and standard deviation of 8.81 m 3/s. The percent difference in streamflow at Chartiers Creek at Carnegie, PA ranged from 0% to 11% with an average difference of 5% and standard deviation of 0.28 m 3/s. New generation equipment is being tested and developed to advance the use of <span class="hlt">radar</span>-derived surface-water velocity and instantaneous streamflow to facilitate the collection and transmission of real-time streamflow that can be used to parameterize hydraulic routing models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70033186','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70033186"><span>Measuring real-time streamflow using emerging technologies: <span class="hlt">Radar</span>, hydroacoustics, and the probability concept</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fulton, J.; Ostrowski, J.</p> <p>2008-01-01</p> <p>Forecasting streamflow during extreme hydrologic events such as floods can be problematic. This is particularly true when flow is unsteady, and river forecasts rely on models that require uniform-flow rating curves to route water from one forecast <span class="hlt">point</span> to another. As a result, alternative methods for measuring streamflow are needed to properly route flood waves and account for inertial and pressure forces in natural channels dominated by nonuniform-flow conditions such as mild water surface slopes, backwater, tributary inflows, and reservoir operations. The objective of the demonstration was to use emerging technologies to measure instantaneous streamflow in open channels at two existing US Geological Survey streamflow-gaging stations in Pennsylvania. Surface-water and instream-<span class="hlt">point</span> velocities were measured using hand-held <span class="hlt">radar</span> and hydroacoustics. Streamflow was computed using the probability concept, which requires velocity data from a single <span class="hlt">vertical</span> containing the maximum instream velocity. The percent difference in streamflow at the Susquehanna River at Bloomsburg, PA ranged from 0% to 8% with an average difference of 4% and standard deviation of 8.81 m3/s. The percent difference in streamflow at Chartiers Creek at Carnegie, PA ranged from 0% to 11% with an average difference of 5% and standard deviation of 0.28 m3/s. New generation equipment is being tested and developed to advance the use of <span class="hlt">radar</span>-derived surface-water velocity and instantaneous streamflow to facilitate the collection and transmission of real-time streamflow that can be used to parameterize hydraulic routing models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860060387&hterms=speckle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dspeckle','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860060387&hterms=speckle&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dspeckle"><span>Use of speckle for determining the response characteristics of Doppler imaging <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tilley, D. G.</p> <p>1986-01-01</p> <p>An optical model is developed for imaging optical <span class="hlt">radars</span> such as the SAR on Seasat and the Shuttle Imaging <span class="hlt">Radar</span> (SIR-B) by analyzing the Doppler shift of individual speckles in the image. The signal received at the spacecraft is treated in terms of a Fresnel-Kirchhoff integration over all backscattered radiation within a Huygen aperture at the earth. Account is taken of the movement of the spacecraft along the orbital path between emission and reception. The individual <span class="hlt">points</span> are described by integration of the <span class="hlt">point</span> source amplitude with a Green's function scattering kernel. Doppler data at each <span class="hlt">point</span> furnishes the coordinates for visual representations. A Rayleigh-Poisson model of the surface scattering characteristics is used with Monte Carlo methods to generate simulations of Doppler <span class="hlt">radar</span> speckle that compare well with Seasat SAR data SIR-B data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.1240S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.1240S"><span>Multifractal analysis of different hydrological products of X-band <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Skouri-Plakali, Ilektra; Da Silva Rocha Paz, Igor; Ichiba, Abdellah; Gires, Auguste; Tchiguirinskaia, Ioulia; Schertzer, Daniel</p> <p>2017-04-01</p> <p>Rainfall is widely considered as the hydrological process that triggers all the others. Its accurate measurements are crucial especially when they are used afterwards for the hydrological modeling of urban and peri-urban catchments for decision-making. Rainfall is a complex process and is scale dependent in space and time. Hence a high spatial and temporal resolution of the data is more appropriate for urban modeling. Therefore, a great interest of high-resolution measurements of precipitation in space and time is manifested. <span class="hlt">Radar</span> technologies have not stopped evolving since their first appearance about the mid-twentieth. Indeed, the turning <span class="hlt">point</span> work by Marshall-Palmer (1948) has established the Z - R power-law relation that has been widely used, with major scientific efforts being devoted to find "the best choice" of the two associated parameters. Nowadays X-band <span class="hlt">radars</span>, being provided with dual-polarization and Doppler means, offer more accurate data of higher resolution. The fact that drops are oblate induces a differential phase shift between the two polarizations. The quantity most commonly used for the rainfall rate computation is actually the specific differential phase shift, which is the gradient of the differential phase shift along the radial beam direction. It is even stronger correlated to the rain rate R than reflectivity Z. Hence the rain rate can be computed with a different power-law relation, which again depends on only two parameters. Furthermore, an attenuation correction is needed to adjust the loss of <span class="hlt">radar</span> energy due to the absorption and scattering as it passes through the atmosphere. Due to natural variations of reflectivity with altitude, <span class="hlt">vertical</span> profile of reflectivity should be corrected as well. There are some other typical <span class="hlt">radar</span> data filtering procedures, all resulting in various hydrological products. In this work, we use the Universal Multifractal framework to analyze and to inter-compare different products of X-band <span class="hlt">radar</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1424987-toward-exploring-synergy-between-cloud-radar-polarimetry-doppler-spectral-analysis-deep-cold-precipitating-systems-arctic','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1424987-toward-exploring-synergy-between-cloud-radar-polarimetry-doppler-spectral-analysis-deep-cold-precipitating-systems-arctic"><span>Toward Exploring the Synergy Between Cloud <span class="hlt">Radar</span> Polarimetry and Doppler Spectral Analysis in Deep Cold Precipitating Systems in the Arctic</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>Oue, Mariko; Kollias, Pavlos; Ryzhkov, Alexander</p> <p></p> <p>The study of Arctic ice and mixed-phase clouds, which are characterized by a variety of ice particle types in the same cloudy volume, is challenging research. This study illustrates a new approach to qualitative and quantitative analysis of the complexity of ice and mixed-phase microphysical processes in Arctic deep precipitating systems using the combination of Ka-band zenith-<span class="hlt">pointing</span> <span class="hlt">radar</span> Doppler spectra and quasi-<span class="hlt">vertical</span> profiles of polarimetric <span class="hlt">radar</span> variables measured by a Ka/W-band scanning <span class="hlt">radar</span>. The results illustrate the frequent occurrence of multimodal Doppler spectra in the dendritic/planar growth layer, where locally generated, slower-falling particle populations are well separated from faster-falling populationsmore » in terms of Doppler velocity. The slower-falling particle populations contribute to an increase of differential reflectivity (Z DR), while an enhanced specific differential phase (K DP) in this dendritic growth temperature range is caused by both the slower and faster-falling particle populations. Another area with frequent occurrence of multimodal Doppler spectra is in mixed-phase layers, where both populations produce Z DR and K DP values close to 0, suggesting the occurrence of a riming process. A Joint analysis of the Doppler spectra and the polarimetric <span class="hlt">radar</span> variables provides important insight into the microphysics of snow formation and allows the separation of the contributions of ice of different habits to the values of reflectivity and Z DR.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1424987-toward-exploring-synergy-between-cloud-radar-polarimetry-doppler-spectral-analysis-deep-cold-precipitating-systems-arctic','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1424987-toward-exploring-synergy-between-cloud-radar-polarimetry-doppler-spectral-analysis-deep-cold-precipitating-systems-arctic"><span>Toward Exploring the Synergy Between Cloud <span class="hlt">Radar</span> Polarimetry and Doppler Spectral Analysis in Deep Cold Precipitating Systems in the Arctic</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Oue, Mariko; Kollias, Pavlos; Ryzhkov, Alexander; ...</p> <p>2018-03-16</p> <p>The study of Arctic ice and mixed-phase clouds, which are characterized by a variety of ice particle types in the same cloudy volume, is challenging research. This study illustrates a new approach to qualitative and quantitative analysis of the complexity of ice and mixed-phase microphysical processes in Arctic deep precipitating systems using the combination of Ka-band zenith-<span class="hlt">pointing</span> <span class="hlt">radar</span> Doppler spectra and quasi-<span class="hlt">vertical</span> profiles of polarimetric <span class="hlt">radar</span> variables measured by a Ka/W-band scanning <span class="hlt">radar</span>. The results illustrate the frequent occurrence of multimodal Doppler spectra in the dendritic/planar growth layer, where locally generated, slower-falling particle populations are well separated from faster-falling populationsmore » in terms of Doppler velocity. The slower-falling particle populations contribute to an increase of differential reflectivity (Z DR), while an enhanced specific differential phase (K DP) in this dendritic growth temperature range is caused by both the slower and faster-falling particle populations. Another area with frequent occurrence of multimodal Doppler spectra is in mixed-phase layers, where both populations produce Z DR and K DP values close to 0, suggesting the occurrence of a riming process. A Joint analysis of the Doppler spectra and the polarimetric <span class="hlt">radar</span> variables provides important insight into the microphysics of snow formation and allows the separation of the contributions of ice of different habits to the values of reflectivity and Z DR.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.3815P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.3815P"><span>Characteristics of Moderately Deep Tropical Convection Observed by Dual-Polarimetric <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Powell, Scott</p> <p>2017-04-01</p> <p>Moderately deep cumulonimbus clouds (often erroneously called congestus) over the tropical warm pool play an important role in large-scale dynamics by moistening the free troposphere, thus allowing for the upscale growth of convection into mesoscale convective systems. Direct observational analysis of such convection has been limited despite a wealth of <span class="hlt">radar</span> data collected during several field experiments in the tropics. In this study, the structure of isolated cumulonimbus clouds, particularly those in the moderately deep mode with heights of up to 8 km, as observed by RHI scans obtained with the S-PolKa <span class="hlt">radar</span> during DYNAMO is explored. Such elements are first identified following the algorithm of Powell et al (2016); small contiguous regions of echo are considered isolated convection. Within isolated echo objects, echoes are further subdivided into core echoes, which feature <span class="hlt">vertical</span> profiles reflectivity and differential reflectivity that is similar to convection embedded in larger cloud complexes, and fringe echoes, which contain <span class="hlt">vertical</span> profiles of differential reflectivity that are more similar to stratiform regions. Between the surface and 4 km, reflectivities of 30-40 (10-20) dBZ are most commonly observed in isolated convective core (fringe) echoes. Convective cores in echo objects too wide to be considered isolated have a ZDR profile that peaks near the surface (with values of 0.5-1 dB common), and decays linearly to about 0.3 dB at and above an altitude of 6 km. Stratiform echoes have a minimum ZDR below of 0-0.5 dB below the bright band and a constant distribution centered on 0.5 dB above the bright band. The isolated convective core and fringe respectively possess composite <span class="hlt">vertical</span> profiles of ZDR that resemble convective and stratiform echoes. The mode of the distribution of aspect ratios of isolated convection is approximately 2.3, but the long axis of isolated echo objects demonstrates no preferred orientation. An early attempt at illustrating</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01788&hterms=shrubs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dshrubs','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01788&hterms=shrubs&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dshrubs"><span>Space <span class="hlt">Radar</span> Image of Harvard Forest</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1999-01-01</p> <p>This is a <span class="hlt">radar</span> image of the area surrounding the Harvard Forest in north-central Massachusetts that has been operated as a ecological research facility by Harvard University since 1907. At the center of the image is the Quabbin Reservoir, and the Connecticut River is at the lower left of the image. The Harvard Forest itself is just above the reservoir. Researchers are comparing the naturally occurring physical disturbances in the forest and the recent and projected chemical disturbances and their effects on the forest ecosystem. Agricultural land appears dark blue/purple, along with low shrub vegetation and some wetlands. Urban development is bright pink; the yellow to green tints are conifer-dominated vegetation with the pitch pine sand plain at the middle left edge of the image appearing very distinctive. The green tint may indicate pure pine plantation stands, and deciduous broadleaf trees appear gray/pink with perhaps wetter sites being pinker. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered at 42.50 degrees North latitude and 72.33 degrees West longitude and covers an area of 53 kilometers 63 by kilometers (33 miles by 39 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted and horizontally received; green is L-band horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band horizontally transmitted and horizontally received.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01846&hterms=livestock&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dlivestock','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01846&hterms=livestock&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dlivestock"><span>Space <span class="hlt">Radar</span> Image of Tuva, Central Asia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This spaceborne <span class="hlt">radar</span> image shows part of the remote central Asian region of Tuva, an autonomous republic of the Russian Federation. Tuva is a mostly mountainous region that lies between western Mongolia and southern Siberia. This image shows the area just south of the republic's capital of Kyzyl. Most of the red, pink and blue areas in the image are agricultural fields of a large collective farming complex that was developed during the era of the Soviet Union. Traditional agricultural activity in the region, still active in remote areas, revolves around practices of nomadic livestock herding. White areas on the image are north-facing hillsides, which develop denser forests than south-facing slopes. The river in the upper right is one of the two major branches of the Yenesey River. Tuva has received some notoriety in recent years due to the intense interest of the celebrated Caltech physicist Dr. Richard Feynman, chronicled in the book 'Tuva or Bust' by Ralph Leighton. The image was acquired by Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band SyntheticAperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour onOctober 1, 1994. The image is 56 kilometers by 74 kilometers (35 miles by 46 miles) and is centered at 51.5 degrees north latitude, 95.1 degrees east longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> fequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band, horizontally transmitted andreceived; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and verticallyreceived. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to PlanetEarth program.</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('https://ntrs.nasa.gov/search.jsp?R=PIA01793&hterms=Ancient+people&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DAncient%2Bpeople','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01793&hterms=Ancient+people&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DAncient%2Bpeople"><span>Space <span class="hlt">Radar</span> Image of Giza Egypt - with enlargement</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image shows the area west of the Nile River near Cairo, Egypt. The Nile River is the dark band along the right side of the image and it flows approximately due North from the bottom to the right. The boundary between dense urbanization and the desert can be clearly seen between the bright and dark areas in the center of the image. This boundary represents the approximate extent of yearly Nile flooding which played an important part in determining where people lived in ancient Egypt. This land usage pattern persists to this day. The pyramids at Giza appear as three bright triangles aligned with the image top just at the boundary of the urbanized area. They are also shown enlarged in the inset box in the top left of the image. The Great Pyramid of Khufu (Cheops in Greek) is the northern most of the three Giza pyramids. The side-looking <span class="hlt">radar</span> illuminates the scene from the top, the two sides of the pyramids facing the <span class="hlt">radar</span> reflect most of the energy back to the antenna and appear <span class="hlt">radar</span> bright; the two sides away from the <span class="hlt">radar</span> reflect less energy back and appear dark Two additional pyramids can be seen left of center in the lower portion of the image. The modern development in the desert on the left side of the image is the Sixth of October City, an area of factories and residences started by Anwar Sadat to relieve urban crowding. The image was taken on April 19, 1994 by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered on latitude 29.72 degrees North latitude and 30.83 degrees East longitude. The area shown is approximately 20 kilometers by 30 kilometers. The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is C</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01793.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01793.html"><span>Space <span class="hlt">Radar</span> Image of Giza Egypt - with Enlargement</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This <span class="hlt">radar</span> image shows the area west of the Nile River near Cairo, Egypt. The Nile River is the dark band along the right side of the image and it flows approximately due North from the bottom to the right. The boundary between dense urbanization and the desert can be clearly seen between the bright and dark areas in the center of the image. This boundary represents the approximate extent of yearly Nile flooding which played an important part in determining where people lived in ancient Egypt. This land usage pattern persists to this day. The pyramids at Giza appear as three bright triangles aligned with the image top just at the boundary of the urbanized area. They are also shown enlarged in the inset box in the top left of the image. The Great Pyramid of Khufu (Cheops in Greek) is the northern most of the three Giza pyramids. The side-looking <span class="hlt">radar</span> illuminates the scene from the top, the two sides of the pyramids facing the <span class="hlt">radar</span> reflect most of the energy back to the antenna and appear <span class="hlt">radar</span> bright; the two sides away from the <span class="hlt">radar</span> reflect less energy back and appear dark Two additional pyramids can be seen left of center in the lower portion of the image. The modern development in the desert on the left side of the image is the Sixth of October City, an area of factories and residences started by Anwar Sadat to relieve urban crowding. The image was taken on April 19, 1994 by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered on latitude 29.72 degrees North latitude and 30.83 degrees East longitude. The area shown is approximately 20 kilometers by 30 kilometers. The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted, horizontally received; green is C</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19810021786','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19810021786"><span><span class="hlt">Radar</span> altimeter waveform modeled parameter recovery. [SEASAT-1 data</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>Satellite-borne <span class="hlt">radar</span> altimeters include waveform sampling gates providing <span class="hlt">point</span> samples of the transmitted <span class="hlt">radar</span> pulse after its scattering from the ocean's surface. Averages of the waveform sampler data can be fitted by varying parameters in a model mean return waveform. The theoretical waveform model used is described as well as a general iterative nonlinear least squares procedures used to obtain estimates of parameters characterizing the modeled waveform for SEASAT-1 data. The six waveform parameters recovered by the fitting procedure are: (1) amplitude; (2) time origin, or track <span class="hlt">point</span>; (3) ocean surface rms roughness; (4) noise baseline; (5) ocean surface skewness; and (6) altitude or off-nadir angle. Additional practical processing considerations are addressed and FORTRAN source listing for subroutines used in the waveform fitting are included. While the description is for the Seasat-1 altimeter waveform data analysis, the work can easily be generalized and extended to other <span class="hlt">radar</span> altimeter systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014SPIE.8941E..1EB','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014SPIE.8941E..1EB"><span>THz impulse <span class="hlt">radar</span> for biomedical sensing: nonlinear system behavior</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brown, E. R.; Sung, Shijun; Grundfest, W. S.; Taylor, Z. D.</p> <p>2014-03-01</p> <p>The THz impulse <span class="hlt">radar</span> is an "RF-inspired" sensor system that has performed remarkably well since its initial development nearly six years ago. It was developed for ex vivo skin-burn imaging, and has since shown great promise in the sensitive detection of hydration levels in soft tissues of several types, such as in vivo corneal and burn samples. An intriguing aspect of the impulse <span class="hlt">radar</span> is its hybrid architecture which combines the high-peak-power of photoconductive switches with the high-responsivity and -bandwidth (RF and video) of Schottky-diode rectifiers. The result is a very sensitive sensor system in which the post-detection signal-to-noise ratio depends super-linearly on average signal power up to a <span class="hlt">point</span> where the diode is "turned on" in the forward direction, and then behaves quasi-linearly beyond that <span class="hlt">point</span>. This paper reports the first nonlinear systems analysis done on the impulse <span class="hlt">radar</span> using MATLAB.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=138891&Lab=NRMRL&keyword=climatology&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=138891&Lab=NRMRL&keyword=climatology&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>SUB-PIXEL RAINFALL VARIABILITY AND THE IMPLICATIONS FOR UNCERTAINTIES IN <span class="hlt">RADAR</span> RAINFALL ESTIMATES</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p><span class="hlt">Radar</span> estimates of rainfall are subject to significant measurement uncertainty. Typically, uncertainties are measured by the discrepancies between real rainfall estimates based on <span class="hlt">radar</span> reflectivity and <span class="hlt">point</span> rainfall records of rain gauges. This study investigates how the disc...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010122735&hterms=death&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddeath','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010122735&hterms=death&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddeath"><span>Imaging <span class="hlt">Radar</span> in the Mojave Desert-Death Valley Region</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Farr, Tom G.</p> <p>2001-01-01</p> <p>The Mojave Desert-Death Valley region has had a long history as a test bed for remote sensing techniques. Along with visible-near infrared and thermal IR sensors, imaging <span class="hlt">radars</span> have flown and orbited over the area since the 1970's, yielding new insights into the geologic applications of these technologies. More recently, <span class="hlt">radar</span> interferometry has been used to derive digital topographic maps of the area, supplementing the USGS 7.5' digital quadrangles currently available for nearly the entire area. As for their shorter-wavelength brethren, imaging <span class="hlt">radars</span> were tested early in their civilian history in the Mojave Desert-Death Valley region because it contains a variety of surface types in a small area without the confounding effects of vegetation. The earliest imaging <span class="hlt">radars</span> to be flown over the region included military tests of short-wavelength (3 cm) X-band sensors. Later, the Jet Propulsion Laboratory began its development of imaging <span class="hlt">radars</span> with an airborne sensor, followed by the Seasat orbital <span class="hlt">radar</span> in 1978. These systems were L-band (25 cm). Following Seasat, JPL embarked upon a series of Space Shuttle Imaging <span class="hlt">Radars</span>: SIRA (1981), SIR-B (1984), and SIR-C (1994). The most recent in the series was the most capable <span class="hlt">radar</span> sensor flown in space and acquired large numbers of data swaths in a variety of test areas around the world. The Mojave Desert-Death Valley region was one of those test areas, and was covered very well with 3 wavelengths, multiple polarizations, and at multiple angles. At the same time, the JPL aircraft <span class="hlt">radar</span> program continued improving and collecting data over the Mojave Desert Death Valley region. Now called AIRSAR, the system includes 3 bands (P-band, 67 cm; L-band, 25 cm; C-band, 5 cm). Each band can collect all possible polarizations in a mode called polarimetry. In addition, AIRSAR can be operated in the TOPSAR mode wherein 2 antennas collect data interferometrically, yielding a digital elevation model (DEM). Both L-band and C-band can be</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmRe.203..216G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmRe.203..216G"><span>Pseudo-<span class="hlt">radar</span> algorithms with two extremely wet months of disdrometer data in the Paris area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gires, A.; Tchiguirinskaia, I.; Schertzer, D.</p> <p>2018-05-01</p> <p>Disdrometer data collected during the two extremely wet months of May and June 2016 at the Ecole des Ponts ParisTech are used to get insights on <span class="hlt">radar</span> algorithms. The rain rate and pseudo-<span class="hlt">radar</span> quantities (horizontal and <span class="hlt">vertical</span> reflectivity, specific differential phase shift) are all estimated over several durations with the help of drop size distributions (DSD) collected at 30 s time steps. The pseudo-<span class="hlt">radar</span> quantities are defined with simplifying hypotheses, in particular on the DSD homogeneity. First it appears that the parameters of the standard <span class="hlt">radar</span> relations Zh - R, R - Kdp and R - Zh - Zdr for these pseudo-<span class="hlt">radar</span> quantities exhibit strong variability between events and even within an event. Second an innovative methodology that relies on checking the ability of a given algorithm to reproduce the good scale invariant multifractal behaviour (on scales 30 s - few h) observed on rainfall time series is implemented. In this framework, the classical hybrid model (Zh - R for low rain rates and R - Kdp for great ones) performs best, as well as the local estimates of the <span class="hlt">radar</span> relations' parameters. However, we emphasise that due to the hypotheses on which they rely these observations cannot be straightforwardly extended to real <span class="hlt">radar</span> quantities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008amos.confE..55B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008amos.confE..55B"><span>The Precision Expandable <span class="hlt">Radar</span> Calibration Sphere (PERCS) With Applications for Laser Imaging and Ranging</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bernhardt, P.; Nicholas, A.; Thomas, L.; Davis, M.; Hoberman, C.; Davis, M.</p> <p></p> <p>The Naval Research Laboratory will provide an orbiting calibration sphere to be used with ground-based laser imaging telescopes and HF radio systems. The Precision Expandable <span class="hlt">Radar</span> Calibration Sphere (PERCS) is a practical, reliable, high-performance HF calibration sphere and laser imaging target to orbit at about 600 km altitude. The sphere will be made of a spherical wire frame with aspect independent <span class="hlt">radar</span> cross section in the 3 to 35 MHz frequency range. The necessary launch vehicle to place the PERCS in orbit will be provided by the Department of Defense Space Test Program. The expandable calibration target has a stowed diameter of 1 meter and a fully deployed diameter of 10.2 meters. A separate deployment mechanism is provided for the sphere. After deployment, the Precision Expandable <span class="hlt">Radar</span> Calibration Sphere (PERCS) with 180 <span class="hlt">vertices</span> will be in a high inclination orbit to scatter radio pulses from a number of ground systems, including (1) over-the-horizon (OTH) <span class="hlt">radars</span> operated by the United States and Australia; (2) high power HF facilities such as HAARP in Alaska, EISCAT in Norway, and Arecibo in Puerto Rico; (3) the chain of high latitude SuperDARN <span class="hlt">radars</span> used for auroral region mapping; and (4) HF direction finding for Navy ships. With the PERCS satellite, the accuracy of HF <span class="hlt">radars</span> can be periodically checked for range, elevation, and azimuth errors. In addition, each of the 360 <span class="hlt">vertices</span> on the PERCS sphere will support an optical retro-reflector for operations with ground laser facilities used to track satellites. The ground laser systems will be used to measure the precise location of the sphere within one cm accuracy and will provide the spatial orientation of the sphere as well as the rotation rate. The Department of Defense facilities that can use the corner-cube reflectors on the PERCS include (1) the Air Force Maui Optical Site (AMOS), (2) the Starfire Optical Range (SOR), and (3) the NRL Optical Test Facility (OTF).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01798&hterms=tailings&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dtailings','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01798&hterms=tailings&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dtailings"><span>Space <span class="hlt">Radar</span> Image of Salt Lake City, Utah</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This <span class="hlt">radar</span> image of Salt Lake City, Utah, illustrates the different land use patterns that are present in the Utah Valley. Salt Lake City lies between the shores of the Great Salt Lake (the dark area on the left side of the image) and the Wasatch Front Range (the mountains in the upper half of the image). The Salt Lake City area is of great interest to urban planners because of the combination of lake, valley and alpine environments that coexist in the region. Much of the southern shore of the Great Salt Lake is a waterfowl management area. The green grid pattern in the right center of the image is Salt Lake City and its surrounding communities. The Salt Lake City airport is visible as the brown rectangle near the center of the image. Interstate Highway 15 runs from the middle right edge to the upper left of the image. The bright white patch east of Interstate 15 is the downtown area, including Temple Square and the state capitol. The University of Utah campus is the yellowish area that lies at the base of the mountains, east of Temple Square. The large reservoir in the lower left center is a mine tailings pond. The semi-circular feature in the mountains at the bottom edge of the image is the Kennecott Copper Mine. The area shown is 60 kilometers by 40 kilometers (37 miles by 25 miles) and is centered at 40.6 degrees north latitude, 112.0 degrees west longitude. North is toward the upper left. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on April 10, 1994. The colors in this image represent the following <span class="hlt">radar</span> channels and polarizations: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01798.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01798.html"><span>Space <span class="hlt">Radar</span> Image of Salt Lake City, Utah</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This <span class="hlt">radar</span> image of Salt Lake City, Utah, illustrates the different land use patterns that are present in the Utah Valley. Salt Lake City lies between the shores of the Great Salt Lake (the dark area on the left side of the image) and the Wasatch Front Range (the mountains in the upper half of the image). The Salt Lake City area is of great interest to urban planners because of the combination of lake, valley and alpine environments that coexist in the region. Much of the southern shore of the Great Salt Lake is a waterfowl management area. The green grid pattern in the right center of the image is Salt Lake City and its surrounding communities. The Salt Lake City airport is visible as the brown rectangle near the center of the image. Interstate Highway 15 runs from the middle right edge to the upper left of the image. The bright white patch east of Interstate 15 is the downtown area, including Temple Square and the state capitol. The University of Utah campus is the yellowish area that lies at the base of the mountains, east of Temple Square. The large reservoir in the lower left center is a mine tailings pond. The semi-circular feature in the mountains at the bottom edge of the image is the Kennecott Copper Mine. The area shown is 60 kilometers by 40 kilometers (37 miles by 25 miles) and is centered at 40.6 degrees north latitude, 112.0 degrees west longitude. North is toward the upper left. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on April 10, 1994. The colors in this image represent the following <span class="hlt">radar</span> channels and polarizations: red is L-band, horizontally transmitted and received; green is L-band, horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program. http</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01781&hterms=Rafael&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DRafael','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01781&hterms=Rafael&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3DRafael"><span>Space <span class="hlt">Radar</span> Image of San Rafael Glacier, Chile</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>A NASA <span class="hlt">radar</span> instrument has been successfully used to measure some of the fastest moving and most inaccessible glaciers in the world -- in Chile's huge, remote Patagonia ice fields -- demonstrating a technique that could produce more accurate predictions of glacial response to climate change and corresponding sea level changes. This image, produced with interferometric measurements made by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) flown on the Space Shuttle last fall, has provided the first detailed measurements of the mass and motion of the San Rafael Glacier. Very few measurements have been made of the Patagonian ice fields, which are the world's largest mid-latitude ice masses and account for more than 60 percent of the Southern Hemisphere's glacial area outside of Antarctica. These features make the area essential for climatologists attempting to understand the response of glaciers on a global scale to changes in climate, but the region's inaccessibility and inhospitable climate have made it nearly impossible for scientists to study its glacial topography, meteorology and changes over time. Currently, topographic data exist for only a few glaciers while no data exist for the vast interior of the ice fields. Velocity has been measured on only five of the more than 100 glaciers, and the data consist of only a few single-<span class="hlt">point</span> measurements. The interferometry performed by the SIR-C/X-SAR was used to generate both a digital elevation model of the glaciers and a map of their ice motion on a pixel-per-pixel basis at very high resolution for the first time. The data were acquired from nearly the same position in space on October 9, 10 and 11, 1994, at L-band frequency (24-cm wavelength), <span class="hlt">vertically</span> transmitted and received polarization, as the Space Shuttle Endeavor flew over several Patagonian outlet glaciers of the San Rafael Laguna. The area shown in these two images is 50 kilometers by 30 kilometers (30 miles by 18 miles) in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999BAMS...80..653T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999BAMS...80..653T"><span>A Single-<span class="hlt">Radar</span> Technique for Estimating the Winds in Tropical Cyclones.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tuttle, John; Gall, Robert</p> <p>1999-04-01</p> <p>A method for determining horizontal wind speeds in hurricanes using ground-based <span class="hlt">radars</span> is presented and evaluated. The method makes use of the tracking reflectivity echos by correlation (TREC) method where individual features in <span class="hlt">radar</span> reflectivity are tracked, from <span class="hlt">radar</span> sweeps several minutes apart, by finding the maxima in the cross-correlation function between the two times. This method has been applied successfully in determining motions within the clear boundary layer where reflectors are insects and refractive index variations, but it generally has failed when applied to determining air motions by tracking precipitation elements in strong environmental shear. It appears to work in the lower few kilometers of the hurricane where the <span class="hlt">vertical</span> wind shear is relatively weak.Examples are presented where the TREC algorithm is applied to three landfalling hurricanes: Hurricanes Hugo and Erin in the United States and Typhoon Herb in Taiwan. The results from Hugo, where the <span class="hlt">radar</span> data were provided by a WSR-57, were compared to in situ wind measurements by the National Oceanic and Atmospheric Administration P-3 research aircraft. In Erin and Herb, Doppler <span class="hlt">radar</span> data are available and the radial winds (with respect to the <span class="hlt">radar</span>) computed by TREC could be compared.The results were very promising. In Hugo, the agreement between the TREC analysis and the aircraft winds was generally to within 10%. In Erin and Herb less than 20% of the difference between radial-Doppler wind estimations by TREC and the actual Doppler wind measurements was greater than 5 m s-1. When Herb was closer to the <span class="hlt">radar</span>, however, the error rates were much higher due to the interference of ground clutter.TREC promises to provide a quick and reasonably accurate method for continuously computing fully two-dimensional winds from land-based <span class="hlt">radars</span> as hurricanes approach the coast. Such information would complement that provided by Doppler <span class="hlt">radars</span> where it could estimate the tangential component to the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01731&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01731&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia"><span>Space <span class="hlt">Radar</span> Image of Kliuchevskoi Volcano, Russia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is an image of the Kliuchevskoi volcano, Kamchatka, Russia, which began to erupt on September 30, 1994. Kliuchevskoi is the bright white peak surrounded by red slopes in the lower left portion of the image. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> aboard the space shuttle Endeavour on its 25th orbit on October 1, 1994. The image shows an area approximately 30 kilometers by 60 kilometers (18.5 miles by 37 miles) that is centered at 56.18 degrees north latitude and 160.78 degrees east longitude. North is toward the top of the image. The Kamchatka volcanoes are among the most active volcanoes in the world. The volcanic zone sits above a tectonic plate boundary, where the Pacific plate is sinking beneath the northeast edge of the Eurasian plate. The Endeavour crew obtained dramatic video and photographic images of this region during the eruption, which will assist scientists in analyzing the dynamics of the current activity. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). The Kamchatka River runs from left to right across the image. An older, dormant volcanic region appears in green on the north side of the river. The current eruption included massive ejections of gas, vapor and ash, which reached altitudes of 20,000 meters (65,000 feet). New lava flows are visible on the flanks of Kliuchevskoi, appearing yellow/green in the image, superimposed on the red surfaces in the lower center. Melting snow triggered mudflows on the north flank of the volcano, which may threaten agricultural zones and other settlements in the valley to the north. Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The <span class="hlt">radars</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01768&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01768&hterms=Sunlight+cities&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3DSunlight%2Bcities"><span>Space <span class="hlt">Radar</span> Image of Taal Volcano, Philippines</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is an image of Taal volcano, near Manila on the island of Luzon in the Philippines. The black area in the center is Taal Lake, which nearly fills the 30-kilometer-diameter (18-mile) caldera. The caldera rim consists of deeply eroded hills and cliffs. The large island in Taal Lake, which itself contains a crater lake, is known as Volcano Island. The bright yellow patch on the southwest side of the island marks the site of an explosion crater that formed during a deadly eruption of Taal in 1965. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 78th orbit on October 5, 1994. The image shows an area approximately 56 kilometers by 112 kilometers (34 miles by 68 miles) that is centered at 14.0 degrees north latitude and 121.0 degrees east longitude. North is toward the upper right of the image. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). Since 1572, Taal has erupted at least 34 times. Since early 1991, the volcano has been restless, with swarms of earthquakes, new steaming areas, ground fracturing, and increases in water temperature of the lake. Volcanologists and other local authorities are carefully monitoring Taal to understand if the current activity may foretell an eruption. Taal is one of 15 'Decade Volcanoes' that have been identified by the volcanology community as presenting large potential hazards to population centers. The bright area in the upper right of the image is the densely populated city of Manila, only 50 kilometers (30 miles) north of the central crater. Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The <span class="hlt">radars</span> illuminate Earth</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.9822R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.9822R"><span>STRING: A new drifter for HF <span class="hlt">radar</span> validation.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rammou, Anna-Maria; Zervakis, Vassilis; Bellomo, Lucio; Kokkini, Zoi; Quentin, Celine; Mantovani, Carlo; Kalampokis, Alkiviadis</p> <p>2015-04-01</p> <p>High-Frequency <span class="hlt">radars</span> (HFR) are an effective mean of remotely monitoring sea-surface currents, based on recording the Doppler-shift of radio-waves backscattered on the sea surface. Validation of HFRs' measurements takes place via comparisons either with in-situ Eulerian velocity data (usually obtained by surface current-meters attached on moorings) or to Lagrangian velocity fields (recorded by surface drifters). The most common surface drifter used for this purpose is the CODE-type drifter (Davis, 1985), an industry-standard design to record the <span class="hlt">vertical</span> average velocity of the upper 1 m layer of the water column. In this work we claim that the observed differences between the HFR-derived velocities and Lagrangian measurements can be attributed not just to the different spatial scales recorded by the above instruments but also due to the fact that while the HFR-derived velocity corresponds to exponentially weighted <span class="hlt">vertical</span> average of the velocity field from the surface to 1 m depth (Stewart and Joy, 1974) the velocity estimated by the CODE drifters corresponds to boxcar-type weighted <span class="hlt">vertical</span> average due to the orthogonal shape of the CODE drifters' sails. After analyzing the theoretical behavior of a drifter under the influence of wind and current, we proceed to propose a new design of exponentially-shaped sails for the drogues of CODE-based drifters, so that the HFR-derived velocities and the drifter-based velocities will be directly comparable, regarding the way of <span class="hlt">vertically</span> averaging the velocity field.The new drifter, codenamed STRING, exhibits identical behavior to the classical CODE design under relatively homogeneous conditions in the upper 1 m layer, however it is expected to follow a significantly different track in conditions of high <span class="hlt">vertical</span> shear and stratification. Thus, we suggest that the new design is the instrument of choice for validation of HFR installations, as it can be used in all conditions and behaves identically to CODEs when <span class="hlt">vertical</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017Ap%26SS.362..136Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Ap%26SS.362..136Q"><span>Analytical and numerical construction of <span class="hlt">vertical</span> periodic orbits about triangular libration <span class="hlt">points</span> based on polynomial expansion relations among directions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qian, Ying-Jing; Yang, Xiao-Dong; Zhai, Guan-Qiao; Zhang, Wei</p> <p>2017-08-01</p> <p>Innovated by the nonlinear modes concept in the vibrational dynamics, the <span class="hlt">vertical</span> periodic orbits around the triangular libration <span class="hlt">points</span> are revisited for the Circular Restricted Three-body Problem. The ζ -component motion is treated as the dominant motion and the ξ and η -component motions are treated as the slave motions. The slave motions are in nature related to the dominant motion through the approximate nonlinear polynomial expansions with respect to the ζ -position and ζ -velocity during the one of the periodic orbital motions. By employing the relations among the three directions, the three-dimensional system can be transferred into one-dimensional problem. Then the approximate three-dimensional <span class="hlt">vertical</span> periodic solution can be analytically obtained by solving the dominant motion only on ζ -direction. To demonstrate the effectiveness of the proposed method, an accuracy study was carried out to validate the polynomial expansion (PE) method. As one of the applications, the invariant nonlinear relations in polynomial expansion form are used as constraints to obtain numerical solutions by differential correction. The nonlinear relations among the directions provide an alternative <span class="hlt">point</span> of view to explore the overall dynamics of periodic orbits around libration <span class="hlt">points</span> with general rules.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JGRD..118.1814C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JGRD..118.1814C"><span>Empirical conversion of the <span class="hlt">vertical</span> profile of reflectivity from Ku-band to S-band frequency</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Qing; Hong, Yang; Qi, Youcun; Wen, Yixin; Zhang, Jian; Gourley, Jonathan J.; Liao, Liang</p> <p>2013-02-01</p> <p>ABSTRACT This paper presents an empirical method for converting reflectivity from Ku-band (13.8 GHz) to S-band (2.8 GHz) for several hydrometeor species, which facilitates the incorporation of Tropical Rainfall Measuring Mission (TRMM) Precipitation <span class="hlt">Radar</span> (PR) measurements into quantitative precipitation estimation (QPE) products from the U.S. Next-Generation <span class="hlt">Radar</span> (NEXRAD). The development of empirical dual-frequency relations is based on theoretical simulations, which have assumed appropriate scattering and microphysical models for liquid and solid hydrometeors (raindrops, snow, and ice/hail). Particle phase, shape, orientation, and density (especially for snow particles) have been considered in applying the T-matrix method to compute the scattering amplitudes. Gamma particle size distribution (PSD) is utilized to model the microphysical properties in the ice region, melting layer, and raining region of precipitating clouds. The variability of PSD parameters is considered to study the characteristics of dual-frequency reflectivity, especially the variations in <span class="hlt">radar</span> dual-frequency ratio (DFR). The empirical relations between DFR and Ku-band reflectivity have been derived for particles in different regions within the <span class="hlt">vertical</span> structure of precipitating clouds. The reflectivity conversion using the proposed empirical relations has been tested using real data collected by TRMM-PR and a prototype polarimetric WSR-88D (Weather Surveillance <span class="hlt">Radar</span> 88 Doppler) <span class="hlt">radar</span>, KOUN. The processing and analysis of collocated data demonstrate the validity of the proposed empirical relations and substantiate their practical significance for reflectivity conversion, which is essential to the TRMM-based <span class="hlt">vertical</span> profile of reflectivity correction approach in improving NEXRAD-based QPE.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9986E..04G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9986E..04G"><span>Multibeam monopulse <span class="hlt">radar</span> for airborne sense and avoid system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gorwara, Ashok; Molchanov, Pavlo</p> <p>2016-10-01</p> <p>The multibeam monopulse <span class="hlt">radar</span> for Airborne Based Sense and Avoid (ABSAA) system concept is the next step in the development of passive monopulse direction finder proposed by Stephen E. Lipsky in the 80s. In the proposed system the multibeam monopulse <span class="hlt">radar</span> with an array of directional antennas is positioned on a small aircaraft or Unmanned Aircraft System (UAS). <span class="hlt">Radar</span> signals are simultaneously transmitted and received by multiple angle shifted directional antennas with overlapping antenna patterns and the entire sky, 360° for both horizontal and <span class="hlt">vertical</span> coverage. Digitizing of amplitude and phase of signals in separate directional antennas relative to reference signals provides high-accuracy high-resolution range and azimuth measurement and allows to record real time amplitude and phase of reflected from non-cooperative aircraft signals. High resolution range and azimuth measurement provides minimal tracking errors in both position and velocity of non-cooperative aircraft and determined by sampling frequency of the digitizer. High speed sampling with high-accuracy processor clock provides high resolution phase/time domain measurement even for directional antennas with wide Field of View (FOV). Fourier transform (frequency domain processing) of received <span class="hlt">radar</span> signals provides signatures and dramatically increases probability of detection for non-cooperative aircraft. Steering of transmitting power and integration, correlation period of received reflected signals for separate antennas (directions) allows dramatically decreased ground clutter for low altitude flights. An open architecture, modular construction allows the combination of a <span class="hlt">radar</span> sensor with Automatic Dependent Surveillance - Broadcast (ADS-B), electro-optic, acoustic sensors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFM.C23B0753Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFM.C23B0753Y"><span>Ground-Truthing a Next Generation Snow <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yan, S.; Brozena, J. M.; Gogineni, P. S.; Abelev, A.; Gardner, J. M.; Ball, D.; Liang, R.; Newman, T.</p> <p>2016-12-01</p> <p>During the early spring of 2016 the Naval Research Laboratory (NRL) performed a test of a next generation airborne snow <span class="hlt">radar</span> over ground truth data collected on several areas of fast ice near Barrow, AK. The <span class="hlt">radar</span> was developed by the Center for Remote Sensing of Ice Sheets (CReSIS) at the University of Kansas, and includes several improvements compared to their previous snow <span class="hlt">radar</span>. The new unit combines the earlier Ku-band and snow <span class="hlt">radars</span> into a single unit with an operating frequency spanning the entire 2-18 GHz, an enormous bandwidth which provides the possibility of snow depth measurements with 1.5 cm range resolution. Additionally, the <span class="hlt">radar</span> transmits on dual polarizations (H and V), and receives the signal through two orthogonally polarized Vivaldi arrays, each with 128 phase centers. The 8 sets of along-track phase centers are combined in hardware to improve SNR and narrow the beamwidth in the along-track, resulting in 8 cross-track effective phase centers which are separately digitized to allow for beam sharpening and forming in post-processing. Tilting the receive arrays 30 degrees from the horizontal also allows the formation of SAR images and the potential for estimating snow-water equivalent (SWE). Ground truth data (snow depth, density, salinity and SWE) were collected over several 60 m wide swaths that were subsequently overflown with the snow <span class="hlt">radar</span> mounted on a Twin Otter. The <span class="hlt">radar</span> could be operated in nadir (by beam steering the receive antennas to <span class="hlt">point</span> beneath the aircraft) or side-looking modes. Results from the comparisons will be shown.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.fs.usda.gov/treesearch/pubs/54587','TREESEARCH'); return false;" href="https://www.fs.usda.gov/treesearch/pubs/54587"><span>Tropical-Forest Structure and Biomass Dynamics from TanDEM-X <span class="hlt">Radar</span> Interferometry</span></a></p> <p><a target="_blank" href="http://www.fs.usda.gov/treesearch/">Treesearch</a></p> <p>Robert Treuhaft; Yang Lei; Fabio Gonçalves; Michael Keller; João Santos; Maxim Neumann; André Almeida</p> <p>2017-01-01</p> <p>Changes in tropical-forest structure and aboveground biomass (AGB) contribute directly to atmospheric changes in CO2, which, in turn, bear on global climate. This paper demonstrates the capability of <span class="hlt">radar</span>-interferometric phase-height time series at X-band (wavelength = 3 cm) to monitor changes in <span class="hlt">vertical</span> structure and AGB, with sub-hectare and monthly spatial and...</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/2017AGUFM.A21P..07K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A21P..07K"><span>Phase-partitioning in mixed-phase clouds - An approach to characterize the entire <span class="hlt">vertical</span> column</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kalesse, H.; Luke, E. P.; Seifert, P.</p> <p>2017-12-01</p> <p>The characterization of the entire <span class="hlt">vertical</span> profile of phase-partitioning in mixed-phase clouds is a challenge which can be addressed by synergistic profiling measurements with ground-based polarization lidars and cloud <span class="hlt">radars</span>. While lidars are sensitive to small particles and can thus detect supercooled liquid (SCL) layers, cloud <span class="hlt">radar</span> returns are dominated by larger particles (like ice crystals). The maximum lidar observation height is determined by complete signal attenuation at a penetrated optical depth of about three. In contrast, cloud <span class="hlt">radars</span> are able to penetrate multiple liquid layers and can thus be used to expand the identification of cloud phase to the entire <span class="hlt">vertical</span> column beyond the lidar extinction height, if morphological features in the <span class="hlt">radar</span> Doppler spectrum can be related to the existence of SCL. Relevant spectral signatures such as bimodalities and spectral skewness can be related to cloud phase by training a neural network appropriately in a supervised learning scheme, with lidar measurements functioning as supervisor. The neural network output (prediction of SCL location) derived using cloud <span class="hlt">radar</span> Doppler spectra can be evaluated with several parameters such as liquid water path (LWP) detected by microwave radiometer (MWR) and (liquid) cloud base detected by ceilometer or Raman lidar. The technique has been previously tested on data from Department of Energy (DOE) Atmospheric Radiation Measurement (ARM) instruments in Barrow, Alaska and is in this study utilized for observations from the Leipzig Aerosol and Cloud Remote Observations System (LACROS) during the Analysis of the Composition of Clouds with Extended Polarization Techniques (ACCEPT) field experiment in Cabauw, Netherlands in Fall 2014. Comparisons to supercooled-liquid layers as classified by CLOUDNET are provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01731.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01731.html"><span>Space <span class="hlt">Radar</span> Image of Kliuchevskoi Volcano, Russia</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This is an image of the Kliuchevskoi volcano, Kamchatka, Russia, which began to erupt on September 30, 1994. Kliuchevskoi is the bright white peak surrounded by red slopes in the lower left portion of the image. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> aboard the space shuttle Endeavour on its 25th orbit on October 1, 1994. The image shows an area approximately 30 kilometers by 60 kilometers (18.5 miles by 37 miles) that is centered at 56.18 degrees north latitude and 160.78 degrees east longitude. North is toward the top of the image. The Kamchatka volcanoes are among the most active volcanoes in the world. The volcanic zone sits above a tectonic plate boundary, where the Pacific plate is sinking beneath the northeast edge of the Eurasian plate. The Endeavour crew obtained dramatic video and photographic images of this region during the eruption, which will assist scientists in analyzing the dynamics of the current activity. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). The Kamchatka River runs from left to right across the image. An older, dormant volcanic region appears in green on the north side of the river. The current eruption included massive ejections of gas, vapor and ash, which reached altitudes of 20,000 meters (65,000 feet). New lava flows are visible on the flanks of Kliuchevskoi, appearing yellow/green in the image, superimposed on the red surfaces in the lower center. Melting snow triggered mudflows on the north flank of the volcano, which may threaten agricultural zones and other settlements in the valley to the north. http://photojournal.jpl.nasa.gov/catalog/PIA01731</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01768.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01768.html"><span>Space <span class="hlt">Radar</span> Image of Taal Volcano, Philippines</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This is an image of Taal volcano, near Manila on the island of Luzon in the Philippines. The black area in the center is Taal Lake, which nearly fills the 30-kilometer-diameter (18-mile) caldera. The caldera rim consists of deeply eroded hills and cliffs. The large island in Taal Lake, which itself contains a crater lake, is known as Volcano Island. The bright yellow patch on the southwest side of the island marks the site of an explosion crater that formed during a deadly eruption of Taal in 1965. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 78th orbit on October 5, 1994. The image shows an area approximately 56 kilometers by 112 kilometers (34 miles by 68 miles) that is centered at 14.0 degrees north latitude and 121.0 degrees east longitude. North is toward the upper right of the image. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). Since 1572, Taal has erupted at least 34 times. Since early 1991, the volcano has been restless, with swarms of earthquakes, new steaming areas, ground fracturing, and increases in water temperature of the lake. Volcanologists and other local authorities are carefully monitoring Taal to understand if the current activity may foretell an eruption. Taal is one of 15 "Decade Volcanoes" that have been identified by the volcanology community as presenting large potential hazards to population centers. The bright area in the upper right of the image is the densely populated city of Manila, only 50 kilometers (30 miles) north of the central crater. http://photojournal.jpl.nasa.gov/catalog/PIA01768</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ThApC.tmp..256K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ThApC.tmp..256K"><span><span class="hlt">Vertical</span> structure of precipitating shallow echoes observed from TRMM during Indian summer monsoon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Shailendra</p> <p>2017-08-01</p> <p>The present study explores the properties of precipitating shallow echoes (PSEs) over the tropical areas (30°S-30°N) during Indian summer monsoon season using attenuated corrected <span class="hlt">radar</span> reflectivity factor (Ze) measured by the Tropical Rainfall Measuring Mission satellite. <span class="hlt">Radar</span> echoes observed in study are less than the freezing height, so they belong to warm precipitation. <span class="hlt">Radar</span> echoes with at least 0.75 km wide are considered for finding the shallow echoes climatology. Western Ghats and adjoining ocean (Arabian sea) have the highest PSEs followed by Myanmar and Burma coast, whereas the overall west coast of Latin America consists of the lowest PSEs. Tropical oceanic areas contain fewer PSEs compared to coastal areas. Average <span class="hlt">vertical</span> profiles show nearly similar Ze characteristics which peaks between 1.5 and 2 km altitude with model value 32-34 dBZ. Slope of Ze is higher for intense PSEs as <span class="hlt">radar</span> reflectivity decreases more rapidly in intense PSEs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225461p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225461p/"><span>View of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Transmitter ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>View of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Transmitter Building foundation, showing Fire Control Stations (Buildings 621 and 622) and concrete stairway (top left) camera facing southwest - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340242p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340242p/"><span>4. VIEW NORTHEAST, <span class="hlt">radar</span> tower (unknown function), prime search <span class="hlt">radar</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>4. VIEW NORTHEAST, <span class="hlt">radar</span> tower (unknown function), prime search <span class="hlt">radar</span> tower, emergency power building, and height finder <span class="hlt">radar</span> tower - Fort Custer Military Reservation, P-67 <span class="hlt">Radar</span> Station, .25 mile north of Dickman Road, east of Clark Road, Battle Creek, Calhoun County, MI</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017P%26SS..143..199H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017P%26SS..143..199H"><span>AMSNEXRAD-Automated detection of meteorite strewnfields in doppler weather <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hankey, Michael; Fries, Marc; Matson, Rob; Fries, Jeff</p> <p>2017-09-01</p> <p>For several years meteorite recovery in the United States has been greatly enhanced by using Doppler weather <span class="hlt">radar</span> images to determine possible fall zones for meteorites produced by witnessed fireballs. While most fireball events leave no record on the Doppler <span class="hlt">radar</span>, some large fireballs do. Based on the successful recovery of 10 meteorite falls 'under the <span class="hlt">radar</span>', and the discovery of <span class="hlt">radar</span> on more than 10 historic falls, it is believed that meteoritic dust and or actual meteorites falling to the ground have been recorded on Doppler weather <span class="hlt">radar</span> (Fries et al., 2014). Up until this <span class="hlt">point</span>, the process of detecting the <span class="hlt">radar</span> signatures associated with meteorite falls has been a manual one and dependent on prior accurate knowledge of the fall time and estimated ground track. This manual detection process is labor intensive and can take several hours per event. Recent technological developments by NOAA now help enable the automation of these tasks. This in combination with advancements by the American Meteor Society (Hankey et al., 2014) in the tracking and plotting of witnessed fireballs has opened the possibility for automatic detection of meteorites in NEXRAD <span class="hlt">Radar</span> Archives. Here in the processes for fireball triangulation, search area determination, <span class="hlt">radar</span> interfacing, data extraction, storage, search, detection and plotting are explained.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340243p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340243p/"><span>5. VIEW EAST, height finder <span class="hlt">radar</span> towers, <span class="hlt">radar</span> tower (unknown ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>5. VIEW EAST, height finder <span class="hlt">radar</span> towers, <span class="hlt">radar</span> tower (unknown function), prime search <span class="hlt">radar</span> tower, operations building, and central heating plant - Fort Custer Military Reservation, P-67 <span class="hlt">Radar</span> Station, .25 mile north of Dickman Road, east of Clark Road, Battle Creek, Calhoun County, MI</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01767&hterms=water+villages&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bvillages','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01767&hterms=water+villages&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dwater%2Bvillages"><span>Space <span class="hlt">Radar</span> Image of Rabaul Volcano, New Guinea</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is a <span class="hlt">radar</span> image of the Rabaul volcano on the island of New Britain, Papua, New Guinea taken almost a month after its September 19, 1994, eruption that killed five people and covered the town of Rabaul and nearby villages with up to 75 centimeters (30 inches) of ash. More than 53,000 people have been displaced by the eruption. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 173rd orbit on October 11, 1994. This image is centered at 4.2 degrees south latitude and 152.2 degrees east longitude in the southwest Pacific Ocean. The area shown is approximately 21 kilometers by 25 kilometers (13 miles by 15.5 miles). North is toward the upper right. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). Most of the Rabaul volcano is underwater and the caldera (crater) creates Blanche Bay, the semi-circular body of water that occupies most of the center of the image. Volcanic vents within the caldera are visible in the image and include Vulcan, on a peninsula on the west side of the bay, and Rabalanakaia and Tavurvur (the circular purple feature near the mouth of the bay) on the east side. Both Vulcan and Tavurvur were active during the 1994 eruption. Ash deposits appear red-orange on the image, and are most prominent on the south flanks of Vulcan and north and northwest of Tavurvur. A faint blue patch in the water in the center of the image is a large raft of floating pumice fragments that were ejected from Vulcan during the eruption and clog the inner bay. Visible on the east side of the bay are the grid-like patterns of the streets of Rabaul and an airstrip, which appears as a dark northwest-trending band at the right-center of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225459p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225459p/"><span>Contextual view of <span class="hlt">Point</span> Bonita Ridge, showing Bonita Ridge access ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Contextual view of <span class="hlt">Point</span> Bonita Ridge, showing Bonita Ridge access road retaining wall and location of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Transmitter Building foundation (see stake at center left), camera facing north - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140007333','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140007333"><span>Assimilation of Dual-Polarimetric <span class="hlt">Radar</span> Observations with WRF GSI</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Li, Xuanli; Mecikalski, John; Fehnel, Traci; Zavodsky, Bradley; Srikishen, Jayanthi</p> <p>2014-01-01</p> <p>Dual-polarimetric (dual-pol) <span class="hlt">radar</span> typically transmits both horizontally and <span class="hlt">vertically</span> polarized radio wave pulses. From the two different reflected power returns, more accurate estimate of liquid and solid cloud and precipitation can be provided. The upgrade of the traditional NWS WSR-88D <span class="hlt">radar</span> to include dual-pol capabilities will soon be completed for the entire NEXRAD network. Therefore, the use of dual-pol <span class="hlt">radar</span> network will have a broad impact in both research and operational communities. The assimilation of dual-pol <span class="hlt">radar</span> data is especially challenging as few guidelines have been provided by previous research. It is our goal to examine how to best use dual-pol <span class="hlt">radar</span> data to improve forecast of severe storm and forecast initialization. In recent years, the Development Testbed Center (DTC) has released the community Gridpoint Statistical Interpolation (GSI) DA system for the Weather Research and Forecasting (WRF) model. The community GSI system runs in independently environment, yet works functionally equivalent to operational centers. With collaboration with the NASA Short-term Prediction Research and Transition (SPoRT) Center, this study explores regional assimilation of the dual-pol <span class="hlt">radar</span> variables from the WSR-88D <span class="hlt">radars</span> for real case storms. Our presentation will highlight our recent effort on incorporating the horizontal reflectivity (ZH), differential reflectivity (ZDR), specific differential phase (KDP), and radial velocity (VR) data for initializing convective storms, with a significant focus being on an improved representation of hydrometeor fields. In addition, discussion will be provided on the development of enhanced assimilation procedures in the GSI system with respect to dual-pol variables. Beyond the dual-pol variable assimilation procedure developing within a GSI framework, highresolution (=1 km) WRF model simulations and storm scale data assimilation experiments will be examined, emphasizing both model initialization and short-term forecast</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01788.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01788.html"><span>Space <span class="hlt">Radar</span> Image of Harvard Forest</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This is a <span class="hlt">radar</span> image of the area surrounding the Harvard Forest in north-central Massachusetts that has been operated as a ecological research facility by Harvard University since 1907. At the center of the image is the Quabbin Reservoir, and the Connecticut River is at the lower left of the image. The Harvard Forest itself is just above the reservoir. Researchers are comparing the naturally occurring physical disturbances in the forest and the recent and projected chemical disturbances and their effects on the forest ecosystem. Agricultural land appears dark blue/purple, along with low shrub vegetation and some wetlands. Urban development is bright pink; the yellow to green tints are conifer-dominated vegetation with the pitch pine sand plain at the middle left edge of the image appearing very distinctive. The green tint may indicate pure pine plantation stands, and deciduous broadleaf trees appear gray/pink with perhaps wetter sites being pinker. This image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and the United States space agencies, is part of NASA's Mission to Planet Earth. The image is centered at 42.50 degrees North latitude and 72.33 degrees West longitude and covers an area of 53 kilometers 63 by kilometers (33 miles by 39 miles). The colors in the image are assigned to different frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band horizontally transmitted and horizontally received; green is L-band horizontally transmitted and <span class="hlt">vertically</span> received; and blue is C-band horizontally transmitted and horizontally received. http://photojournal.jpl.nasa.gov/catalog/PIA01788</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21164990','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21164990"><span>Terahertz <span class="hlt">radar</span> cross section measurements.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Iwaszczuk, Krzysztof; Heiselberg, Henning; Jepsen, Peter Uhd</p> <p>2010-12-06</p> <p>We perform angle- and frequency-resolved <span class="hlt">radar</span> cross section (RCS) measurements on objects at terahertz frequencies. Our RCS measurements are performed on a scale model aircraft of size 5-10 cm in polar and azimuthal configurations, and correspond closely to RCS measurements with conventional <span class="hlt">radar</span> on full-size objects. The measurements are performed in a terahertz time-domain system with freely propagating terahertz pulses generated by tilted pulse front excitation of lithium niobate crystals and measured with sub-picosecond time resolution. The application of a time domain system provides ranging information and also allows for identification of scattering <span class="hlt">points</span> such as weaponry attached to the aircraft. The shapes of the models and positions of reflecting parts are retrieved by the filtered back projection algorithm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01767.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01767.html"><span>Space <span class="hlt">Radar</span> Image of Rabaul Volcano, New Guinea</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>This is a <span class="hlt">radar</span> image of the Rabaul volcano on the island of New Britain, Papua, New Guinea taken almost a month after its September 19, 1994, eruption that killed five people and covered the town of Rabaul and nearby villages with up to 75 centimeters (30 inches) of ash. More than 53,000 people have been displaced by the eruption. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 173rd orbit on October 11, 1994. This image is centered at 4.2 degrees south latitude and 152.2 degrees east longitude in the southwest Pacific Ocean. The area shown is approximately 21 kilometers by 25 kilometers (13 miles by 15.5 miles). North is toward the upper right. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). Most of the Rabaul volcano is underwater and the caldera (crater) creates Blanche Bay, the semi-circular body of water that occupies most of the center of the image. Volcanic vents within the caldera are visible in the image and include Vulcan, on a peninsula on the west side of the bay, and Rabalanakaia and Tavurvur (the circular purple feature near the mouth of the bay) on the east side. Both Vulcan and Tavurvur were active during the 1994 eruption. Ash deposits appear red-orange on the image, and are most prominent on the south flanks of Vulcan and north and northwest of Tavurvur. A faint blue patch in the water in the center of the image is a large raft of floating pumice fragments that were ejected from Vulcan during the eruption and clog the inner bay. Visible on the east side of the bay are the grid-like patterns of the streets of Rabaul and an airstrip, which appears as a dark northwest-trending band at the right-center of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017SPIE10188E..16N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017SPIE10188E..16N"><span><span class="hlt">Radar</span> research at The Pennsylvania State University <span class="hlt">Radar</span> and Communications Laboratory</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Narayanan, Ram M.</p> <p>2017-05-01</p> <p>The <span class="hlt">Radar</span> and Communications Laboratory (RCL) at The Pennsylvania State University is at the forefront of <span class="hlt">radar</span> technology and is engaged in cutting edge research in all aspects of <span class="hlt">radar</span>, including modeling and simulation studies of novel <span class="hlt">radar</span> paradigms, design and development of new types of <span class="hlt">radar</span> architectures, and extensive field measurements in realistic scenarios. This paper summarizes the research at The Pennsylvania State University's <span class="hlt">Radar</span> and Communications Laboratory and relevant collaborative research with several groups over the past 15 years in the field of <span class="hlt">radar</span> and related technologies, including communications, radio frequency identification (RFID), and spectrum sensing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225463p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225463p/"><span>Detail view of northwest side of Signal Corps <span class="hlt">Radar</span> (S.C.R.) ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Detail view of northwest side of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 Transmitter Building foundation, showing portion of concrete gutter drainage system and asphalt floor tiles, camera facing north - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01765&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01765&hterms=Russia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DRussia"><span>Space <span class="hlt">Radar</span> Image of Kiluchevskoi, Volcano, Russia</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This is an image of the area of Kliuchevskoi volcano, Kamchatka, Russia, which began to erupt on September 30, 1994. Kliuchevskoi is the blue triangular peak in the center of the image, towards the left edge of the bright red area that delineates bare snow cover. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 88th orbit on October 5, 1994. The image shows an area approximately 75 kilometers by 100 kilometers (46 miles by 62 miles) that is centered at 56.07 degrees north latitude and 160.84 degrees east longitude. North is toward the bottom of the image. The <span class="hlt">radar</span> illumination is from the top of the image. The Kamchatka volcanoes are among the most active volcanoes in the world. The volcanic zone sits above a tectonic plate boundary, where the Pacific plate is sinking beneath the northeast edge of the Eurasian plate. The Endeavour crew obtained dramatic video and photographic images of this region during the eruption, which will assist scientists in analyzing the dynamics of the recent activity. The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received); blue represents the C-band (horizontally transmitted and <span class="hlt">vertically</span> received). In addition to Kliuchevskoi, two other active volcanoes are visible in the image. Bezymianny, the circular crater above and to the right of Kliuchevskoi, contains a slowly growing lava dome. Tolbachik is the large volcano with a dark summit crater near the upper right edge of the red snow covered area. The Kamchatka River runs from right to left across the bottom of the image. The current eruption of Kliuchevskoi included massive ejections of gas, vapor and ash, which reached altitudes of 15,000 meters (50,000 feet). Melting snow mixed with volcanic ash triggered mud flows on the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890020486','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890020486"><span>Equatorial <span class="hlt">radar</span> system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rukao, S.; Tsuda, T.; Sato, T.; Kato, S.</p> <p>1989-01-01</p> <p>A large clear air <span class="hlt">radar</span> with the sensitivity of an incoherent scatter <span class="hlt">radar</span> for observing the whole equatorial atmosphere up to 1000 km altitude is now being designed in Japan. The <span class="hlt">radar</span>, called the Equatorial <span class="hlt">Radar</span>, will be built in Pontianak, Kalimantan Island, Indonesia (0.03 N, 109.3 E). The system is a 47 MHz monostatic Doppler <span class="hlt">radar</span> with an active phased array configuration similar to that of the MU <span class="hlt">radar</span> in Japan, which has been in successful operation since 1983. It will have a PA product of more than 5 x 10(9) sq. Wm (P = average transmitter power, A = effective antenna aperture) with sensitivity more than 10 times that of the MU <span class="hlt">radar</span>. This system configuration enables pulse-to-pulse beam steering within 25 deg from the zenith. As is the case of the MU <span class="hlt">radar</span>, a variety of sophisticated operations will be made feasible under the supervision of the <span class="hlt">radar</span> controller. A brief description of the system configuration is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030061424&hterms=relationships&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D80%26Ntt%3Drelationships','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030061424&hterms=relationships&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DTitle%26N%3D0%26No%3D80%26Ntt%3Drelationships"><span>Relationships between Electrical and <span class="hlt">Radar</span> Characteristics of Thunderstorms Observed During ACES</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Buechler, Dennis E.; Mach, Douglas M.; Blakeslee, Richard J.</p> <p>2003-01-01</p> <p>The Altus Cumulus Electrification Study (ACES) took place near Key West, Florida during August 2002. A high altitude, remotely piloted aircraft obtained optical pulse and electric field data over a number of thunderstorms during the study period. Measurements of the <span class="hlt">vertical</span> electric field and cross sections of <span class="hlt">radar</span> reflectivity along the flight track are shown for 2 overpasses of a thunderstorm that occurred on 10 August 2002.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A11A3011C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A11A3011C"><span>Architectures for Rainfall Property Estimation From Polarimetric <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Collis, S. M.; Giangrande, S. E.; Helmus, J.; Troemel, S.</p> <p>2014-12-01</p> <p><span class="hlt">Radars</span> that transmit and receive signals in polarizations aligned both horizontal and <span class="hlt">vertical</span> to the horizon collect a number of measurements. The relation both between these measurements and between measurements and desired microphysical quantities (such as rainfall rate) is complicated due to a number of scattering mechanisms. The result is that there ends up being an intractable number of often incompatible techniques for extracting geophysical insight. This presentation will discuss methods developed by the Atmospheric Measurement Climate (ARM) Research Facility to streamline the creation of application chains for retrieving rainfall properties for the purposes of fine scale model evaluation. By using a Common Data Model (CDM) approach and working in the popular open source Python scientific environment analysis techniques such as Linear Programming (LP) can be bought to bear on the task of retrieving insight from <span class="hlt">radar</span> signals. This presentation will outline how we have used these techniques to detangle polarimetric phase signals, estimate a three-dimensional precipitation field and then objectively compare to cloud resolving model derived rainfall fields from the NASA/DoE Mid-Latitude Continental Convective Clouds Experiment (MC3E). All techniques show will be available, open source, in the Python-ARM <span class="hlt">Radar</span> Toolkit (Py-ART).</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01726&hterms=silk+road&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dsilk%2Broad','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01726&hterms=silk+road&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dsilk%2Broad"><span>Space <span class="hlt">Radar</span> Image of the Silk route in Niya, Taklamak, China</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1999-01-01</p> <p>This composite image is of an area thought to contain the ruins of the ancient settlement of Niya. It is located in the southwest corner of the Taklamakan Desert in China's Sinjiang Province. This region was part of some of China's earliest dynasties and from the third century BC on was traversed by the famous Silk Road. The Silk Road, passing east-west through this image, was an ancient trade route that led across Central Asia's desert to Persia, Byzantium and Rome. The multi-frequency, multi-polarized <span class="hlt">radar</span> imagery was acquired on orbit 106 of the space shuttle Endeavour on April 16, 1994 by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span>. The image is centered at 37.78 degrees north latitude and 82.41 degrees east longitude. The area shown is approximately 35 kilometers by 83 kilometers (22 miles by 51 miles). The image is a composite of an image from an Earth-orbiting satellite called Systeme Probatoire d'Observation de la Terre (SPOT)and a SIR-C multi-frequency, multi-polarized <span class="hlt">radar</span> image. The false-color <span class="hlt">radar</span> image was created by displaying the C-band (horizontally transmitted and received) return in red, the L-band (horizontally transmitted and received) return in green, and the L-band (horizontally transmitted and <span class="hlt">vertically</span> received) return in blue. The prominent east/west pink formation at the bottom of the image is most likely a ridge of loosely consolidated sedimentary rock. The Niya River -- the black feature in the lower right of the French satellite image -- meanders north-northeast until it clears the sedimentary ridge, at which <span class="hlt">point</span> it abruptly turns northwest. Sediment and evaporite deposits left by the river over millennia dominate the center and upper right of the <span class="hlt">radar</span> image (in light pink). High ground, ridges and dunes are seen among the riverbed meanderings as mottled blue. Through image enhancement and analysis, a new feature probably representing a man-made canal has been discovered and mapped. Spaceborne Imaging <span class="hlt">Radar</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040040149','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040040149"><span>The 94 GHz Cloud <span class="hlt">Radar</span> System on a NASA ER-2 Aircraft</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Li, Lihua; Heymsfield, Gerald M.; Racette, Paul E.; Tian, Lin; Zenker, Ed</p> <p>2003-01-01</p> <p>The 94-GHz (W-band) Cloud <span class="hlt">Radar</span> System (CRS) has been developed and flown on a NASA ER-2 high-altitude (20 km) aircraft. The CRS is a fully coherent, polarimeteric Doppler <span class="hlt">radar</span> that is capable of detecting clouds and precipitation from the surface up to the aircraft altitude in the lower stratosphere. The <span class="hlt">radar</span> is especially well suited for cirrus cloud studies because of its high sensitivity and fine spatial resolution. This paper describes the CRS motivation, instrument design, specifications, calibration, and preliminary data &om NASA s Cirrus Regional Study of Tropical Anvils and Cirrus Layers - Florida Area Cirrus Experiment (CRYSTAL-FACE) field campaign. The unique combination of CRS with other sensors on the ER-2 provides an unprecedented opportunity to study cloud radiative effects on the global energy budget. CRS observations are being used to improve our knowledge of atmospheric scattering and attenuation characteristics at 94 GHz, and to provide datasets for algorithm implementation and validation for the upcoming NASA CloudSat mission that will use a 94-GHz spaceborne cloud <span class="hlt">radar</span> to provide the first direct global survey of the <span class="hlt">vertical</span> structure of cloud systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01735.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01735.html"><span>Space <span class="hlt">Radar</span> Image of Manaus, Brazil</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p>These two false-color images of the Manaus region of Brazil in South America were acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> on board the space shuttle Endeavour. The image at left was acquired on April 12, 1994, and the image at right was acquired on October 3, 1994. The area shown is approximately 8 kilometers by 40 kilometers (5 miles by 25 miles). The two large rivers in this image, the Rio Negro (at top) and the Rio Solimoes (at bottom), combine at Manaus (west of the image) to form the Amazon River. The image is centered at about 3 degrees south latitude and 61 degrees west longitude. North is toward the top left of the images. The false colors were created by displaying three L-band polarization channels: red areas correspond to high backscatter, horizontally transmitted and received, while green areas correspond to high backscatter, horizontally transmitted and <span class="hlt">vertically</span> received. Blue areas show low returns at <span class="hlt">vertical</span> transmit/receive polarization; hence the bright blue colors of the smooth river surfaces can be seen. Using this color scheme, green areas in the image are heavily forested, while blue areas are either cleared forest or open water. The yellow and red areas are flooded forest or floating meadows. The extent of the flooding is much greater in the April image than in the October image and appears to follow the 10-meter (33-foot) annual rise and fall of the Amazon River. The flooded forest is a vital habitat for fish, and floating meadows are an important source of atmospheric methane. These images demonstrate the capability of SIR-C/X-SAR to study important environmental changes that are impossible to see with optical sensors over regions such as the Amazon, where frequent cloud cover and dense forest canopies block monitoring of flooding. Field studies by boat, on foot and in low-flying aircraft by the University of California at Santa Barbara, in collaboration with Brazil's Instituto Nacional de Pesguisas</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110009939','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110009939"><span>Quality Control and Calibration of the Dual-Polarization <span class="hlt">Radar</span> at Kwajalein, RMI</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Marks, David A.; Wolff, David B.; Carey, Lawrence D.; Tokay, Ali</p> <p>2010-01-01</p> <p>Weather <span class="hlt">radars</span>, recording information about precipitation around the globe, will soon be significantly upgraded. Most of today s weather <span class="hlt">radars</span> transmit and receive microwave energy with horizontal orientation only, but upgraded systems have the capability to send and receive both horizontally and <span class="hlt">vertically</span> oriented waves. These enhanced "dual-polarimetric" (DP) <span class="hlt">radars</span> peer into precipitation and provide information on the size, shape, phase (liquid / frozen), and concentration of the falling particles (termed hydrometeors). This information is valuable for improved rain rate estimates, and for providing data on the release and absorption of heat in the atmosphere from condensation and evaporation (phase changes). The heating profiles in the atmosphere influence global circulation, and are a vital component in studies of Earth s changing climate. However, to provide the most accurate interpretation of <span class="hlt">radar</span> data, the <span class="hlt">radar</span> must be properly calibrated and data must be quality controlled (cleaned) to remove non-precipitation artifacts; both of which are challenging tasks for today s weather <span class="hlt">radar</span>. The DP capability maximizes performance of these procedures using properties of the observed precipitation. In a notable paper published in 2005, scientists from the Cooperative Institute for Mesoscale Meteorological Studies (CIMMS) at the University of Oklahoma developed a method to calibrate <span class="hlt">radars</span> using statistically averaged DP measurements within light rain. An additional publication by one of the same scientists at the National Severe Storms Laboratory (NSSL) in Norman, Oklahoma introduced several techniques to perform quality control of <span class="hlt">radar</span> data using DP measurements. Following their lead, the Topical Rainfall Measuring Mission (TRMM) Satellite Validation Office at NASA s Goddard Space Flight Center has fine-tuned these methods for specific application to the weather <span class="hlt">radar</span> at Kwajalein Island in the Republic of the Marshall Islands, approximately 2100 miles</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025065','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025065"><span>Radiation pattern of a borehole <span class="hlt">radar</span> antenna</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ellefsen, K.J.; Wright, D.L.</p> <p>2002-01-01</p> <p>To understand better how a borehole antenna radiates <span class="hlt">radar</span> waves into a formation, this phenomenon is simulated numerically using the finite-difference, time-domain method. The simulations are of two different antenna models that include features like a driving <span class="hlt">point</span> fed by a coaxial cable, resistive loading of the antenna, and a water-filled borehole. For each model, traces are calculated in the far-field region, and then, from these traces, radiation patterns are calculated. The radiation patterns show that the amplitude of the <span class="hlt">radar</span> wave is strongly affected by its frequency, its propagation direction, and the resistive loading of the antenna.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA02705&hterms=Hydroelectric+power&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DHydroelectric%2Bpower','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA02705&hterms=Hydroelectric+power&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DHydroelectric%2Bpower"><span><span class="hlt">Radar</span> image with color as height, Bahia State, Brazil</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>This <span class="hlt">radar</span> image is the first to show the full 240-kilometer-wide (150 mile)swath collected by the Shuttle <span class="hlt">Radar</span> Topography Mission (SRTM). The area shown is in the state of Bahia in Brazil. The semi-circular mountains along the leftside of the image are the Serra Da Jacobin, which rise to 1100 meters (3600 feet) above sea level. The total relief shown is approximately 800 meters (2600 feet). The top part of the image is the Sertao, a semi-arid region, that is subject to severe droughts during El Nino events. A small portion of the San Francisco River, the longest river (1609 kilometers or 1000 miles) entirely within Brazil, cuts across the upper right corner of the image. This river is a major source of water for irrigation and hydroelectric power. Mapping such regions will allow scientists to better understand the relationships between flooding cycles, drought and human influences on ecosystems.<p/>This image combines two types of data from the Shuttle <span class="hlt">Radar</span> Topography Mission. The image brightness corresponds to the strength of the <span class="hlt">radar</span> signal reflected from the ground, while colors show the elevation as measured by SRTM. The three dark <span class="hlt">vertical</span> stripes show the boundaries where four segments of the swath are merged to form the full scanned swath. These will be removed in later processing. Colors range from green at the lowest elevations to reddish at the highest elevations.<p/>The Shuttle <span class="hlt">Radar</span> Topography Mission (SRTM), launched on February 11, 2000, uses the same <span class="hlt">radar</span> instrument that comprised the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) that flew twice on the Space Shuttle Endeavour in 1994. The mission is designed to collect three-dimensional measurements of the Earth's surface. To collect the 3-D data, engineers added a 60-meter-long (200-foot) mast, an additional C-band imaging antenna and improved tracking and navigation devices. The mission is a cooperative project between the National Aeronautics and Space</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFMSA51D..06E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFMSA51D..06E"><span>Effect of ray and speed perturbations on ionospheric tomography by over-the-horizon <span class="hlt">radar</span>: A new method, useful for SuperDarn <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Eisenbeis, J.; Roy, C.; Bland, E. C.; Occhipinti, G.</p> <p>2017-12-01</p> <p>Most recent methods in ionospheric tomography are based on the inversion of the total electron content measured by ground-based GPS receivers. As a consequence of the high frequency of the GPS signal and the absence of horizontal raypaths, the electron density structure is mainly reconstructed in the F2 region (300 km), where the ionosphere reaches the maximum of ionization, and is not sensitive to the lower ionospheric structure. We propose here a new tomographic method of the lower ionosphere (Roy et al., 2014), based on the full inversion of over-the-horizon (OTH) <span class="hlt">radar</span> data and applicable to SuperDarn data. The major advantage of our methodology is taking into account, numerically and jointly, the effect that the electron density perturbations induce not only in the speed of electromagnetic waves but also on the raypath geometry. This last <span class="hlt">point</span> is extremely critical for OTH/SuperDarn data inversions as the emitted signal propagates through the ionosphere between a fixed starting <span class="hlt">point</span> (the <span class="hlt">radar</span>) and an unknown end <span class="hlt">point</span> on the Earth surface where the signal is backscattered. We detail our ionospheric tomography method with the aid of benchmark tests in order to highlight the sensitivity of the <span class="hlt">radar</span> related to the explored observational parameters: frequencies, elevations, azimuths. Having proved the necessity to take into account both effects simultaneously, we apply our method to real backscattered data from Super Darn and OTH <span class="hlt">radar</span>. The preliminary solution obtained with the Hokkaido East SuperDARN with only two frequencies (10MHz and 11MHz), showed here, is stable and push us to deeply explore a more complete dataset that we will present at the AGU 2017. This is, in our knowledge, the first time that an ionospheric tomography has been estimated with SuperDarn backscattered data. Reference: Roy, C., G. Occhipinti, L. Boschi, J.-P. Moliné, and M. Wieczorek (2014), Effect of ray and speed perturbations on ionospheric tomography by over-the-horizon <span class="hlt">radar</span>: A</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20000053500','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20000053500"><span>The Shuttle <span class="hlt">Radar</span> Topography Mission: A Global DEM</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Farr, Tom G.; Kobrick, Mike</p> <p>2000-01-01</p> <p>Digital topographic data are critical for a variety of civilian, commercial, and military applications. Scientists use Digital Elevation Models (DEM) to map drainage patterns and ecosystems, and to monitor land surface changes over time. The mountain-building effects of tectonics and the climatic effects of erosion can also be modeled with DEW The data's military applications include mission planning and rehearsal, modeling and simulation. Commercial applications include determining locations for cellular phone towers, enhanced ground proximity warning systems for aircraft, and improved maps for backpackers. The Shuttle <span class="hlt">Radar</span> Topography Mission (SRTM) (Fig. 1), is a cooperative project between NASA and the National Imagery and Mapping Agency (NIMA) of the U.S. Department of Defense. The mission is designed to use a single-pass <span class="hlt">radar</span> interferometer to produce a digital elevation model of the Earth's land surface between about 60 degrees north and south latitude. The DEM will have 30 m pixel spacing and about 15 m <span class="hlt">vertical</span> errors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010100392','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010100392"><span>Measurement of Attenuation with Airborne and Ground-Based <span class="hlt">Radar</span> in Convective Storms Over Land and Its Microphysical Implications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tian, Lin; Heymsfield, G. M.; Srivastava, R. C.; Starr, D. OC. (Technical Monitor)</p> <p>2001-01-01</p> <p>Observations by the airborne X-band Doppler <span class="hlt">radar</span> (EDOP) and the NCAR S-band polarimetric (S-POL) <span class="hlt">radar</span> from two field experiments are used to evaluate the Surface ref'ercnce technique (SRT) for measuring the path integrated attenuation (PIA) and to study attenuation in deep convective storms. The EDOP, flying at an altitude of 20 km, uses a nadir beam and a forward <span class="hlt">pointing</span> beam. It is found that over land, the surface scattering cross-section is highly variable at nadir incidence but relatively stable at forward incidence. It is concluded that measurement by the forward beam provides a viable technique for measuring PIA using the SRT. <span class="hlt">Vertical</span> profiles of peak attenuation coefficient are derived in vxo deep convective storms by the dual-wavelength method. Using the measured Doppler velocity, the reflectivities at. the two wavelengths, the differential reflectivity and the estimated attenuation coefficients, it is shown that: supercooled drops and dry ice particles probably co-existed above the melting level in regions of updraft, that water-coated partially melted ice particles probably contributed to high attenuation below the melting level, and that the data are not readil explained in terms of a gamma function raindrop size distribution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990111506&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dbiomass%2Bforest','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990111506&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dbiomass%2Bforest"><span>Spaceborne Applications of P Band Imaging <span class="hlt">Radars</span> for Measuring Forest Biomass</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rignot, Eric J.; Zimmermann, Reiner; vanZyl, Jakob J.</p> <p>1995-01-01</p> <p>In three sites of boreal and temperate forests, P band HH, HV, and VV polarization data combined estimate total aboveground dry woody biomass within 12 to 27% of the values derived from allometric equations, depending on forest complexity. Biomass estimates derived from HV-polarization data only are 2 to 14% less accurate. When the <span class="hlt">radar</span> operates at circular polarization, the errors exceed 100% over flooded forests, wet or damaged trees and sparse open tall forests because double-bounce reflections of the <span class="hlt">radar</span> signals yield <span class="hlt">radar</span> signatures similar to that of tall and massive forests. Circular polarizations, which minimize the effect of Faraday rotation in spaceborne applications, are therefore of limited use for measuring forest biomass. In the tropical rain forest of Manu, in Peru, where forest biomass ranges from 4 kg/sq m in young forest succession up to 50 kg/sq m in old, undisturbed floodplain stands, the P band horizontal and <span class="hlt">vertical</span> polarization data combined separate biomass classes in good agreement with forest inventory estimates. The worldwide need for large scale, updated, biomass estimates, achieved with a uniformly applied method, justifies a more in-depth exploration of multi-polarization long wavelength imaging <span class="hlt">radar</span> applications for tropical forests inventories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20020034155','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20020034155"><span>TRMM Precipitation <span class="hlt">Radar</span> and Microwave Imager Observations of Convective and Stratiform Rain Over Land and Their Theoretical Implications</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Prabhakara, C.; Iacovazzi, R., Jr.; Yoo, J.-M.; Weinman, J. A.; Lau, William K. M. (Technical Monitor)</p> <p>2001-01-01</p> <p>Observations of brightness temperature, Tb made over land regions by the Tropical Rainfall Measuring Mission (TRMM) Microwave Imager (TMI) radiometer have been analyzed along with the nearly simultaneous measurements of the <span class="hlt">vertical</span> profiles of reflectivity factor, Z, made by the Precipitation <span class="hlt">Radar</span> (PR) onboard the TRMM satellite. This analysis is performed to explore the interrelationship between the TMI and PR data in areas that are covered predominantly by convective or stratiform rain. In particular, we have compared on a scale of 20 km, average <span class="hlt">vertical</span> profiles of Z with the averages of Tbs in the 19, 37 and 85 GHz channels. Generally, we find from these data that as Z increases, Tbs in the three channels decrease due to extinction. In order to explain physically the relationship between the Tb and Z observations, we have performed radiative transfer simulations utilizing <span class="hlt">vertical</span> profiles of hydrometeors applicable to convective and stratiform rain regions. These profiles are constructed taking guidance from the Z observations of PR and recent LDR and ZDR measurements made by land-based polarimetric <span class="hlt">radars</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01735&hterms=flying+fish&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dflying%2Bfish','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01735&hterms=flying+fish&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dflying%2Bfish"><span>Space <span class="hlt">Radar</span> Image of Manaus, Brazil</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>These two false-color images of the Manaus region of Brazil in South America were acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> on board the space shuttle Endeavour. The image at left was acquired on April 12, 1994, and the image at right was acquired on October 3, 1994. The area shown is approximately 8 kilometers by 40 kilometers (5 miles by 25 miles). The two large rivers in this image, the Rio Negro (at top) and the Rio Solimoes (at bottom), combine at Manaus (west of the image) to form the Amazon River. The image is centered at about 3 degrees south latitude and 61 degrees west longitude. North is toward the top left of the images. The false colors were created by displaying three L-band polarization channels: red areas correspond to high backscatter, horizontally transmitted and received, while green areas correspond to high backscatter, horizontally transmitted and <span class="hlt">vertically</span> received. Blue areas show low returns at <span class="hlt">vertical</span> transmit/receive polarization; hence the bright blue colors of the smooth river surfaces can be seen. Using this color scheme, green areas in the image are heavily forested, while blue areas are either cleared forest or open water. The yellow and red areas are flooded forest or floating meadows. The extent of the flooding is much greater in the April image than in the October image and appears to follow the 10-meter (33-foot) annual rise and fall of the Amazon River. The flooded forest is a vital habitat for fish, and floating meadows are an important source of atmospheric methane. These images demonstrate the capability of SIR-C/X-SAR to study important environmental changes that are impossible to see with optical sensors over regions such as the Amazon, where frequent cloud cover and dense forest canopies block monitoring of flooding. Field studies by boat, on foot and in low-flying aircraft by the University of California at Santa Barbara, in collaboration with Brazil's Instituto Nacional de Pesguisas</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20150019752','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20150019752"><span>Preliminary Findings of Inflight Icing Field Test to Support Icing Remote Sensing Technology Assessment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>King, Michael; Reehorst, Andrew; Serke, Dave</p> <p>2015-01-01</p> <p>NASA and the National Center for Atmospheric Research have developed an icing remote sensing technology that has demonstrated skill at detecting and classifying icing hazards in a <span class="hlt">vertical</span> column above an instrumented ground station. This technology has recently been extended to provide volumetric coverage surrounding an airport. Building on the existing <span class="hlt">vertical</span> <span class="hlt">pointing</span> system, the new method for providing volumetric coverage will utilize a <span class="hlt">vertical</span> <span class="hlt">pointing</span> cloud <span class="hlt">radar</span>, a multifrequency microwave radiometer with azimuth and elevation <span class="hlt">pointing</span>, and a NEXRAD <span class="hlt">radar</span>. The new terminal area icing remote sensing system processes the data streams from these instruments to derive temperature, liquid water content, and cloud droplet size for each examined <span class="hlt">point</span> in space. These data are then combined to ultimately provide icing hazard classification along defined approach paths into an airport.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01857&hterms=retreated&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dretreated','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01857&hterms=retreated&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dretreated"><span>Space <span class="hlt">Radar</span> Image of Cape Cod, Massachusetts</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This spaceborne <span class="hlt">radar</span> image shows the famous 'hook' of Cape Cod, Massachusetts. The Cape, which juts out into the Atlantic Ocean about 100 kilometers (62 miles) southeast of Boston, actually consists of sandy debris left behind by the great continental ice sheets when they last retreated from southern New England about 20,000 years ago. Today's landscape consists of sandy forests, fields of scrub oak and other bushes and grasses, salt marshes, freshwater ponds, as well as the famous beaches and sand dunes. In this image, thickly forested areas appear green, marshes are dark blue, ponds and sandy areas are black, and developed areas are mostly pink. The dark L-shape in the lower center is the airport runways in Hyannis, the Cape's largest town. The dark X-shape left of the center is Otis Air Force Base. The Cape Cod Canal, above and left of center, connects Buzzards Bay on the left with Cape Cod Bay on the right. The northern tip of the island of Martha's Vineyard is seen in the lower left. The tip of the Cape, in the upper right, includes the community of Provincetown, which appears pink, and the protected National Seashore areas of sand dunes that parallel the Atlantic coast east of Provincetown. Scientists are using <span class="hlt">radar</span> images like this one to study delicate coastal environments and the effects of human activities on the ecosystem and landscape. This image was acquired by Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) onboard the space shuttle Endeavour on April 15, 1994. The image is 81.7 kilometers by 43.1 kilometers (50.7 miles by 26.7 miles) and is centered at 41.8 degrees north latitude, 70.3 degrees west longitude. North is toward the upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations of the <span class="hlt">radar</span> as follows: red is L-band, horizontally transmitted and received; green is C-band, horizontally transmitted, <span class="hlt">vertically</span> received; and blue is C-band, horizontally transmitted and received. SIR</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JAsGe...6..256M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JAsGe...6..256M"><span>Joint application of Geoelectrical Resistivity and Ground Penetrating <span class="hlt">Radar</span> techniques for the study of hyper-saturated zones. Case study in Egypt</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mesbah, Hany S.; Morsy, Essam A.; Soliman, Mamdouh M.; Kabeel, Khamis</p> <p>2017-06-01</p> <p>This paper presents the results of the application of the Geoelectrical Resistivity Sounding (GRS) and Ground Penetrating <span class="hlt">Radar</span> (GPR) for outlining and investigating of surface springing out (flow) of groundwater to the base of an service building site, and determining the reason(s) for the zone of maximum degree of saturation; in addition to provide stratigraphic information for this site. The studied economic building is constructed lower than the ground surface by about 7 m. A <span class="hlt">Vertical</span> Electrical Sounding (VES) survey was performed at 12 <span class="hlt">points</span> around the studied building in order to investigate the <span class="hlt">vertical</span> and lateral extent of the subsurface sequence, three VES's were conducted at each side of the building at discrete distances. And a total of 9 GPR profiles with 100- and 200-MHz antennae were conducted, with the objective of evaluating the depth and the degree of saturation of the subsurface layers. The qualitative and quantitative interpretation of the acquired VES's showed easily the levels of saturations close to and around the studied building. From the interpretation of GPR profiles, it was possible to locate and determine the saturated layers. The <span class="hlt">radar</span> signals are penetrated and enabled the identification of the subsurface reflectors. The results of GPR and VES showed a good agreement and the integrated interpretations were supported by local geology. Finally, the new constructed geoelectrical resistivity cross-sections (in contoured-form), are easily clarifying the direction of groundwater flow toward the studied building.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70027587','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70027587"><span><span class="hlt">Radar</span> stage uncertainty</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Fulford, J.M.; Davies, W.J.</p> <p>2005-01-01</p> <p>The U.S. Geological Survey is investigating the performance of <span class="hlt">radars</span> used for stage (or water-level) measurement. This paper presents a comparison of estimated uncertainties and data for <span class="hlt">radar</span> water-level measurements with float, bubbler, and wire weight water-level measurements. The <span class="hlt">radar</span> sensor was also temperature-tested in a laboratory. The uncertainty estimates indicate that <span class="hlt">radar</span> measurements are more accurate than uncorrected pressure sensors at higher water stages, but are less accurate than pressure sensors at low stages. Field data at two sites indicate that <span class="hlt">radar</span> sensors may have a small negative bias. Comparison of field <span class="hlt">radar</span> measurements with wire weight measurements found that the <span class="hlt">radar</span> tends to measure slightly lower values as stage increases. Copyright ASCE 2005.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988STIN...8920373U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988STIN...8920373U"><span>A portable CW/FM-CW Doppler <span class="hlt">radar</span> for local investigation of severe storms</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Unruh, Wesley P.; Wolf, Michael A.; Bluestein, Howard B.</p> <p></p> <p>During the 1987 spring storm season we used a portable 1-W X-band CW Doppler <span class="hlt">radar</span> to probe a tornado, a funnel cloud, and a wall cloud in Oklahoma and Texas. This same device was used during the spring storm season in 1988 to probe a wall cloud in Texas. The <span class="hlt">radar</span> was battery powered and highly portable, and thus convenient to deploy from our chase vehicle. The device separated the receding and approaching Doppler velocities in real time and, while the <span class="hlt">radar</span> was being used, it allowed convenient stereo data recording for later spectral analysis and operator monitoring of the Doppler signals in stereo headphones. This aural monitoring, coupled with the ease with which an operator can be trained to recognize the nature of the signals heard, made the <span class="hlt">radar</span> very easy to operate reliably and significantly enhanced the quality of the data being recorded. At the end of the 1988 spring season, the <span class="hlt">radar</span> was modified to include FM-CW ranging and processing. These modifications were based on a unique combination of video recording and FM chirp generation, which incorporated a video camera and recorder as an integral part of the <span class="hlt">radar</span>. After modification, the <span class="hlt">radar</span> retains its convenient portability and the operational advantage of being able to listen to the Doppler signals directly. The original mechanical design was unaffected by these additions. During the summer of 1988, this modified device was used at the Langmuir Laboratory at Socorro, New Mexico in an attempt to measure <span class="hlt">vertical</span> convective flow in a thunderstorm.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20050234670','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20050234670"><span>The Tropical Convective Spectrum. Part 1; Archetypal <span class="hlt">Vertical</span> Structures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Boccippio, Dennis J.; Petersen, Walter A.; Cecil, Daniel J.</p> <p>2005-01-01</p> <p>A taxonomy of tropical convective and stratiform <span class="hlt">vertical</span> structures is constructed through cluster analysis of 3 yr of Tropical Rainfall Measuring Mission (TRMM) "warm-season" (surface temperature greater than 10 C) precipitation <span class="hlt">radar</span> (PR) <span class="hlt">vertical</span> profiles, their surface rainfall, and associated <span class="hlt">radar</span>-based classifiers (convective/ stratiform and brightband existence). Twenty-five archetypal profile types are identified, including nine convective types, eight stratiform types, two mixed types, and six anvil/fragment types (nonprecipitating anvils and sheared deep convective profiles). These profile types are then hierarchically clustered into 10 similar families, which can be further combined, providing an objective and physical reduction of the highly multivariate PR data space that retains <span class="hlt">vertical</span> structure information. The taxonomy allows for description of any storm or local convective spectrum by the profile types or families. The analysis provides a quasi-independent corroboration of the TRMM 2A23 convective/ stratiform classification. The global frequency of occurrence and contribution to rainfall for the profile types are presented, demonstrating primary rainfall contribution by midlevel glaciated convection (27%) and similar depth decaying/stratiform stages (28%-31%). Profiles of these types exhibit similar 37- and 85-GHz passive microwave brightness temperatures but differ greatly in their frequency of occurrence and mean rain rates, underscoring the importance to passive microwave rain retrieval of convective/stratiform discrimination by other means, such as polarization or texture techniques, or incorporation of lightning observations. Close correspondence is found between deep convective profile frequency and annualized lightning production, and pixel-level lightning occurrence likelihood directly tracks the estimated mean ice water path within profile types.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014acm..conf..441R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014acm..conf..441R"><span><span class="hlt">Radar</span>-derived asteroid shapes <span class="hlt">point</span> to a 'zone of stability' for topography slopes and surface erosion rates</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Richardson, J.; Graves, K.; Bowling, T.</p> <p>2014-07-01</p> <p>Previous studies of the combined effects of asteroid shape, spin, and self-gravity have focused primarily upon the failure limits for bodies with a variety of standard shapes, friction, and cohesion values [1,2,3]. In this study, we look in the opposite direction and utilize 22 asteroid shape-models derived from <span class="hlt">radar</span> inversion [4] and 7 small body shape-models derived from spacecraft observations [5] to investigate the region in shape/spin space [1,2] wherein self-gravity and rotation combine to produce a stable minimum state with respect to surface potential differences, dynamic topography, slope magnitudes, and erosion rates. This erosional minimum state is self-correcting, such that changes in the body's rotation rate, either up or down, will increase slope magnitudes across the body, thereby driving up erosion rates non-linearly until the body has once again reached a stable, minimized surface state [5]. We investigated this phenomenon in a systematic fashion using a series of synthesized, increasingly prolate spheroid shape models. Adjusting the rotation rate of each synthetic shape to minimize surface potential differences, dynamic topography, and slope magnitudes results in the magenta curve of the figure (right side), defining the zone of maximum surface stability (MSS). This MSS zone is invariant both with respect to body size (gravitational potential and rotational potential scale together with radius), and density when the scaled-spin of [2] is used. Within our sample of observationally derived small-body shape models, slow rotators (Group A: blue <span class="hlt">points</span>), that are not in the maximum surface stability (MSS) zone and where gravity dominates the slopes, will generally experience moderate erosion rates (left plot) and will tend to move up and to the right in shape/spin space as the body evolves (right plot). Fast rotators (Group C: red <span class="hlt">points</span>), that are not in the MSS zone and where spin dominates the slopes, will generally experience high erosion rates</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19840025810&hterms=temperature+classes&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtemperature%2Bclasses','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19840025810&hterms=temperature+classes&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dtemperature%2Bclasses"><span>Use of multi-frequency, multi-polarization, multi-angle airborne <span class="hlt">radars</span> for class discrimination in a southern temperature forest</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mehta, N. C.</p> <p>1984-01-01</p> <p>The utility of <span class="hlt">radar</span> scatterometers for discrimination and characterization of natural vegetation was investigated. Backscatter measurements were acquired with airborne multi-frequency, multi-polarization, multi-angle <span class="hlt">radar</span> scatterometers over a test site in a southern temperate forest. Separability between ground cover classes was studied using a two-class separability measure. Very good separability is achieved between most classes. Longer wavelength is useful in separating trees from non-tree classes, while shorter wavelength and cross polarization are helpful for discrimination among tree classes. Using the maximum likelihood classifier, 50% overall classification accuracy is achieved using a single, short-wavelength scatterometer channel. Addition of multiple incidence angles and another <span class="hlt">radar</span> band improves classification accuracy by 20% and 50%, respectively, over the single channel accuracy. Incorporation of a third <span class="hlt">radar</span> band seems redundant for vegetation classification. <span class="hlt">Vertical</span> transmit polarization is critically important for all classes.</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://ntrs.nasa.gov/search.jsp?R=19850008930&hterms=Treetops+algorithm&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DTreetops%2Balgorithm','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850008930&hterms=Treetops+algorithm&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DTreetops%2Balgorithm"><span>Information for space <span class="hlt">radar</span> designers: Required dynamic range vs resolution and antenna calibration using the Amazon rain forest</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Moore, R. K.; Frost, V. S.</p> <p>1985-01-01</p> <p>Calibration of the <span class="hlt">vertical</span> pattern of the antennas for the SEASAT scatterometer was accomplished using the nearly-uniform <span class="hlt">radar</span> return from the Amazon rain forest. A similar calibration will be attempted for the SIR-B antenna. Thick calibration is important to establish the radiometric calibration across the swath of the SIR-B, and the developed methodology will provide an important tool in the evaluation of future spaceborne imaging <span class="hlt">radars</span>. This calibration was made by the very-wide-beam SEASAT scatterometer antennas because at 14.65 GHz the scattering coefficient of the rain forest is almost independent of angle of incidence. It is expected that the variation in scattering coefficient for the rain forest across the relatively narrow <span class="hlt">vertical</span> beam of the SIR-B will be very small; even at L band the forest should be essentially impenetrable for <span class="hlt">radar</span> signals, the volume scatter from the treetops will predominate as at higher frequencies. The basic research elements include: (1) examination of SIR-B images over the rain forest to establish the variability of the scattering coefficient at finer resolutions than that of the SEASAT scatterometer; (2) analysis of the variability of SIR-B data detected prior to processing for either azimuth compression or; possibly, range compression so that averages over relatively large footprints can be used; (3) processing of data of the form of (2) using algorithms that can recover the <span class="hlt">vertical</span> pattern of the antenna.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009AnGeo..27...65D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009AnGeo..27...65D"><span>Aspect sensitive E- and F-region SPEAR-enhanced incoherent backscatter observed by the EISCAT Svalbard <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dhillon, R. S.; Robinson, T. R.; Yeoman, T. K.</p> <p>2009-01-01</p> <p>Previous studies of the aspect sensitivity of heater-enhanced incoherent <span class="hlt">radar</span> backscatter in the high-latitude ionosphere have demonstrated the directional dependence of incoherent scatter signatures corresponding to artificially excited electrostatic waves, together with consistent field-aligned signatures that may be related to the presence of artificial field-aligned irregularities. These earlier high-latitude results have provided motivation for repeating the investigation in the different geophysical conditions that obtain in the polar cap ionosphere. The Space Plasma Exploration by Active <span class="hlt">Radar</span> (SPEAR) facility is located within the polar cap and has provided observations of RF-enhanced ion and plasma line spectra recorded by the EISCAT Svalbard UHF incoherent scatter <span class="hlt">radar</span> system (ESR), which is collocated with SPEAR. In this paper, we present observations of aspect sensitive E- and F-region SPEAR-induced ion and plasma line enhancements that indicate excitation of both the purely growing mode and the parametric decay instability, together with sporadic E-layer results that may indicate the presence of cavitons. We note consistent enhancements from field-aligned, <span class="hlt">vertical</span> and also from 5° south of field-aligned. We attribute the prevalence of <span class="hlt">vertical</span> scatter to the importance of the Spitze region, and of that from field-aligned to possible wave/irregularity coupling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01300.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01300.html"><span>Space <span class="hlt">Radar</span> Image of New Orleans, Louisiana</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1998-04-14</p> <p>This image of the area surrounding the city of New Orleans, Louisiana in the southeastern United States demonstrates the ability of multi-frequency imaging <span class="hlt">radar</span> to distinguish different types of land cover. The dark area in the center is Lake Pontchartrain. The thin line running across the lake is a causeway connecting New Orleans to the city of Mandeville. Lake Borgne is the dark area in the lower right of the image. The Mississippi River appears as a dark, wavy line in the lower left. The white dots on the Mississippi are ships. The French Quarter is the brownish square near the left center of the image. Lakefront Airport, a field used mostly for general aviation, is the bright spot near the center, jutting out into Lake Pontchartrain. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span> C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) during orbit 39 of space shuttle Endeavour on October 2, 1994. The area is located at 30.10 degrees north latitude and 89.1 degrees west longitude. The area shown is approximately 100 kilometers (60 miles) by 50 kilometers (30 miles). The colors in this image were obtained using the following <span class="hlt">radar</span> channels: red represents the L-band (horizontally transmitted and received); green represents the C-band (horizontally transmitted and received); blue represents the L-band (<span class="hlt">vertically</span> transmitted and received). The green areas are primarily vegetation consisting of swamp land and swamp forest (bayou) growing on sandy soil, while the pink areas are associated with reflections from buildings in urban and suburban areas. Different tones and colors in the vegetation areas will be studied by scientists to see how effective imaging <span class="hlt">radar</span> data is in discriminating between different types of wetlands. Accurate maps of coastal wetland areas are important to ecologists studying wild fowl and the coastal environment. http://photojournal.jpl.nasa.gov/catalog/PIA01300</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1997JASTP..59.1035P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1997JASTP..59.1035P"><span>RAPIER: a new relocatable VHF coherent <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Popple, M.; Chapman, P. J.; Thomas, E. C.; Jones, T. B.</p> <p>1997-06-01</p> <p>VHF coherent <span class="hlt">radar</span> observations of the high latitude ionosphere have contributed significantly to our understanding of the complex processes which couple the ionosphere, magnetosphere and the solar wind. In addition, these observations have also improved our knowledge of the physics of the ionospheric plasma irregularities and their scattering properties. In this article the design of a new mobile coherent <span class="hlt">radar</span> system is described. The new system, RAPIER (Relocatable Auroral Polar Ionospheric <span class="hlt">Radar</span>), was initially collocated with the existing SABRE <span class="hlt">radar</span> and simultaneous operations undertaken to evaluate RAPIER's performance in its beam scanning mode. In this way the performance of the new system was quantitatively compared with that of a well established auroral <span class="hlt">radar</span> facility. The velocities measured by the new RAPIER system are well correlated with those observed by SABRE. The received backscatter powers observed by the two systems were, however, less well correlated, mainly caused by differences between their respective antenna elevation polar diagrams. As expected from system considerations, SABRE was found to be more sensitive than RAPIER at slant ranges corresponding to the maxima in the SABRE elevation polar diagrams. However, RAPIER's improved elevation polar diagram, superior instantaneous dynamic range and its ability to alter its receiver gain with <span class="hlt">pointing</span> direction ensured that it could accurately measure targets over a much greater spatial region than SABRE. This effect became more pronounced when regions of intense backscatter were monitored.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012EGUGA..1410417S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012EGUGA..1410417S"><span>Alpine <span class="hlt">radar</span> conversion for LAWR</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Savina, M.; Burlando, P.</p> <p>2012-04-01</p> <p> class of <span class="hlt">radars</span>, because it accounts for the large variability of hydrometeors reflectivity and <span class="hlt">vertical</span> hydrometeors positioning (echo-top), which is strongly influenced by the high location of the <span class="hlt">radar</span>. The ARCOM procedure is in addition embedded in a multistep quality control framework, which also includes the calibration on raingauge observations, and can be summarized as follow: 1) correction of both LAWR and raingauge observations for known errors (e.g. magnetron decay and heated-related water loss) 2) evaluation of the local Pearson's correlation coefficient (PCC) as estimator of the linear correlation between raingauge and LAWR observations (logarithmic receiver); 3) computation of the local ACF in the form of the local linear regression coefficient between raingauge and LAWR observations; 4) calibration of the ARCOM, i.e. definition of the parametrization able to reproduce the spatial variability of ACF as function of the local sP, being the PCCs used as weight in the calibration procedure. The resulting calibrated ARCOM finally allows, in any ungauged mountain spot, to convert LAWR observations into precipitation rate. The temporal and the spatial transferability of the ARCOM are evaluated via split-sample and a take-one-out cross validation. The results revealed good spatial transferability and a seasonal bias within 7%, thus opening new opportunities for local range distributed measurements of precipitation in mountain regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011SPIE.8044E..05M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011SPIE.8044E..05M"><span>Fusion of <span class="hlt">radar</span> and satellite target measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Moy, Gabriel; Blaty, Donald; Farber, Morton; Nealy, Carlton</p> <p>2011-06-01</p> <p>A potentially high payoff for the ballistic missile defense system (BMDS) is the ability to fuse the information gathered by various sensor systems. In particular, it may be valuable in the future to fuse measurements made using ground based <span class="hlt">radars</span> with passive measurements obtained from satellite-based EO/IR sensors. This task can be challenging in a multitarget environment in view of the widely differing resolution between active ground-based <span class="hlt">radar</span> and an observation made by a sensor at long range from a satellite platform. Additionally, each sensor system could have a residual <span class="hlt">pointing</span> bias which has not been calibrated out. The problem is further compounded by the possibility that an EO/IR sensor may not see exactly the same set of targets as a microwave <span class="hlt">radar</span>. In order to better understand the problems involved in performing the fusion of metric information from EO/IR satellite measurements with active microwave <span class="hlt">radar</span> measurements, we have undertaken a study of this data fusion issue and of the associated data processing techniques. To carry out this analysis, we have made use of high fidelity simulations to model the <span class="hlt">radar</span> observations from a missile target and the observations of the same simulated target, as gathered by a constellation of satellites. In the paper, we discuss the improvements seen in our tests when fusing the state vectors, along with the improvements in sensor bias estimation. The limitations in performance due to the differing phenomenology between IR and microwave <span class="hlt">radar</span> are discussed as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990111590','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990111590"><span>Ocean Turbulence. Paper 3; Two-<span class="hlt">Point</span> Closure Model Momentum, Heat and Salt <span class="hlt">Vertical</span> Diffusivities in the Presence of Shear</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Canuto, V. M.; Dubovikov, M. S.; Howard, A.; Cheng, Y.</p> <p>1999-01-01</p> <p>In papers 1 and 2 we have presented the results of the most updated 1-<span class="hlt">point</span> closure model for the turbulent <span class="hlt">vertical</span> diffusivities of momentum, heat and salt, K(sub m,h,s). In this paper, we derive the analytic expressions for K(sub m,h,s) using a new 2-<span class="hlt">point</span> closure model that has recently been developed and successfully tested against some approx. 80 turbulence statistics for different flows. The new model has no free parameters. The expressions for K(sub m, h. s) are analytical functions of two stability parameters: the Turner number R(sub rho) (salinity gradient/temperature gradient) and the Richardson number R(sub i) (temperature gradient/shear). The turbulent kinetic energy K and its rate of dissipation may be taken local or non-local (K-epsilon model). Contrary to all previous models that to describe turbulent mixing below the mixed layer (ML) have adopted three adjustable "background diffusivities" for momentum. heat and salt, we propose a model that avoids such adjustable diffusivities. We assume that below the ML, K(sub m,h,s) have the same functional dependence on R(sub i) and R(sub rho) derived from the turbulence model. However, in order to compute R(sub i) below the ML, we use data of <span class="hlt">vertical</span> shear due to wave-breaking measured by Gargett et al. (1981). The procedure frees the model from adjustable background diffusivities and indeed we use the same model throughout the entire <span class="hlt">vertical</span> extent of the ocean. Using the new K(sub m,h, s), we run an O-GCM and present a variety of results that we compare with Levitus and the KPP model. Since the traditional 1-<span class="hlt">point</span> (used in papers 1 and 2) and the new 2-<span class="hlt">point</span> closure models used here represent different modeling philosophies and procedures, testing them in an O-GCM is indispensable. The basic motivation is to show that the new 2-<span class="hlt">point</span> closure model gives results that are overall superior to the 1-<span class="hlt">point</span> closure in spite of the fact that the latter rely on several adjustable parameters while the new 2-<span class="hlt">point</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A31A2153Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A31A2153Q"><span>Enhance the accuracy of <span class="hlt">radar</span> snowfall estimation with Multi new Z-S relationships in MRMS system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qi, Y.</p> <p>2017-12-01</p> <p>Snow may have negative affects on roadways and human lives, but the result of the melted snow/ice is good for farm, humans, and animals. For example, in the Southwest and West mountainous area of United States, water shortage is a very big concern. However, snowfall in the winter can provide humans, animals and crops an almost unlimited water supply. So, using <span class="hlt">radar</span> to accurately estimate the snowfall is very important for human life and economic development in the water lacking area. The current study plans to analyze the characteristics of the horizontal and <span class="hlt">vertical</span> variations of dry/wet snow using dual polarimetric <span class="hlt">radar</span> observations, relative humidity and in situ snow water equivalent observations from the National Weather Service All Weather Prediction Accumulation Gauges (AWPAG) across the CONUS, and establish the relationships between the reflectivity (Z) and ground snow water equivalent (S). The new Z-S relationships will be evaluated with independent CoCoRaHS (Community Collaborative Rain, Hail & Snow Network) gauge observations and eventually implemented in the Multi-<span class="hlt">Radar</span> Multi-Sensor system for improved quantitative precipitation estimation for snow. This study will analyze the characteristics of the horizontal and <span class="hlt">vertical</span> variations of dry/wet snow using dual polarimetric <span class="hlt">radar</span> observations, relative humidity and in situ snow water equivalent observations from the National Weather Service All Weather Prediction Accumulation Gauges (AWPAG) across the CONUS, and establish the relationships between the reflectivity (Z) and ground snow water equivalent (S). The new Z-S relationships will be used to reduce the error of snowfall estimation in Multi <span class="hlt">Radar</span> and Multi Sensors (MRMS) system, and tested in MRMS system and evaluated with the COCORaHS observations. Finally, it will be ingested in MRMS sytem, and running in NWS/NCAR operationally</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890062568&hterms=water+effects&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dwater%2Beffects','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890062568&hterms=water+effects&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dwater%2Beffects"><span>A study of rain effects on <span class="hlt">radar</span> scattering from water waves</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bliven, Larry F.; Giovanangeli, Jean-Paul; Norcross, George</p> <p>1988-01-01</p> <p>Results are presented from a laboratory investigation of microwave power return due to rain-generated short waves on a wind wave surface. The wind wave tank, sensor, and data processing methods used in the study are described. The study focuses on the response of a 36-GHz <span class="hlt">radar</span> system, orientated 30 deg from nadir and <span class="hlt">pointing</span> upwind, to surface waves generated by various combinations of rain and wind. The results show stronger <span class="hlt">radar</span> signal levels due to short surface waves generated by rain impacting the wind wave surface, supporting the results of Moore et al. (1979) for a 14-GHz <span class="hlt">radar</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.4268L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.4268L"><span>Cassini <span class="hlt">RADAR</span> at Titan : Results in 2014/2015</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lorenz, Ralph D.</p> <p>2015-04-01</p> <p>Since the last EGU meeting, two Cassini flybys of Titan will have featured significant <span class="hlt">RADAR</span> observations, illuminating our understanding of this enigmatic, complex world and its hydrocarbon seas in particular. T104, which executed in August 2014, featured a nadir-<span class="hlt">pointed</span> altimetry swath over the northern part of Kraken Mare, Titan's largest sea. The echo characteristics showed that the sea surface was generally flat (to within a few mm), although a couple of areas appear to show some evidence of roughness. Intriguingly, altimetry processing which yielded (Mastrogiuseppe et al., GRL, 2014) the detection of a prominent bottom echo 160m beneath the surface of Ligeia Mare on T91 failed to yield a similar echo over most of Kraken on T104, suggesting either that Kraken is very deep (perhaps consistent with rather steep shoreline topography) or that the liquid in Kraken is more <span class="hlt">radar</span>-absorbing than that in Ligeia, or both. The absorbing-liquid scenario may be consistent with a hydrological model for Titan's seas (Lorenz, GRL, 2014) wherein the most northerly seas receive more 'fresh' methane input, flushing ethane and other lower-volatility (and more <span class="hlt">radar</span>-absorbing) solutes south into Kraken. T108, the last northern seas <span class="hlt">radar</span> observation until T126 at the very end of the Cassini tour in 2017, is planned to execute on 11th January 2015, and preliminary results will be presented at the EGU meeting. This flyby features altimetry over part of Punga Mare, which will provide surface roughness information and possible bathymetry, permitting comparison of nadir-<span class="hlt">pointed</span> data over all of Titan's three seas (Ligeia on T91; Kraken Mare on T104). The flyby also includes SAR observation of the so-called Ligeia 'Magic Island', the best-observed of several areas of varying <span class="hlt">radar</span> brightness on Titan's seas. This brightness may be due to sediments suspended by currents, or by roughening of the surface either by local wind stress ('catspaw') or non-local stress (wind-driven currents</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/460790','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/460790"><span>Ionospheric effects on synthetic aperture <span class="hlt">radar</span> at VHF</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>Fitzgerald, T.J.</p> <p>1997-02-01</p> <p>Synthetic aperture <span class="hlt">radars</span> (SAR) operated from airplanes have been used at VHF because of their enhanced foliage and ground penetration compared to <span class="hlt">radars</span> operated at UHF. A satellite-borne VHF SAR would have considerable utility but in order to operate with high resolution it would have to use both a large relative bandwidth and a large aperture. The presence of the ionosphere in the propagation path of the <span class="hlt">radar</span> will cause a deterioration of the imaging because of dispersion over the bandwidth and group path changes in the imaged area over the collection aperture. In this paper we present calculations ofmore » the effects of a deterministic ionosphere on SAR imaging for a <span class="hlt">radar</span> operated with a 100 MHz bandwidth centered at 250 MHz and over an angular aperture of 23{degrees}. The ionosphere induces a <span class="hlt">point</span> spread function with an approximate half-width of 150 m in the slant-range direction and of 25 m in the cross-range direction compared to the nominal resolution of 1.5 m in both directions.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AMT....10..221B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AMT....10..221B"><span>Observing relationships between lightning and cloud profiles by means of a satellite-borne cloud <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Buiat, Martina; Porcù, Federico; Dietrich, Stefano</p> <p>2017-01-01</p> <p>Cloud electrification and related lightning activity in thunderstorms have their origin in the charge separation and resulting distribution of charged iced particles within the cloud. So far, the ice distribution within convective clouds has been investigated mainly by means of ground-based meteorological <span class="hlt">radars</span>. In this paper we show how the products from Cloud Profiling <span class="hlt">Radar</span> (CPR) on board CloudSat, a polar satellite of NASA's Earth System Science Pathfinder (ESSP), can be used to obtain information from space on the <span class="hlt">vertical</span> distribution of ice particles and ice content and relate them to the lightning activity. The analysis has been carried out, focusing on 12 convective events over Italy that crossed CloudSat overpasses during significant lightning activity. The CPR products considered here are the <span class="hlt">vertical</span> profiles of cloud ice water content (IWC) and the effective radius (ER) of ice particles, which are compared with the number of strokes as measured by a ground lightning network (LINET). Results show a strong correlation between the number of strokes and the <span class="hlt">vertical</span> distribution of ice particles as depicted by the 94 GHz CPR products: in particular, cloud upper and middle levels, high IWC content and relatively high ER seem to be favourable contributory causes for CG (cloud to ground) stroke occurrence.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014acm..conf..185G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014acm..conf..185G"><span>Physical properties of meteoroids based on middle and upper atmosphere <span class="hlt">radar</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>Gritsevich, M.; Kero, J.; Virtanen, J.; Szasz, C.; Nakamura, T.; Peltoniemi, J.; Koschny, D.</p> <p>2014-07-01</p> <p>We present a novel approach to reliably interpret the meteor head-echo scattering measurements detected by the 46.5 MHz MU <span class="hlt">radar</span> system near Shigaraki, Japan. A meteor head echo is caused by radio waves scattered from the dense region of plasma surrounding and co-moving with a meteoroid during atmospheric flight. The signal Doppler shift and/or range rate of the target can therefore be used to determine meteoroid velocity. The data reduction steps include determining the exact trajectory of the meteoroids entering the observation volume of the antenna beam and calculating meteoroid mass and velocity as a function of time. The model is built using physically-based parametrization. The considered observation volume is narrow, elongated in the <span class="hlt">vertical</span> direction, and its area of greatest sensitivity covers a circular area of about 10 km diameter at an altitude of 100 km above the <span class="hlt">radar</span>. Over 100,000 meteor head echoes have been detected over past years of observations. Most of the events are faint with no alternative to be detected visually or with intensified video (ICCD) cameras. In this study we are focusing on objects which have entered the atmosphere with almost <span class="hlt">vertical</span> trajectories, to ensure the observed segment of the trajectory to be as complete as possible, without loss of its beginning or end part due to beam-pattern-related loss of signal power. The analysis output parameters are range, altitude, radial velocity, meteoroid velocity, instantaneous target position, <span class="hlt">Radar</span> Cross Section (RCS), meteor radiant, meteoroid ballistic and ablation coefficients, mass loss parameter and meteoroid mass, with possibility to derive other parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014cosp...40E1077G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014cosp...40E1077G"><span>Physical Properties of Meteoroids based on Middle and Upper Atmosphere <span class="hlt">Radar</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>Gritsevich, Maria; Nakamura, Takuji; Kero, Johan; Szasz, Csilla; Virtanen, Jenni; Peltoniemi, Jouni; Koschny, Detlef</p> <p></p> <p>We present a novel approach to reliably interpret the meteor head echo scattering measurements detected by the 46.5 MHz MU <span class="hlt">radar</span> system near Shigaraki, Japan. A meteor head echo is caused by radio waves scattered from the dense region of plasma surrounding and co-moving with a meteoroid during atmospheric flight. The signal Doppler shift and/or range rate of the target can therefore be used to determine meteoroid velocity. The data reduction steps include determining the exact trajectory of the meteoroids entering the observation volume of the antenna beam and calculating meteoroid mass and velocity as a function of time. The model is built using physically based parameterization. The considered observation volume is narrow, elongated in the <span class="hlt">vertical</span> direction, and its area of greatest sensitivity covers a circular area of about 10 km diameter at an altitude of 100 km above the <span class="hlt">radar</span>. Over 100000 meteor head echoes have been detected over past years of observations. Most of the events are faint with no alternative to be detected visually or with intensified video (ICCD) cameras. In this study we are focusing on objects which have entered the atmosphere with almost <span class="hlt">vertical</span> trajectories, to ensure the observed segment of the trajectory to be as complete as possible, without loss of its beginning or end part due to beam-pattern related loss of signal power. The analysis output parameters are range, altitude, radial velocity, meteoroid velocity, instantaneous target position, <span class="hlt">Radar</span> Cross Section (RCS), meteor radiant, meteoroid ballistic and ablation coefficients, mass loss parameter and meteoroid mass, with possibility to derive other parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340241p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/mi0425.photos.340241p/"><span>3. VIEW NORTHWEST, height finder <span class="hlt">radar</span> towers, and <span class="hlt">radar</span> tower ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>3. VIEW NORTHWEST, height finder <span class="hlt">radar</span> towers, and <span class="hlt">radar</span> tower (unknown function) - Fort Custer Military Reservation, P-67 <span class="hlt">Radar</span> Station, .25 mile north of Dickman Road, east of Clark Road, Battle Creek, Calhoun County, MI</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405970-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-oliktok-point-radar-installation-alaska-volume-includes-appendices','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405970-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-oliktok-point-radar-installation-alaska-volume-includes-appendices"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron, Elmendorf AFB, Alaska. Remedial investigation and feasibility study: Oliktok <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Volume 1. (Includes appendices a - b)</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>NONE</p> <p>1996-04-15</p> <p>This report presents the findings of Remedial Investigations and Feasibility Studies at sites located at the Oliktok <span class="hlt">Point</span> <span class="hlt">radar</span> installation in northern Alaska. The sites were characterized based on sampling and analyses conducted during Remedial Investigation activities performed during August and September 1993.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A21P..01K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A21P..01K"><span>New Cloud and Precipitation Research Avenues Enabled by low-cost Phased-array <span class="hlt">Radar</span> Technology</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kollias, P.; Oue, M.; Fridlind, A. M.; Matsui, T.; McLaughlin, D. J.</p> <p>2017-12-01</p> <p>For over half a century, <span class="hlt">radars</span> operating in a wide range of frequencies have been the primary source of observational insights of clouds and precipitation microphysics and dynamics and contributed to numerous significant advancements in the field of cloud and precipitation physics. The development of multi-wavelength and polarization diversity techniques has further strengthened the quality of microphysical and dynamical retrievals from <span class="hlt">radars</span> and has assisted in overcoming some of the limitations imposed by the physics of scattering. Atmospheric <span class="hlt">radars</span> have historically employed a mechanically-scanning dish antenna and their ability to <span class="hlt">point</span> to, survey, and revisit specific <span class="hlt">points</span> or regions in the atmosphere is limited by mechanical inertia. Electronically scanned, or phased-array, <span class="hlt">radars</span> capable of high-speed, inertialess beam steering, have been available for several decades, but the cost of this technology has limited its use to military applications. During the last 10 years, lower power and lower-cost versions of electronically scanning <span class="hlt">radars</span> have been developed, and this presents an attractive and affordable new tool for the atmospheric sciences. The operational and research communities are currently exploring phased array advantages in signal processing (i.e. beam multiplexing, improved clutter rejection, cross beam wind estimation, adaptive sensing) and science applications (i.e. tornadic storm morphology studies). Here, we will present some areas of atmospheric research where inertia-less <span class="hlt">radars</span> with ability to provide rapid volume imaging offers the potential to advance cloud and precipitation research. We will discuss the added value of single phased-array <span class="hlt">radars</span> as well as networks of these <span class="hlt">radars</span> for several problems including: multi-Doppler wind retrieval techniques, cloud lifetime studies and aerosol-convection interactions. The performance of current (dish) and future (e-scan) <span class="hlt">radar</span> systems for these atmospheric studies will be evaluated using</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AtmRe.203..286G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AtmRe.203..286G"><span><span class="hlt">Radar</span>-derived quantitative precipitation estimation in complex terrain over the eastern Tibetan Plateau</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gou, Yabin; Ma, Yingzhao; Chen, Haonan; Wen, Yixin</p> <p>2018-05-01</p> <p>Quantitative precipitation estimation (QPE) is one of the important applications of weather <span class="hlt">radars</span>. However, in complex terrain such as Tibetan Plateau, it is a challenging task to obtain an optimal Z-R relation due to the complex spatial and temporal variability in precipitation microphysics. This paper develops two <span class="hlt">radar</span> QPE schemes respectively based on Reflectivity Threshold (RT) and Storm Cell Identification and Tracking (SCIT) algorithms using observations from 11 Doppler weather <span class="hlt">radars</span> and 3264 rain gauges over the Eastern Tibetan Plateau (ETP). These two QPE methodologies are evaluated extensively using four precipitation events that are characterized by different meteorological features. Precipitation characteristics of independent storm cells associated with these four events, as well as the storm-scale differences, are investigated using short-term <span class="hlt">vertical</span> profile of reflectivity (VPR) clusters. Evaluation results show that the SCIT-based rainfall approach performs better than the simple RT-based method for all precipitation events in terms of score comparison using validation gauge measurements as references. It is also found that the SCIT-based approach can effectively mitigate the local error of <span class="hlt">radar</span> QPE and represent the precipitation spatiotemporal variability better than the RT-based scheme.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870047847&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dbiomass%2Bforest','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870047847&hterms=biomass+forest&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dbiomass%2Bforest"><span>Forest biomass, canopy structure, and species composition relationships with multipolarization L-band synthetic aperture <span class="hlt">radar</span> data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sader, Steven A.</p> <p>1987-01-01</p> <p>The effect of forest biomass, canopy structure, and species composition on L-band synthetic aperature <span class="hlt">radar</span> data at 44 southern Mississippi bottomland hardwood and pine-hardwood forest sites was investigated. Cross-polarization mean digital values for pine forests were significantly correlated with green weight biomass and stand structure. Multiple linear regression with five forest structure variables provided a better integrated measure of canopy roughness and produced highly significant correlation coefficients for hardwood forests using HV/VV ratio only. Differences in biomass levels and canopy structure, including branching patterns and <span class="hlt">vertical</span> canopy stratification, were important sources of volume scatter affecting multipolarization <span class="hlt">radar</span> data. Standardized correction techniques and calibration of aircraft data, in addition to development of canopy models, are recommended for future investigations of forest biomass and structure using synthetic aperture <span class="hlt">radar</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405975-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-decision-document-further-response-action-planned-oliktok-point-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405975-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-decision-document-further-response-action-planned-oliktok-point-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron, Elmendorf AFB, Alaska. Decision document for no further response action planned Oliktok <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-06-03</p> <p>This Decision Document discusses the selection of no further action as the recommended action for four sites located at the Oliktok <span class="hlt">Point</span> <span class="hlt">radar</span> installation. The United States Air Force (Air Force) completed a Remedial Investigation/Feasibility Study and a Risk Assessment for the eight sites located at the Oliktok <span class="hlt">Point</span> installation (U.S. Air Force 1996a,b). Based on the findings of these activities, four sites are recommended for no further action.</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://www.dtic.mil/docs/citations/ADA361115','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA361115"><span>Oceanography - High Frequency <span class="hlt">Radar</span> and Ocean Thin Layers, Volume 10, No. 2</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1999-03-11</p> <p>near Monterey Bay. A major advantage of HF <span class="hlt">radar</span> measurements is their ability to describe these processes in two dimensions. Complicating this...Seabreeze cycle in the winds is a broad- band process centered near the diurnal period. Harmonic analyses of coastal surface currents at periods...accurate representations of a near -surface process related to wind forcing, whereas the semidiurnal oscillations have longer <span class="hlt">vertical</span> scales and are</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225466p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/ca3450.photos.225466p/"><span>Topographic and location map of Bonita <span class="hlt">Point</span> Coast Guard and ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>Topographic and location map of Bonita <span class="hlt">Point</span> Coast Guard and lighthouse station, June 1940, this drawing shows the Bonita Ridge access road retaining wall and general conditions at Fort Barry and Bonita Ridge (upper left) before the construction of Signal Corps <span class="hlt">Radar</span> (S.C.R.) 296 Station 5 - Fort Barry, Signal Corps <span class="hlt">Radar</span> 296, Station 5, Transmitter Building Foundation, <span class="hlt">Point</span> Bonita, Marin Headlands, Sausalito, Marin County, CA</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20060028449','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20060028449"><span>Debris Flux Comparisons From The Goldstone <span class="hlt">Radar</span>, Haystack <span class="hlt">Radar</span>, and Hax <span class="hlt">Radar</span> Prior, During, and After the Last Solar Maximum</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Stokely, C. L.; Stansbery, E. G.; Goldstein, R. M.</p> <p>2006-01-01</p> <p>The continual monitoring of low Earth orbit (LEO) debris environment using highly sensitive <span class="hlt">radars</span> is essential for an accurate characterization of these dynamic populations. Debris populations are continually evolving since there are new debris sources, previously unrecognized debris sources, and debris loss mechanisms that are dependent on the dynamic space environment. Such <span class="hlt">radar</span> data are used to supplement, update, and validate existing orbital debris models. NASA has been utilizing <span class="hlt">radar</span> observations of the debris environment for over a decade from three complementary <span class="hlt">radars</span>: the NASA JPL Goldstone <span class="hlt">radar</span>, the MIT Lincoln Laboratory (MIT/LL) Long Range Imaging <span class="hlt">Radar</span> (known as the Haystack <span class="hlt">radar</span>), and the MIT/LL Haystack Auxiliary <span class="hlt">radar</span> (HAX). All of these systems are highly sensitive <span class="hlt">radars</span> that operate in a fixed staring mode to statistically sample orbital debris in the LEO environment. Each of these <span class="hlt">radars</span> is ideally suited to measure debris within a specific size region. The Goldstone <span class="hlt">radar</span> generally observes objects with sizes from 2 mm to 1 cm. The Haystack <span class="hlt">radar</span> generally measures from 5 mm to several meters. The HAX <span class="hlt">radar</span> generally measures from 2 cm to several meters. These overlapping size regions allow a continuous measurement of cumulative debris flux versus diameter from 2 mm to several meters for a given altitude window. This is demonstrated for all three <span class="hlt">radars</span> by comparing the debris flux versus diameter over 200 km altitude windows for 3 nonconsecutive years from 1998 through 2003. These years correspond to periods before, during, and after the peak of the last solar cycle. Comparing the year to year flux from Haystack for each of these altitude regions indicate statistically significant changes in subsets of the debris populations. Potential causes of these changes are discussed. These analysis results include error bars that represent statistical sampling errors, and are detailed in this paper.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19820003099','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19820003099"><span>Planetary <span class="hlt">radar</span> studies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Thompson, T. W.; Cutts, J. A.</p> <p>1981-01-01</p> <p>A catalog of lunar and <span class="hlt">radar</span> anomalies was generated to provide a base for comparison with Venusian <span class="hlt">radar</span> signatures. The relationships between lunar <span class="hlt">radar</span> anomalies and regolith processes were investigated, and a consortium was formed to compare lunar and Venusian <span class="hlt">radar</span> images of craters. Time was scheduled at the Arecibo Observatory to use the 430 MHz <span class="hlt">radar</span> to obtain high resolution <span class="hlt">radar</span> maps of six areas of the lunar suface. Data from 1978 observations of Mare Serenitas and Plato are being analyzed on a PDP 11/70 computer to construct the computer program library necessary for the eventual reduction of the May 1981 and subsequent data acquisitions. Papers accepted for publication are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EGUGA..19.2262L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EGUGA..19.2262L"><span>Orographic precipitation and <span class="hlt">vertical</span> velocity characteristics from drop size and fall velocity spectra observed by disdrometers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Dong-In; Kim, Dong-Kyun; Kim, Ji-Hyeon; Kang, Yunhee; Kim, Hyeonjoon</p> <p>2017-04-01</p> <p>During a summer monsoon season each year, severe weather phenomena caused by front, mesoscale convective systems, or typhoons often occur in the southern Korean Peninsula where is mostly comprised of complex high mountains. These areas play an important role in controlling formation, amount, and distribution of rainfall. As precipitation systems move over the mountains, they can develop rapidly and produce localized heavy rainfall. Thus observational analysis in the mountainous areas is required for studying terrain effects on the rapid rainfall development and its microphysics. We performed intensive field observations using two s-band operational weather <span class="hlt">radars</span> around Mt. Jiri (1950 m ASL) during summertime on June and July in 2015-2016. Observation data of DSD (Drop Size Distribution) from Parsivel disdrometer and (w component) <span class="hlt">vertical</span> velocity data from ultrasonic anemometers were analyzed for Typhoon Chanhom on 12 July 2015 and the heavy rain event on 1 July 2016. During the heavy rain event, a dual-Doppler <span class="hlt">radar</span> analysis using Jindo <span class="hlt">radar</span> and Gunsan <span class="hlt">radar</span> was also conducted to examine 3-D wind fields and <span class="hlt">vertical</span> structure of reflectivity in these areas. For examining up-/downdrafts in the windward or leeward side of Mt. Jiri, we developed a new scheme technique to estimate <span class="hlt">vertical</span> velocities (w) from drop size and fall velocity spectra of Parsivel disdrometers at different stations. Their comparison with the w values observed by the 3D anemometer showed quite good agreement each other. The Z histogram with regard to the estimated w was similar to that with regard to R, indicating that Parsivel-estimated w is quite reasonable for classifying strong and weak rain, corresponding to updraft and downdraft, respectively. Mostly, positive w values (upward) were estimated in heavy rainfall at the windward side (D1 and D2). Negative w values (downward) were dominant even during large rainfall at the leeward side (D4). For D1 and D2, the upward w percentages were</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19820003100','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19820003100"><span>Planetary <span class="hlt">radar</span> studies. [<span class="hlt">radar</span> mapping of the Moon and <span class="hlt">radar</span> signatures of lunar and Venus craters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Thompson, T. W.; Cutts, J. A.</p> <p>1981-01-01</p> <p>Progress made in studying the evolution of Venusian craters and the evolution of infrared and <span class="hlt">radar</span> signatures of lunar crater interiors is reported. Comparison of <span class="hlt">radar</span> images of craters on Venus and the Moon present evidence for a steady state Venus crater population. Successful observations at the Arecibo Observatory yielded good data on five nights when data for a mix of inner and limb areas were acquired. Lunar craters with <span class="hlt">radar</span> bright ejects are discussed. An overview of infrared <span class="hlt">radar</span> crater catalogs in the data base is included.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720016521','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720016521"><span>Apollo experience report: Lunar module landing <span class="hlt">radar</span> and rendezvous <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rozas, P.; Cunningham, A. R.</p> <p>1972-01-01</p> <p>A developmental history of the Apollo lunar module landing and rendezvous <span class="hlt">radar</span> subsystems is presented. The Apollo <span class="hlt">radar</span> subsystems are discussed from initial concept planning to flight configuration testing. The major <span class="hlt">radar</span> subsystem accomplishments and problems are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199422p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199422p/"><span>30. Perimeter acquisition <span class="hlt">radar</span> building room #318, showing <span class="hlt">radar</span> control. ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>30. Perimeter acquisition <span class="hlt">radar</span> building room #318, showing <span class="hlt">radar</span> control. Console and line printers - Stanley R. Mickelsen Safeguard Complex, Perimeter Acquisition <span class="hlt">Radar</span> Building, Limited Access Area, between Limited Access Patrol Road & Service Road A, Nekoma, Cavalier County, ND</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140010279','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140010279"><span>A User Guide for Smoothing Air Traffic <span class="hlt">Radar</span> Data</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bach, Ralph E.; Paielli, Russell A.</p> <p>2014-01-01</p> <p>Matlab software was written to provide smoothing of <span class="hlt">radar</span> tracking data to simulate ADS-B (Automatic Dependent Surveillance-Broadcast) data in order to test a tactical conflict probe. The probe, called TSAFE (Tactical Separation-Assured Flight Environment), is designed to handle air-traffic conflicts left undetected or unresolved when loss-of-separation is predicted to occur within approximately two minutes. The data stream that is down-linked from an aircraft equipped with an ADS-B system would include accurate GPS-derived position and velocity information at sample rates of 1 Hz. Nation-wide ADS-B equipage (mandated by 2020) should improve surveillance accuracy and TSAFE performance. Currently, position data are provided by Center <span class="hlt">radar</span> (nominal 12-sec samples) and Terminal <span class="hlt">radar</span> (nominal 4.8-sec samples). Aircraft ground speed and ground track are estimated using real-time filtering, causing lags up to 60 sec, compromising performance of a tactical resolution tool. Offline smoothing of <span class="hlt">radar</span> data reduces wild-<span class="hlt">point</span> errors, provides a sample rate as high as 1 Hz, and yields more accurate and lag-free estimates of ground speed, ground track, and climb rate. Until full ADS-B implementation is available, smoothed <span class="hlt">radar</span> data should provide reasonable track estimates for testing TSAFE in an ADS-B-like environment. An example illustrates the smoothing of <span class="hlt">radar</span> data and shows a comparison of smoothed-<span class="hlt">radar</span> and ADS-B tracking. This document is intended to serve as a guide for using the smoothing software.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20160008414&hterms=TYPES+RADAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DTYPES%2BOF%2BRADAR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20160008414&hterms=TYPES+RADAR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DTYPES%2BOF%2BRADAR"><span>Planetary <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Neish, Catherine D.; Carter, Lynn M.</p> <p>2015-01-01</p> <p>This chapter describes the principles of planetary <span class="hlt">radar</span>, and the primary scientific discoveries that have been made using this technique. The chapter starts by describing the different types of <span class="hlt">radar</span> systems and how they are used to acquire images and accurate topography of planetary surfaces and probe their subsurface structure. It then explains how these products can be used to understand the properties of the target being investigated. Several examples of discoveries made with planetary <span class="hlt">radar</span> are then summarized, covering solar system objects from Mercury to Saturn. Finally, opportunities for future discoveries in planetary <span class="hlt">radar</span> are outlined and discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20160012017','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20160012017"><span>Effect of <span class="hlt">Vertical</span> Rate Error on Recovery from Loss of Well Clear Between UAS and Non-Cooperative Intruders</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cone, Andrew; Thipphavong, David; Lee, Seung Man; Santiago, Confesor</p> <p>2016-01-01</p> <p>When an Unmanned Aircraft System (UAS) encounters an intruder and is unable to maintain required temporal and spatial separation between the two vehicles, it is referred to as a loss of well-clear. In this state, the UAS must make its best attempt to regain separation while maximizing the minimum separation between itself and the intruder. When encountering a non-cooperative intruder (an aircraft operating under visual flight rules without ADS-B or an active transponder) the UAS must rely on the <span class="hlt">radar</span> system to provide the intruders location, velocity, and heading information. As many UAS have limited climb and descent performance, <span class="hlt">vertical</span> position andor <span class="hlt">vertical</span> rate errors make it difficult to determine whether an intruder will pass above or below them. To account for that, there is a proposal by RTCA Special Committee 228 to prohibit guidance systems from providing <span class="hlt">vertical</span> guidance to regain well-clear to UAS in an encounter with a non-cooperative intruder unless their <span class="hlt">radar</span> system has <span class="hlt">vertical</span> position error below 175 feet (95) and <span class="hlt">vertical</span> velocity errors below 200 fpm (95). Two sets of fast-time parametric studies was conducted, each with 54000 pairwise encounters between a UAS and non-cooperative intruder to determine the suitability of offering <span class="hlt">vertical</span> guidance to regain well clear to a UAS in the presence of <span class="hlt">radar</span> sensor noise. The UAS was not allowed to maneuver until it received well-clear recovery guidance. The maximum severity of the loss of well-clear was logged and used as the primary indicator of the separation achieved by the UAS. One set of 54000 encounters allowed the UAS to maneuver either <span class="hlt">vertically</span> or horizontally, while the second permitted horizontal maneuvers, only. Comparing the two data sets allowed researchers to see the effect of allowing <span class="hlt">vertical</span> guidance to a UAS for a particular encounter and <span class="hlt">vertical</span> rate error. Study results show there is a small reduction in the average severity of a loss of well-clear when <span class="hlt">vertical</span> maneuvers</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://water.usgs.gov/ogw/bgas/publications/SIR2005-5087/toc.html','USGSPUBS'); return false;" href="http://water.usgs.gov/ogw/bgas/publications/SIR2005-5087/toc.html"><span>Analysis of Borehole-<span class="hlt">Radar</span> Reflection Data from Machiasport, Maine, December 2003</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Johnson, Carole D.; Joesten, Peter K.</p> <p>2005-01-01</p> <p>-reflection data. There are several steeply dipping reflectors with orientations similar to the fracture patterns observed with borehole imaging techniques and in outcrops. The <span class="hlt">radar</span>-reflection data showed that the vitrophyre in borehole MW09 was more highly fractured than the underlying gabbroic unit. The velocities of <span class="hlt">radar</span> waves in the bedrock surrounding the boreholes were determined using single-hole <span class="hlt">vertical</span> <span class="hlt">radar</span> profiling. Velocities of 114 and 125 meters per microsecond were used to determine the distance to reflectors, the radial depth of penetration, and the dip of reflectors. The bimodal volcanic units appear to be ideal for <span class="hlt">radar</span>-wave propagation. For the <span class="hlt">radar</span> surveys collected at this site, <span class="hlt">radar</span> reflections were detected up to 40 m into the rock from the borehole. These results indicate that boreholes could conservatively be spaced about 15-20 m apart for hole-to-hole <span class="hlt">radar</span> methods to be effective for imaging between the boreholes and monitoring remediation. Integrated analysis of drilling and borehole-geophysical logs indicates the vitrophyric formation is more fractured than the more mafic gabbroic units in these boreholes. There does not, however, appear to be a quantifiable difference in the <span class="hlt">radar</span>-wave penetration in these two rock units.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970003539','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970003539"><span>Observations of <span class="hlt">Radar</span> Backscatter at Ku and C Bands in the Presence of Large Waves during the Surface Wave Dynamics Experiment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nghiem, S. V.; Li, Fuk K.; Lou, Shu-Hsiang; Neumann, Gregory; McIntosh, Robert E.; Carson, Steven C.; Carswell, James R.; Walsh, Edward J.; Donelan, Mark A.; Drennan, William M.</p> <p>1995-01-01</p> <p>Ocean <span class="hlt">radar</span> backscatter in the presence of large waves is investigated using data acquired with the Jet Propulsion Laboratory NUSCAT <span class="hlt">radar</span> at Ku band for horizontal and <span class="hlt">vertical</span> polarizations and the University of Massachusetts CSCAT <span class="hlt">radar</span> at C band for <span class="hlt">vertical</span> polarization during the Surface Wave Dynamics Experiment. Off-nadir backscatter data of ocean surfaces were obtained in the presence of large waves with significant wave height up to 5.6 m. In moderate-wind cases, effects of large waves are not detectable within the measurement uncertainty and no noticeable correlation between backscatter coefficients and wave height is found. Under high-wave light-wind conditions, backscatter is enhanced significantly at large incidence angles with a weaker effect at small incidence angles. Backscatter coefficients in the wind speed range under consideration are compared with SASS-2 (Ku band), CMOD3-H1 (C band), and Plant's model results which confirm the experimental observations. Variations of the friction velocity, which can give rise to the observed backscatter behaviors in the presence of large waves, are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA09075&hterms=mars+climate+orbiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3D%2527mars%2Bclimate%2Borbiter%2527','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA09075&hterms=mars+climate+orbiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3D%2527mars%2Bclimate%2Borbiter%2527"><span>Interpreting <span class="hlt">Radar</span> View near Mars' South Pole, Orbit 1360</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2006-01-01</p> <p><p/> A radargram from the Shallow Subsurface <span class="hlt">Radar</span> instrument (SHARAD) on NASA's Mars Reconnaissance Orbiter is shown in the upper-right panel and reveals detailed structure in the polar layered deposits of the south pole of Mars. <p/> The sounding <span class="hlt">radar</span> collected the data presented here during orbit 1360 of the mission, on Nov. 10, 2006. <p/> The horizontal scale in the radargram is distance along the ground track. It can be referenced to the ground track map shown in the lower right. The <span class="hlt">radar</span> traversed from about 74 degrees to 85 degrees south latitude, or about 650 kilometers (400 miles). The ground track map shows elevation measured by the Mars Orbiter Laser Altimeter on NASA's Mars Global Surveyor orbiter. Green indicates low elevation; reddish-white indicates higher elevation. The traverse proceeds up onto a plateau formed by the layers. <p/> The <span class="hlt">vertical</span> scale on the radargram is time delay of the <span class="hlt">radar</span> signals reflected back to Mars Reconnaissance Orbiter from the surface and subsurface. For reference, using an assumed velocity of the <span class="hlt">radar</span> waves in the subsurface, time is converted to depth below the surface at one place: about 800 meters (2,600 feet) to one of the strongest subsurface reflectors. This reflector marks the base of the polar layered deposits. The color scale varies from black for weak reflections to white for strong reflections. <p/> The middle panel shows mapping of the major subsurface reflectors, some of which can be traced for a distance of 100 kilometers (60 miles) or more. The layering manifests the recent climate history of Mars, recorded by the deposition and removal of ice and dust. <p/> The Shallow Subsurface <span class="hlt">Radar</span> was provided by the Italian Space Agency (ASI). Its operations are led by the University of Rome and its data are analyzed by a joint U.S.-Italian science team. NASA's Jet Propulsion Laboratory, a division of the California Institute of Technology, Pasadena, manages the Mars Reconnaissance Orbiter for the NASA Science</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhRvD..90f4052P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhRvD..90f4052P"><span><span class="hlt">Radar</span> orthogonality and <span class="hlt">radar</span> length in Finsler and metric spacetime geometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pfeifer, Christian</p> <p>2014-09-01</p> <p>The <span class="hlt">radar</span> experiment connects the geometry of spacetime with an observers measurement of spatial length. We investigate the <span class="hlt">radar</span> experiment on Finsler spacetimes which leads to a general definition of <span class="hlt">radar</span> orthogonality and <span class="hlt">radar</span> length. The directions <span class="hlt">radar</span> orthogonal to an observer form the spatial equal time surface an observer experiences and the <span class="hlt">radar</span> length is the physical length the observer associates to spatial objects. We demonstrate these concepts on a forth order polynomial Finsler spacetime geometry which may emerge from area metric or premetric linear electrodynamics or in quantum gravity phenomenology. In an explicit generalization of Minkowski spacetime geometry we derive the deviation from the Euclidean spatial length measure in an observers rest frame explicitly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20140010287','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20140010287"><span>Wind Retrieval Algorithms for the IWRAP and HIWRAP Airborne Doppler <span class="hlt">Radars</span> with Applications to Hurricanes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Guimond, Stephen Richard; Tian, Lin; Heymsfield, Gerald M.; Frasier, Stephen J.</p> <p>2013-01-01</p> <p>Algorithms for the retrieval of atmospheric winds in precipitating systems from downward-<span class="hlt">pointing</span>, conically-scanning airborne Doppler <span class="hlt">radars</span> are presented. The focus in the paper is on two <span class="hlt">radars</span>: the Imaging Wind and Rain Airborne Profiler(IWRAP) and the High-altitude IWRAP (HIWRAP). The IWRAP is a dual-frequency (Cand Ku band), multi-beam (incidence angles of 30 50) system that flies on the NOAAWP-3D aircraft at altitudes of 2-4 km. The HIWRAP is a dual-frequency (Ku and Kaband), dual-beam (incidence angles of 30 and 40) system that flies on the NASA Global Hawk aircraft at altitudes of 18-20 km. Retrievals of the three Cartesian wind components over the entire <span class="hlt">radar</span> sampling volume are described, which can be determined using either a traditional least squares or variational solution procedure. The random errors in the retrievals are evaluated using both an error propagation analysis and a numerical simulation of a hurricane. These analyses show that the <span class="hlt">vertical</span> and along-track wind errors have strong across-track dependence with values of 0.25 m s-1 at nadir to 2.0 m s-1 and 1.0 m s-1 at the swath edges, respectively. The across-track wind errors also have across-track structure and are on average, 3.0 3.5 m s-1 or 10 of the hurricane wind speed. For typical rotated figure four flight patterns through hurricanes, the zonal and meridional wind speed errors are 2 3 m s-1.Examples of measured data retrievals from IWRAP during an eyewall replacement cycle in Hurricane Isabel (2003) and from HIWRAP during the development of Tropical Storm Matthew (2010) are shown.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPC.1967b0036L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPC.1967b0036L"><span>Research on the range side lobe suppression method for modulated stepped frequency <span class="hlt">radar</span> signals</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Yinkai; Shan, Tao; Feng, Yuan</p> <p>2018-05-01</p> <p>The magnitude of time-domain range sidelobe of modulated stepped frequency <span class="hlt">radar</span> affects the imaging quality of inverse synthetic aperture <span class="hlt">radar</span> (ISAR). In this paper, the cause of high sidelobe in modulated stepped frequency <span class="hlt">radar</span> imaging is analyzed first in real environment. Then, the chaos particle swarm optimization (CPSO) is used to select the amplitude and phase compensation factors according to the minimum sidelobe criterion. Finally, the compensated one-dimensional range images are obtained. Experimental results show that the amplitude-phase compensation method based on CPSO algorithm can effectively reduce the sidelobe peak value of one-dimensional range images, which outperforms the common sidelobe suppression methods and avoids the coverage of weak scattering <span class="hlt">points</span> by strong scattering <span class="hlt">points</span> due to the high sidelobes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1336080-fingerprints-riming-event-cloud-radar-doppler-spectra-observations-modeling','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1336080-fingerprints-riming-event-cloud-radar-doppler-spectra-observations-modeling"><span>Fingerprints of a riming event on cloud <span class="hlt">radar</span> Doppler spectra: observations and modeling</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Kalesse, Heike; Szyrmer, Wanda; Kneifel, Stefan; ...</p> <p>2016-03-09</p> <p>In this paper, <span class="hlt">Radar</span> Doppler spectra measurements are exploited to study a riming event when precipitating ice from a seeder cloud sediment through a supercooled liquid water (SLW) layer. The focus is on the "golden sample" case study for this type of analysis based on observations collected during the deployment of the Atmospheric Radiation Measurement Program's (ARM) mobile facility AMF2 at Hyytiälä, Finland, during the Biogenic Aerosols – Effects on Clouds and Climate (BAECC) field campaign. The presented analysis of the height evolution of the <span class="hlt">radar</span> Doppler spectra is a state-of-the-art retrieval with profiling cloud <span class="hlt">radars</span> in SLW layers beyondmore » the traditional use of spectral moments. Dynamical effects are considered by following the particle population evolution along slanted tracks that are caused by horizontal advection of the cloud under wind shear conditions. In the SLW layer, the identified liquid peak is used as an air motion tracer to correct the Doppler spectra for <span class="hlt">vertical</span> air motion and the ice peak is used to study the <span class="hlt">radar</span> profiles of rimed particles. A 1-D steady-state bin microphysical model is constrained using the SLW and air motion profiles and cloud top <span class="hlt">radar</span> observations. The observed <span class="hlt">radar</span> moment profiles of the rimed snow can be simulated reasonably well by the model, but not without making several assumptions about the ice particle concentration and the relative role of deposition and aggregation. In conclusion, this suggests that in situ observations of key ice properties are needed to complement the profiling <span class="hlt">radar</span> observations before process-oriented studies can effectively evaluate ice microphysical parameterizations.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830026005','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830026005"><span>Balloon-borne pressure sensor performance evaluation utilizing tracking <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Norcross, G. A.; Brooks, R. L.</p> <p>1983-01-01</p> <p>The pressure sensors on balloon-borne sondes relate the sonde measurements to height above the Earth's surface through the hypsometric equation. It is crucial that sondes used to explore the <span class="hlt">vertical</span> structure of the atmosphere do not contribute significant height errors to their measurements of atmospheric constituent concentrations and properties. A series of radiosonde flights was conducted. In most cases, each flight consisted of two sondes attached to a single balloon and each flight was tracked by a highly accurate C-band <span class="hlt">radar</span>. For the first 19 radiosonde flights, the standard aneroid cell baroswitch assembly used was the pressure sensor. The last 26 radiosondes were equipped with a premium grade aneroid cell baroswitch assembly sensor and with a hypsometer. It is shown that both aneroid cell baroswitch sensors become increasingly inaccurate with altitude. The hypsometer <span class="hlt">radar</span> differences are not strongly dependent upon altitude and it is found that the standard deviation of the differences at 35 km is 0.179 km.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFMSA51A4080C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFMSA51A4080C"><span>Comparison of <span class="hlt">Vertical</span> Drifts of ISR and Magnetometer Data Measurements at the Magnetic Equator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Condor P, P. J.</p> <p>2014-12-01</p> <p>We compare <span class="hlt">vertical</span> drifts measured with the Jicamarca incoherent scatter <span class="hlt">radar</span> (ISR) and drifts estimated from magnetometer data applying a Neural Network data processing technique. For the application of the Neural Network (NN) method, we use the magnitude of the horizontal (H) component of the magnetic field measured with magnetometers at Jicamarca and Piura (Peru). The data was collected between the years 2002 and 2013. In training the NN we use the difference between the magnitudes of the horizontal components (dH) measured at JRO (placed at the magnetic equator) and Piura (displaced 5° away). Additional parameters used are F10.7 and Ap indexes. The estimates obtained with the NN procedure are very good. We have an RMS error of 3.7 m/s using dH as an input of the NN while the error is 3.9 m/s when we use the component H of JRO as an input. The results are validated using the set of <span class="hlt">vertical</span> drifts observations collected with the Jicamarca incoherent scatter <span class="hlt">radar</span>. The estimated drifts can be accessed using the following website: http://jro.igp.gob.pe/driftnn. In the poster, we show the comparison of <span class="hlt">vertical</span> drifts from 2002 to 2013 where we discuss the agreement between magnetometer and ISR data.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01826&hterms=animal+rights&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Danimal%2Brights','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01826&hterms=animal+rights&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Danimal%2Brights"><span>Space <span class="hlt">Radar</span> Image of Santa Cruz Island, California</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1994-01-01</p> <p>This space <span class="hlt">radar</span> image shows the rugged topography of Santa Cruz Island, part of the Channel Islands National Park in the Pacific Ocean off the coast of Santa Barbara and Ventura, Calif. Santa Cruz, the largest island of the national park, is host to hundreds of species of plants, animals and birds, at least eight of which are known nowhere else in the world. The island is bisected by the Santa Cruz Island fault, which appears as a prominent line running from the upper left to the lower right in this image. The fault is part of the Transverse Range fault system, which extends eastward from this area across Los Angeles to near Palm Springs, Calif. Color variations in this image are related to the different types of vegetation and soils at the surface. For example, grass-covered coastal lowlands appear gold, while chaparral and other scrub areas appear pink and blue. The image is 35 kilometers by 32 kilometers (22 miles by 20 miles) and is centered at 33.8 degrees north latitude, 119.6 degrees west longitude. North is toward upper right. The colors are assigned to different <span class="hlt">radar</span> frequencies and polarizations as follows: red is L-band, horizontally transmitted and received; green is C-band, horizontally transmitted and received; and blue is C-band, horizontally transmitted and <span class="hlt">vertically</span> received. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) on October 10, 1994, onboard the space shuttle Endeavour. SIR-C/X-SAR, a joint mission of the German, Italian and United States space agencies, is part of NASA's Mission to Planet Earth program.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009EGUGA..11..999A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009EGUGA..11..999A"><span>A comparison of selected <span class="hlt">vertical</span> wind measurement techniques on basis of the EUCAARI IMPACT observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arabas, S.; Baehr, C.; Boquet, M.; Dufournet, Y.; Pawlowska, H.; Siebert, H.; Unal, C.</p> <p>2009-04-01</p> <p>The poster presents a comparison of selected methods for determination of the <span class="hlt">vertical</span> wind in the boundary layer used during the EUCAARI IMPACT campaign that took place in May 2008 in The Netherlands. The campaign covered a monthlong intensified ground-based and airborne measurements in the vicinity of the CESAR observatory in Cabauw. Ground-based <span class="hlt">vertical</span> wind remote sensing was carried out using the Leosphere WindCube WLS70 IR Doppler lidar, Vaisala LAP3000 <span class="hlt">radar</span> wind-profiler and the TUDelft TARA S-band <span class="hlt">radar</span>. In-situ airborne measurements were performed using an ultrasonic anemometer (on the ACTOS helicopter underhung platform) and a 5-hole pressure probe (on the SAFIRE ATR-42 airplane radome). Several in-situ anemometers were deployed on the 200-meter high tower of the CESAR observatory. A summary of the characteristics and principles of the considered techniques is presented. A comparison of the results obtained from different platforms depicts the capabilities of each technique and highlights the time, space and velocity resolutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19910017302','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19910017302"><span>The instrumental principles of MST <span class="hlt">radars</span> and incoherent scatter <span class="hlt">radars</span> and the configuration of <span class="hlt">radar</span> system hardware</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Roettger, Juergen</p> <p>1989-01-01</p> <p>The principle of pulse modulation used in the case of coherent scatter <span class="hlt">radars</span> (MST <span class="hlt">radars</span>) is discussed. Coherent detection and the corresponding system configuration is delineated. Antenna requirements and design are outlined and the phase-coherent transmitter/receiver system is described. Transmit/receive duplexers, transmitters, receivers, and quadrature detectors are explained. The <span class="hlt">radar</span> controller, integrator, decoder and correlator design as well as the data transfer and the control and monitoring by the host computer are delineated. Typical operation parameters of some well-known <span class="hlt">radars</span> are summarized.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/405961-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-bullen-point-radar-installation-alaska-final-report','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/405961-united-states-air-force-air-support-group-civil-engineering-squadron-elmendorf-afb-alaska-remedial-investigation-feasibility-study-bullen-point-radar-installation-alaska-final-report"><span>United States Air Force 611th Air Support Group/Civil Engineering Squadron Elmendorf AFB, Alaska. Remedial investigation and feasibility study. Bullen <span class="hlt">Point</span> <span class="hlt">Radar</span> Installation, Alaska. Final report</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>Karmi, S.</p> <p>1996-03-18</p> <p>The United States Air Force (Air Force) has prepared this Remedial investigation/Feasibility Study (RI/FS) report as part of the Installation Restoration Program (IRP) to present results of RI/FS activities at five sites at the Bullen <span class="hlt">Point</span> <span class="hlt">radar</span> installation. The IRP provides for investigating, quantifying, and remediating environmental contamination from past waste management activities at Air Force installations throughout the United States.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA01725&hterms=silk+road&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dsilk%2Broad','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA01725&hterms=silk+road&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dsilk%2Broad"><span>Space <span class="hlt">Radar</span> Image of Niya ruins, Taklamakan desert</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1999-01-01</p> <p>This <span class="hlt">radar</span> image is of an area thought to contain the ruins of the ancient settlement of Niya. It is located in the southwestern corner of the Taklamakan Desert in China's Sinjiang Province. This oasis was part of the famous Silk Road, an ancient trade route from one of China's earliest capitols, Xian, to the West. The image shows a white linear feature trending diagonally from the upper left to the lower right. Scientists believe this newly discovered feature is a man-made canal which presumably diverted river waters toward the settlement of Niya for irrigation purposes. The image was acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) aboard the space shuttle Endeavour on its 106th orbit on April 16, 1994, and is centered at 37.78 degrees north latitude and 82.41 degrees east longitude. The false-color <span class="hlt">radar</span> image was created by displaying the C-band (horizontally transmitted and received) return in red, the L-band (horizontally transmitted and received) return in green, and the L-band (horizontally transmitted and <span class="hlt">vertically</span> received) return in blue. Areas in mottled white and purple are low-lying floodplains of the Niya River. Dark green and black areas between river courses are higher ridges or dunes confining the water flow. Spaceborne Imaging <span class="hlt">Radar</span>-C and X-band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) is part of NASA's Mission to Planet Earth. The <span class="hlt">radars</span> illuminate Earth with microwaves, allowing detailed observations at any time, regardless of weather or sunlight conditions. SIR-C/X-SAR uses three microwave wavelengths: the L-band (24 cm), C-band (6 cm) and X-band (3 cm). The multi-frequency data will be used by the international scientific community to better understand the global environment and how it is changing. The SIR-C/X-SAR data, complemented by aircraft and ground studies, will give scientists clearer insights into those environmental changes which are caused by nature and those changes which are induced by</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 observations made with an airborne synthetic aperture imaging <span class="hlt">radar</span></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 <span class="hlt">radar</span> flown aboard the NASA CV-990 has observed 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 <span class="hlt">radar</span> images are compared to photographic imagery of similar waves taken from Landsat 1. Both the <span class="hlt">radar</span> and Landsat images reveal variations in reflectivity across each wave in a packet that range from low to high to normal. The variations <span class="hlt">point</span> 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('https://www.osti.gov/biblio/5259787-radar-echo-from-flat-conducting-plate-near-far','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/5259787-radar-echo-from-flat-conducting-plate-near-far"><span><span class="hlt">Radar</span> echo from a flat conducting plate - near and far</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>Williams, C.S.</p> <p>1982-01-01</p> <p>Over certain types of terrain, a <span class="hlt">radar</span> fuze (or altimeter), by virtue of the horizontal component of its velocity, is likely to pass over various flat objects of limited size. The echo from such objects could have a duration less than that of one Doppler cycle, where the Doppler frequency is due to the <span class="hlt">vertical</span> component of the velocity. If the terrain is principally made up of such objects, their echoes are in most cases entirely uncorrelated with each other. Hence, the total echo after mixing at the <span class="hlt">radar</span> with the delayed transmitted wave would have a noise-like spectrum notmore » at all confined to the Doppler-frequency band where the desired echo signal is expected. This would seriously degrade the performance of a <span class="hlt">radar</span> that utilizes correlation. This work shows that the echo from a square flat plate will be of duration greater than the time it takes to pass over the plate if the height h above it satisfies h > a/sup 2//lambda where a is the plate-edge dimension and lambda is the <span class="hlt">radar</span> wavelength. The results presented here can be used to determine the spatial region wherein the echo exists, and the magnitude and phase of the echo from such a plate. I infer from these results that the case where the signal has a noise-like spectrum is not impossible but it is unlikely for the applications with which I am familiar.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19940015940&hterms=oceans+behavior&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Doceans%2Bbehavior','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19940015940&hterms=oceans+behavior&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Doceans%2Bbehavior"><span>Ku-band ocean <span class="hlt">radar</span> backscatter observations during SWADE</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nghiem, S. V.; Li, F. K.; Lou, S. H.; Neumann, G.</p> <p>1993-01-01</p> <p>We present results obtained by an airborne Ku-band scatterometer during the Surface Wave Dynamics Experiment (SWADE). The specific objective of this study is to improve our understanding of the relationship between ocean <span class="hlt">radar</span> backscatter and near surface winds. The airborne scatterometer, NUSCAT, was flown on the NASA Ames C-130 over an instrumented oceanic area near 37 deg N and 74 deg W. A total of 10 flights from 27 Feb. to 9 Mar. 1991 were conducted. <span class="hlt">Radar</span> backscatter at incidence angles of 0 to 60 deg were obtained. For each incidence angle, the NUSCAT antenna was azimuthally scanned in multiple complete circles to measure the azimuthal backscatter modulations. Both horizontal and <span class="hlt">vertical</span> polarization backscatter measurements were made. In some of the flights, the cross-polarization backscatter was measured as well. Internal calibrations were carried out throughout each of the flights. Preliminary results indicate that the <span class="hlt">radar</span> was stable to +/-0.3 dB for each flight. In this paper, we present studies of the backscatter measurements over several crossings of the Gulf Stream. In these crossings, large air-sea temperature differences were encountered and substantial changes in the <span class="hlt">radar</span> cross section were observed. We summarize the observations and compare them to the changes of several wind variables across the Gulf Stream boundary. In one of the flights, the apparent wind near the cold side of the Gulf Stream was very low (less than 3 m/s). The behavior of the <span class="hlt">radar</span> cross sections at such low wind speeds and a comparison with models are presented. A case study of the effects of swell on the absolute cross section and the azimuthal modulation pattern is presented. Significant wave heights larger than m were observed during SWADE. The experimentally observed effects of the swell on the <span class="hlt">radar</span> backscatter are discussed. The effects are used to assess the uncertainties in wind retrieval due to underlying waves. A summary of azimuthal modulation from our ten</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/957002','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/957002"><span>Removing interfering clutter associated with <span class="hlt">radar</span> pulses that an airborne <span class="hlt">radar</span> receives from a <span class="hlt">radar</span> transponder</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Ormesher, Richard C.; Axline, Robert M.</p> <p>2008-12-02</p> <p>Interfering clutter in <span class="hlt">radar</span> pulses received by an airborne <span class="hlt">radar</span> system from a <span class="hlt">radar</span> transponder can be suppressed by developing a representation of the incoming echo-voltage time-series that permits the clutter associated with predetermined parts of the time-series to be estimated. These estimates can be used to estimate and suppress the clutter associated with other parts of the time-series.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA286279','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA286279"><span>Recommendation on Transition from Primary/Secondary <span class="hlt">Radar</span> to Secondary- Only <span class="hlt">Radar</span> Capability</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1994-10-01</p> <p><span class="hlt">Radar</span> Beacon Performance Monitor RCIU Remote Control Interface Unit RCL Remote Communications Link R E&D Research, Engineering and Development RML <span class="hlt">Radar</span>...rate. 3.1.2.5 Maintenance The current LRRs have limited remote maintenance monitoring (RMM) capabilities via the Remote Control Interface Unit ( RCIU ...1, -2 and FPS-20 <span class="hlt">radars</span> required an upgrade of some of the <span class="hlt">radar</span> subsystems, namely the RCIU to respond as an RMS and the CD to interface with <span class="hlt">radar</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011SPIE.8060E..09L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011SPIE.8060E..09L"><span>A hardware-in-the-loop simulation program for ground-based <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lam, Eric P.; Black, Dennis W.; Ebisu, Jason S.; Magallon, Julianna</p> <p>2011-06-01</p> <p>A <span class="hlt">radar</span> system created using an embedded computer system needs testing. The way to test an embedded computer system is different from the debugging approaches used on desktop computers. One way to test a <span class="hlt">radar</span> system is to feed it artificial inputs and analyze the outputs of the <span class="hlt">radar</span>. More often, not all of the building blocks of the <span class="hlt">radar</span> system are available to test. This will require the engineer to test parts of the <span class="hlt">radar</span> system using a "black box" approach. A common way to test software code on a desktop simulation is to use breakpoints so that is pauses after each cycle through its calculations. The outputs are compared against the values that are expected. This requires the engineer to use valid test scenarios. We will present a hardware-in-the-loop simulator that allows the embedded system to think it is operating with real-world inputs and outputs. From the embedded system's <span class="hlt">point</span> of view, it is operating in real-time. The hardware in the loop simulation is based on our Desktop PC Simulation (PCS) testbed. In the past, PCS was used for ground-based <span class="hlt">radars</span>. This embedded simulation, called Embedded PCS, allows a rapid simulated evaluation of ground-based <span class="hlt">radar</span> performance in a laboratory environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080037982','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080037982"><span>Reconfigurable L-Band <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rincon, Rafael F.</p> <p>2008-01-01</p> <p>The reconfigurable L-Band <span class="hlt">radar</span> is an ongoing development at NASA/GSFC that exploits the capability inherently in phased array <span class="hlt">radar</span> systems with a state-of-the-art data acquisition and real-time processor in order to enable multi-mode measurement techniques in a single <span class="hlt">radar</span> architecture. The development leverages on the L-Band Imaging Scatterometer, a <span class="hlt">radar</span> system designed for the development and testing of new <span class="hlt">radar</span> techniques; and the custom-built DBSAR processor, a highly reconfigurable, high speed data acquisition and processing system. The <span class="hlt">radar</span> modes currently implemented include scatterometer, synthetic aperture <span class="hlt">radar</span>, and altimetry; and plans to add new modes such as radiometry and bi-static GNSS signals are being formulated. This development is aimed at enhancing the <span class="hlt">radar</span> remote sensing capabilities for airborne and spaceborne applications in support of Earth Science and planetary exploration This paper describes the design of the <span class="hlt">radar</span> and processor systems, explains the operational modes, and discusses preliminary measurements and future plans.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.P23G..05R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.P23G..05R"><span>Super Resolution and Interference Suppression Technique applied to SHARAD <span class="hlt">Radar</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>Raguso, M. C.; Mastrogiuseppe, M.; Seu, R.; Piazzo, L.</p> <p>2017-12-01</p> <p>We will present a super resolution and interference suppression technique applied to the data acquired by the SHAllow <span class="hlt">RADar</span> (SHARAD) on board the NASA's 2005 Mars Reconnaissance Orbiter (MRO) mission, currently operating around Mars [1]. The algorithms allow to improve the range resolution roughly by a factor of 3 and the Signal to Noise Ratio (SNR) by a several decibels. Range compression algorithms usually adopt conventional Fourier transform techniques, which are limited in the resolution by the transmitted signal bandwidth, analogous to the Rayleigh's criterion in optics. In this work, we investigate a super resolution method based on autoregressive models and linear prediction techniques [2]. Starting from the estimation of the linear prediction coefficients from the spectral data, the algorithm performs the <span class="hlt">radar</span> bandwidth extrapolation (BWE), thereby improving the range resolution of the pulse-compressed coherent <span class="hlt">radar</span> data. Moreover, the EMIs (ElectroMagnetic Interferences) are detected and the spectra is interpolated in order to reconstruct an interference free spectrum, thereby improving the SNR. The algorithm can be applied to the single complex look image after synthetic aperture processing (SAR). We apply the proposed algorithm to simulated as well as to real <span class="hlt">radar</span> data. We will demonstrate the effective enhancement on <span class="hlt">vertical</span> resolution with respect to the classical spectral estimator. We will show that the imaging of the subsurface layered structures observed in radargrams is improved, allowing additional insights for the scientific community in the interpretation of the SHARAD <span class="hlt">radar</span> data, which will help to further our understanding of the formation and evolution of known geological features on Mars. References: [1] Seu et al. 2007, Science, 2007, 317, 1715-1718 [2] K.M. Cuomo, "A Bandwidth Extrapolation Technique for Improved Range Resolution of Coherent <span class="hlt">Radar</span> Data", Project Report CJP-60, Revision 1, MIT Lincoln Laboratory (4 Dec. 1992).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JHyd..557..573P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JHyd..557..573P"><span>Accounting for rainfall evaporation using dual-polarization <span class="hlt">radar</span> and mesoscale model data</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pallardy, Quinn; Fox, Neil I.</p> <p>2018-02-01</p> <p>Implementation of dual-polarization <span class="hlt">radar</span> should allow for improvements in quantitative precipitation estimates due to dual-polarization capability allowing for the retrieval of the second moment of the gamma drop size distribution. Knowledge of the shape of the DSD can then be used in combination with mesoscale model data to estimate the motion and evaporation of each size of drop falling from the height at which precipitation is observed by the <span class="hlt">radar</span> to the surface. Using data from Central Missouri at a range between 130 and 140 km from the operational National Weather Service <span class="hlt">radar</span> a rain drop tracing scheme was developed to account for the effects of evaporation, where individual raindrops hitting the ground were traced to the <span class="hlt">point</span> in space and time where they interacted with the <span class="hlt">radar</span> beam. The results indicated evaporation played a significant role in <span class="hlt">radar</span> rainfall estimation in situations where the atmosphere was relatively dry. Improvements in <span class="hlt">radar</span> estimated rainfall were also found in these situations by accounting for evaporation. The conclusion was made that the effects of raindrop evaporation were significant enough to warrant further research into the inclusion high resolution model data in the <span class="hlt">radar</span> rainfall estimation process for appropriate locations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA571801','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA571801"><span>Social <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2012-01-01</p> <p>RTA HFM-201/RSM PAPER 3 - 1 © 2012 The MITRE Corporation. All Rights Reserved. Social <span class="hlt">Radar</span> Barry Costa and John Boiney MITRE Corporation...defenders require an integrated set of capabilities that we refer to as a “ social <span class="hlt">radar</span>.” Such a system would support strategic- to operational-level...situation awareness, alerting, course of action analysis, and measures of effectiveness for each action undertaken. Success of a social <span class="hlt">radar</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ISPAnIII8...93S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ISPAnIII8...93S"><span>Evaluation of <span class="hlt">Vertical</span> Lacunarity Profiles in Forested Areas Using Airborne Laser Scanning <span class="hlt">Point</span> Clouds</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Székely, B.; Kania, A.; Standovár, T.; Heilmeier, H.</p> <p>2016-06-01</p> <p>The horizontal variation and <span class="hlt">vertical</span> layering of the vegetation are important properties of the canopy structure determining the habitat; three-dimensional (3D) distribution of objects (shrub layers, understory vegetation, etc.) is related to the environmental factors (e.g., illumination, visibility). It has been shown that gaps in forests, mosaic-like structures are essential to biodiversity; various methods have been introduced to quantify this property. As the distribution of gaps in the vegetation is a multi-scale phenomenon, in order to capture it in its entirety, scale-independent methods are preferred; one of these is the calculation of lacunarity. We used Airborne Laser Scanning <span class="hlt">point</span> clouds measured over a forest plantation situated in a former floodplain. The flat topographic relief ensured that the tree growth is independent of the topographic effects. The tree pattern in the plantation crops provided various quasi-regular and irregular patterns, as well as various ages of the stands. The <span class="hlt">point</span> clouds were voxelized and layers of voxels were considered as images for two-dimensional input. These images calculated for a certain vicinity of reference <span class="hlt">points</span> were taken as images for the computation of lacunarity curves, providing a stack of lacunarity curves for each reference <span class="hlt">points</span>. These sets of curves have been compared to reveal spatial changes of this property. As the dynamic range of the lacunarity values is very large, the natural logarithms of the values were considered. Logarithms of lacunarity functions show canopy-related variations, we analysed these variations along transects. The spatial variation can be related to forest properties and ecology-specific aspects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840019041','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840019041"><span>An evaluation of the accuracy of some <span class="hlt">radar</span> wind profiling techniques</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Koscielny, A. J.; Doviak, R. J.</p> <p>1983-01-01</p> <p>Major advances in Doppler <span class="hlt">radar</span> measurement in optically clear air have made it feasible to monitor radial velocities in the troposphere and lower stratosphere. For most applications the three dimensional wind vector is monitored rather than the radial velocity. Measurement of the wind vector with a single <span class="hlt">radar</span> can be made assuming a spatially linear, time invariant wind field. The components and derivatives of the wind are estimated by the parameters of a linear regression of the radial velocities on functions of their spatial locations. The accuracy of the wind measurement thus depends on the locations of the radial velocities. The suitability is evaluated of some of the common retrieval techniques for simultaneous measurement of both the <span class="hlt">vertical</span> and horizontal wind components. The techniques considered for study are fixed beam, azimuthal scanning (VAD) and elevation scanning (VED).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870046160&hterms=Dew+point&qs=N%3D0%26Ntk%3DTitle%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDew%2Bpoint','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870046160&hterms=Dew+point&qs=N%3D0%26Ntk%3DTitle%26Ntx%3Dmode%2Bmatchall%26Ntt%3DDew%2Bpoint"><span><span class="hlt">Vertical</span> profiles of ozone, carbon monoxide, and dew-<span class="hlt">point</span> temperature obtained during GTE/CITE 1, October-November 1983. [Chemical Instrumentation Test and Evaluation</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fishman, Jack; Gregory, Gerald L.; Sachse, Glen W.; Beck, Sherwin M.; Hill, Gerald F.</p> <p>1987-01-01</p> <p>A set of 14 pairs of <span class="hlt">vertical</span> profiles of ozone and carbon monoxide, obtained with fast-response instrumentation, is presented. Most of these profiles, which were measured in the remote troposphere, also have supporting fast-response dew-<span class="hlt">point</span> temperature profiles. The data suggest that the continental boundary layer is a source of tropospheric ozone, even in October and November, when photochemical activity should be rather small. In general, the small-scale <span class="hlt">vertical</span> variability between CO and O3 is in phase. At low latitudes this relationship defines levels in the atmosphere where midlatitude air is being transported to lower latitudes, since lower dew-<span class="hlt">point</span> temperatures accompany these higher CO and O3 concentrations. A set of profiles which is suggestive of interhemispheric transport is also presented. Independent meteorological analyses support these interpretations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA579276','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA579276"><span>Effects of Stereoscopic 3D Digital <span class="hlt">Radar</span> Displays on Air Traffic Controller Performance</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>2013-03-01</p> <p>between men and women , but no significant influence was found. Experience in ATC was considered as a potential covariate that would be presumed to have...depicts altitude through the use of stereoscopic disparity, permitting <span class="hlt">vertical</span> separation to be visually represented as differences in disparity...handling information via different sources (e.g., <span class="hlt">radar</span> screen with a series of automated visual cues, paper or electronic flight progress strips, radio</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A34A..08C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A34A..08C"><span>Improving <span class="hlt">Radar</span> Quantitative Precipitation Estimation over Complex Terrain in the San Francisco Bay Area</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cifelli, R.; Chen, H.; Chandrasekar, V.</p> <p>2017-12-01</p> <p>A recent study by the State of California's Department of Water Resources has emphasized that the San Francisco Bay Area is at risk of catastrophic flooding. Therefore, accurate quantitative precipitation estimation (QPE) and forecast (QPF) are critical for protecting life and property in this region. Compared to rain gauge and meteorological satellite, ground based <span class="hlt">radar</span> has shown great advantages for high-resolution precipitation observations in both space and time domain. In addition, the polarization diversity shows great potential to characterize precipitation microphysics through identification of different hydrometeor types and their size and shape information. Currently, all the <span class="hlt">radars</span> comprising the U.S. National Weather Service (NWS) Weather Surveillance <span class="hlt">Radar</span>-1988 Doppler (WSR-88D) network are operating in dual-polarization mode. Enhancement of QPE is one of the main considerations of the dual-polarization upgrade. The San Francisco Bay Area is covered by two S-band WSR-88D <span class="hlt">radars</span>, namely, KMUX and KDAX. However, in complex terrain like the Bay Area, it is still challenging to obtain an optimal rainfall algorithm for a given set of dual-polarization measurements. In addition, the accuracy of rain rate estimates is contingent on additional factors such as bright band contamination, <span class="hlt">vertical</span> profile of reflectivity (VPR) correction, and partial beam blockages. This presentation aims to improve <span class="hlt">radar</span> QPE for the Bay area using advanced dual-polarization rainfall methodologies. The benefit brought by the dual-polarization upgrade of operational <span class="hlt">radar</span> network is assessed. In addition, a pilot study of gap fill X-band <span class="hlt">radar</span> performance is conducted in support of regional QPE system development. This paper also presents a detailed comparison between the dual-polarization <span class="hlt">radar</span>-derived rainfall products with various operational products including the NSSL's Multi-<span class="hlt">Radar</span>/Multi-Sensor (MRMS) system. Quantitative evaluation of various rainfall products is achieved</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AGUFMSA13B2359L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AGUFMSA13B2359L"><span>First Measurements of Polar Mesospheric Summer Echoes by a Tri-static <span class="hlt">Radar</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>La Hoz, C.</p> <p>2015-12-01</p> <p>Polar Mesospheric Summer Echoes (PMSE) have been observed for the first time by a tri-static <span class="hlt">radar</span> system comprising the EISCAT VHF (224 MHz, 0.67 m Bragg wavelength) active <span class="hlt">radar</span> in Tromso (Norway) and passive receiving stations in Kiruna, (Sweden) and Sodankyla (Finland). The antennas at the receiving stations, originally part of the EISCAT tri-static UHF <span class="hlt">radar</span> system at 930 MHz, have been refitted with new feeder systems at the VHF frequency of the transmitter in Tromso. The refitted <span class="hlt">radar</span> system opens new opportunities to study PMSE for its own sake and as a tracer of the dynamics of the polar mesosphere, a region that is difficult to investigate by other means. The measurements show that very frequently both remote receiving antennas detect coherent signals that are much greater than the regular incoherent scattering due to thermal electrons and coinciding in time and space with PMSE measured by the transmitter station in Tromso. This represents further evidence that PMSE is not aspect sensitive, as was already indicated by a less sensitive <span class="hlt">radar</span> system in a bi-static configuration, and implying that the underlying atmospheric turbulence, at least at sub-meter scales, is isotropic in agreement with Kolmogorov's hypothesis. Measurements also show that the <span class="hlt">vertical</span> rate of fall of persistent features of PMSE is the same as the <span class="hlt">vertical</span> line of sight velocity inferred from the doppler shift of the PMSE signals. This equivalence forms the basis for using PMSE as a tracer of the dynamics of the background mesosphere. Thus, it is possible to measure the 3-dimensional velocity field in the PMSE layer over the intersection volume of the three antennas. Since the signals have large signal-to-noise ratios (up to 30 dB), the inferred velocities have high accuracies and good time resolutions. This affords the possibility to make estimates of momentum flux in the mesosphere deposited by overturning gravity waves. Gravity wave momentum flux is believed to be the engine of a</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01770.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01770.html"><span>Space <span class="hlt">Radar</span> Image of Long Valley, California -Interferometry/Topography</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-05-01</p> <p> this area is about 1,320 meters (4,330 feet). Brightness variations come from the <span class="hlt">radar</span> image, which has been geometrically corrected to remove <span class="hlt">radar</span> distortions and rotated to have north toward the top. The image in the lower right is a three-dimensional perspective view of the northeast rim of the Long Valley caldera, looking toward the northwest. SIR-C C-band <span class="hlt">radar</span> image data are draped over topographic data derived from the interferometry processing. No <span class="hlt">vertical</span> exaggeration has been applied. Combining topographic and <span class="hlt">radar</span> image data allows scientists to examine relationships between geologic structures and landforms, and other properties of the land cover, such as soil type, vegetation distribution and hydrologic characteristics. http://photojournal.jpl.nasa.gov/catalog/PIA01770</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1068688','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1068688"><span>Determination of Large-Scale Cloud Ice Water Concentration by Combining Surface <span class="hlt">Radar</span> and Satellite Data in Support of ARM SCM Activities</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>Liu, Guosheng</p> <p>2013-03-15</p> <p>Single-column modeling (SCM) is one of the key elements of Atmospheric Radiation Measurement (ARM) research initiatives for the development and testing of various physical parameterizations to be used in general circulation models (GCMs). The data required for use with an SCM include observed <span class="hlt">vertical</span> profiles of temperature, water vapor, and condensed water, as well as the large-scale <span class="hlt">vertical</span> motion and tendencies of temperature, water vapor, and condensed water due to horizontal advection. Surface-based measurements operated at ARM sites and upper-air sounding networks supply most of the required variables for model inputs, but do not provide the horizontal advection term ofmore » condensed water. Since surface cloud <span class="hlt">radar</span> and microwave radiometer observations at ARM sites are single-<span class="hlt">point</span> measurements, they can provide the amount of condensed water at the location of observation sites, but not a horizontal distribution of condensed water contents. Consequently, observational data for the large-scale advection tendencies of condensed water have not been available to the ARM cloud modeling community based on surface observations alone. This lack of advection data of water condensate could cause large uncertainties in SCM simulations. Additionally, to evaluate GCMs cloud physical parameterization, we need to compare GCM results with observed cloud water amounts over a scale that is large enough to be comparable to what a GCM grid represents. To this end, the <span class="hlt">point</span>-measurements at ARM surface sites are again not adequate. Therefore, cloud water observations over a large area are needed. The main goal of this project is to retrieve ice water contents over an area of 10 x 10 deg. surrounding the ARM sites by combining surface and satellite observations. Built on the progress made during previous ARM research, we have conducted the retrievals of 3-dimensional ice water content by combining surface <span class="hlt">radar</span>/radiometer and satellite measurements, and have produced 3-D</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19810055743&hterms=oil+analysis+training&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doil%2Banalysis%2Btraining','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19810055743&hterms=oil+analysis+training&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doil%2Banalysis%2Btraining"><span>Navigation errors encountered using weather-mapping <span class="hlt">radar</span> for helicopter IFR guidance to oil rigs</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Phillips, J. D.; Bull, J. S.; Hegarty, D. M.; Dugan, D. C.</p> <p>1980-01-01</p> <p>In 1978 a joint NASA-FAA helicopter flight test was conducted to examine the use of weather-mapping <span class="hlt">radar</span> for IFR guidance during landing approaches to oil rig helipads. The following navigation errors were measured: total system error, <span class="hlt">radar</span>-range error, <span class="hlt">radar</span>-bearing error, and flight technical error. Three problem areas were identified: (1) operational problems leading to pilot blunders, (2) poor navigation to the downwind final approach <span class="hlt">point</span>, and (3) pure homing on final approach. Analysis of these problem areas suggests improvement in the <span class="hlt">radar</span> equipment, approach procedure, and pilot training, and gives valuable insight into the development of future navigation aids to serve the off-shore oil industry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20060043796&hterms=TURTLES&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DTURTLES','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060043796&hterms=TURTLES&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DTURTLES"><span>Fine resolution topographic mapping of the Jovian moons: a Ka-band high resolution topographic mapping interferometric synthetic aperture <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Madsen, Soren N.; Carsey, Frank D.; Turtle, Elizabeth P.</p> <p>2003-01-01</p> <p>The topographic data set obtained by MOLA has provided an unprecedented level of information about Mars' geologic features. The proposed flight of JIMO provides an opportunity to accomplish a similar mapping of and comparable scientific discovery for the Jovian moons through us of an interferometric imaging <span class="hlt">radar</span> analogous to the Shuttle <span class="hlt">radar</span> that recently generated a new topographic map of Earth. A Ka-band single pass across-track synthetic aperture <span class="hlt">radar</span> (SAR) interferometer can provide very high resolution surface elevation maps. The concept would use two antennas mounted at the ends of a deployable boom (similar to the Shuttle <span class="hlt">Radar</span> Topographic Mapper) extended orthogonal to the direction of flight. Assuming an orbit altitude of approximately 100 km and a ground velocity of approximately 1.5 km/sec, horizontal resolutions at the 10 meter level and <span class="hlt">vertical</span> resolutions at the sub-meter level are possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030066037&hterms=TURTLES&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DTURTLES','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030066037&hterms=TURTLES&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DTURTLES"><span>Fine Resolution Topographic Mapping of the Jovian Moons: A Ka-Band High Resolution Topographic Mapping Interferometric Synthetic Aperture <span class="hlt">Radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Madsen, S. N.; Carsey, F. D.; Turtle, E. P.</p> <p>2003-01-01</p> <p>The topographic data set obtained by MOLA has provided an unprecedented level of information about Mars' geologic features. The proposed flight of JIMO provides an opportunity to accomplish a similar mapping of and comparable scientific discovery for the Jovian moons through use of an interferometric imaging <span class="hlt">radar</span> analogous to the Shuttle <span class="hlt">radar</span> that recently generated a new topographic map of Earth. A Ka-band single pass across-track synthetic aperture <span class="hlt">radar</span> (SAR) interferometer can provide very high resolution surface elevation maps. The concept would use two antennas mounted at the ends of a deployable boom (similar to the Shuttle <span class="hlt">Radar</span> Topographic Mapper) extended orthogonal to the direction of flight. Assuming an orbit altitude of approximately 100km and a ground velocity of approximately 1.5 km/sec, horizontal resolutions at the 10 meter level and <span class="hlt">vertical</span> resolutions at the sub-meter level are possible.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-sts068-s-054.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-sts068-s-054.html"><span>STS-68 <span class="hlt">radar</span> image: Kilauea, Hawaii</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1994-10-10</p> <p>STS068-S-054 (10 October 1994) --- This is a deformation map of the south flank of Kilauea volcano on the big island of Hawaii, centered at 19.5 degrees north latitude and 155.25 degrees west longitude. The map was created by combining interferometric <span class="hlt">radar</span> data - that is data acquired on different passes of the Space Shuttle Endeavour which are then overlaid to obtain elevation information - acquired by the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-Band Synthetic Aperture <span class="hlt">Radar</span> (SIR-C/X-SAR) during its first flight in April 1994 and its second flight in October 1994. The area shown is approximately 40 by 80 kilometers (25 by 50 miles). North is toward the upper left of the image. The colors indicate the displacement of the surface in that direction that the <span class="hlt">radar</span> instrument was <span class="hlt">pointed</span> (toward the right of the image) in the six months between images. The analysis of ground movement is preliminary, but appears consistent with the motions detected by the Global Positioning System ground receivers that have been used over the past five years. The south flank of the Kilauea volcano is among the most rapidly deforming terrain's on Earth. Several regions show motion over the six-month time period. Most obvious is at the base of Hilina Pali, where 10 centimeters (4 inches) or more of crustal deformation can be seen in a concentrated area near the coastline. On a more localized scale, the currently active Pu'u O'o summit also shows about 10 centimeters (4 inches) of change near the vent area. Finally, there are indications of additional movement along the upper southwest rift zone, just below the Kilauea caldera in the image. Deformation of the south flank is believed to be the result of movements along faults deep beneath the surface of the volcano, as well as injections of magma, or molten rock, into the volcano's "plumbing" system. Detection of ground motions from space has proven to be a unique capability of imaging <span class="hlt">radar</span> technology. Scientists hope to use deformation data</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1995SPIE.2592..169Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1995SPIE.2592..169Z"><span>Millimeter wave scattering characteristics and <span class="hlt">radar</span> cross section measurements of common roadway objects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zoratti, Paul K.; Gilbert, R. Kent; Majewski, Ronald; Ference, Jack</p> <p>1995-12-01</p> <p>Development of automotive collision warning systems has progressed rapidly over the past several years. A key enabling technology for these systems is millimeter-wave <span class="hlt">radar</span>. This paper addresses a very critical millimeter-wave <span class="hlt">radar</span> sensing issue for automotive <span class="hlt">radar</span>, namely the scattering characteristics of common roadway objects such as vehicles, roadsigns, and bridge overpass structures. The data presented in this paper were collected on ERIM's Fine Resolution <span class="hlt">Radar</span> Imaging Rotary Platform Facility and processed with ERIM's image processing tools. The value of this approach is that it provides system developers with a 2D <span class="hlt">radar</span> image from which information about individual <span class="hlt">point</span> scatterers `within a single target' can be extracted. This information on scattering characteristics will be utilized to refine threat assessment processing algorithms and automotive <span class="hlt">radar</span> hardware configurations. (1) By evaluating the scattering characteristics identified in the <span class="hlt">radar</span> image, <span class="hlt">radar</span> signatures as a function of aspect angle for common roadway objects can be established. These signatures will aid in the refinement of threat assessment processing algorithms. (2) Utilizing ERIM's image manipulation tools, total RCS and RCS as a function of range and azimuth can be extracted from the <span class="hlt">radar</span> image data. This RCS information will be essential in defining the operational envelope (e.g. dynamic range) within which any <span class="hlt">radar</span> sensor hardware must be designed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199425p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199425p/"><span>33. Perimeter acquisition <span class="hlt">radar</span> building room #320, perimeter acquisition <span class="hlt">radar</span> ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>33. Perimeter acquisition <span class="hlt">radar</span> building room #320, perimeter acquisition <span class="hlt">radar</span> operations center (PAROC), contains the tactical command and control group equipment required to control the par site. Showing spacetrack monitor console - Stanley R. Mickelsen Safeguard Complex, Perimeter Acquisition <span class="hlt">Radar</span> Building, Limited Access Area, between Limited Access Patrol Road & Service Road A, Nekoma, Cavalier County, ND</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PEPS....4...19K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PEPS....4...19K"><span>Shigaraki UAV-<span class="hlt">Radar</span> Experiment (ShUREX): overview of the campaign with some preliminary results</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kantha, Lakshmi; Lawrence, Dale; Luce, Hubert; Hashiguchi, Hiroyuki; Tsuda, Toshitaka; Wilson, Richard; Mixa, Tyler; Yabuki, Masanori</p> <p>2017-12-01</p> <p>The Shigaraki unmanned aerial vehicle (UAV)-<span class="hlt">Radar</span> Experiment (ShUREX) is an international (USA-Japan-France) observational campaign, whose overarching goal is to demonstrate the utility of small, lightweight, inexpensive, autonomous UAVs in probing and monitoring the lower troposphere and to promote synergistic use of UAVs and very high frequency (VHF) <span class="hlt">radars</span>. The 2-week campaign lasting from June 1 to June 14, 2015, was carried out at the Middle and Upper Atmosphere (MU) Observatory in Shigaraki, Japan. During the campaign, the DataHawk UAV, developed at the University of Colorado, Boulder, and equipped with high-frequency response cold wire and pitot tube sensors (as well as an iMET radiosonde), was flown near and over the VHF-band MU <span class="hlt">radar</span>. Measurements in the atmospheric column in the immediate vicinity of the <span class="hlt">radar</span> were obtained. Simultaneous and continuous operation of the <span class="hlt">radar</span> in range imaging mode enabled fine-scale structures in the atmosphere to be visualized by the <span class="hlt">radar</span>. It also permitted the UAV to be commanded to sample interesting structures, guided in near real time by the <span class="hlt">radar</span> images. This overview provides a description of the ShUREX campaign and some interesting but preliminary results of the very first simultaneous and intensive probing of turbulent structures by UAVs and the MU <span class="hlt">radar</span>. The campaign demonstrated the validity and utility of the <span class="hlt">radar</span> range imaging technique in obtaining very high <span class="hlt">vertical</span> resolution ( 20 m) images of echo power in the atmospheric column, which display evolving fine-scale atmospheric structures in unprecedented detail. The campaign also permitted for the very first time the evaluation of the consistency of turbulent kinetic energy dissipation rates in turbulent structures inferred from the spectral broadening of the backscattered <span class="hlt">radar</span> signal and direct, in situ measurements by the high-frequency response velocity sensor on the UAV. The data also enabled other turbulence parameters such as the temperature</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.A31A2147G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.A31A2147G"><span><span class="hlt">Radar</span>-derived Quantitative Precipitation Estimation in Complex Terrain over the Eastern Tibetan Plateau</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gou, Y.</p> <p>2017-12-01</p> <p>Quantitative Precipitation Estimation (QPE) is one of the important applications of weather <span class="hlt">radars</span>. However, in complex terrain such as Tibetan Plateau, it is a challenging task to obtain an optimal Z-R relation due to the complex space time variability in precipitation microphysics. This paper develops two <span class="hlt">radar</span> QPE schemes respectively based on Reflectivity Threshold (RT) and Storm Cell Identification and Tracking (SCIT) algorithms using observations from 11 Doppler weather <span class="hlt">radars</span> and 3294 rain gauges over the Eastern Tibetan Plateau (ETP). These two QPE methodologies are evaluated extensively using four precipitation events that are characterized by different meteorological features. Precipitation characteristics of independent storm cells associated with these four events, as well as the storm-scale differences, are investigated using short-term <span class="hlt">vertical</span> profiles of reflectivity clusters. Evaluation results show that the SCIT-based rainfall approach performs better than the simple RT-based method in all precipitation events in terms of score comparison using validation gauge measurements as references, with higher correlation (than 75.74%), lower mean absolute error (than 82.38%) and root-mean-square error (than 89.04%) of all the comparative frames. It is also found that the SCIT-based approach can effectively mitigate the <span class="hlt">radar</span> QPE local error and represent precipitation spatiotemporal variability better than RT-based scheme.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9833E..0EC','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9833E..0EC"><span>Capturing atmospheric effects on 3D millimeter wave <span class="hlt">radar</span> propagation patterns</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cook, Richard D.; Fiorino, Steven T.; Keefer, Kevin J.; Stringer, Jeremy</p> <p>2016-05-01</p> <p>Traditional <span class="hlt">radar</span> propagation modeling is done using a path transmittance with little to no input for weather and atmospheric conditions. As <span class="hlt">radar</span> advances into the millimeter wave (MMW) regime, atmospheric effects such as attenuation and refraction become more pronounced than at traditional <span class="hlt">radar</span> wavelengths. The DoD High Energy Laser Joint Technology Offices High Energy Laser End-to-End Operational Simulation (HELEEOS) in combination with the Laser Environmental Effects Definition and Reference (LEEDR) code have shown great promise simulating atmospheric effects on laser propagation. Indeed, the LEEDR radiative transfer code has been validated in the UV through RF. Our research attempts to apply these models to characterize the far field <span class="hlt">radar</span> pattern in three dimensions as a signal propagates from an antenna towards a <span class="hlt">point</span> in space. Furthermore, we do so using realistic three dimensional atmospheric profiles. The results from these simulations are compared to those from traditional <span class="hlt">radar</span> propagation software packages. In summary, a fast running method has been investigated which can be incorporated into computational models to enhance understanding and prediction of MMW propagation through various atmospheric and weather conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AGUFM.P23A1616C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AGUFM.P23A1616C"><span>An enhanced Planetary <span class="hlt">Radar</span> Operating Centre (PROC)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Catallo, C.</p> <p>2010-12-01</p> <p>Planetary exploration by means of <span class="hlt">radar</span> systems, mainly using GPRs is an important role of Italy and numerous scientific international space programs are carried out jointly with ESA and NASA by Italian Space Agency, the scientific community and the industry. Three experiments under Italian leadership ( designed and manufactured by the Italian industry) provided by ASI within a NASA/ESA/ASI joint venture framework are successfully operating: MARSIS on-board MEX, SHARAD on-board MRO and CASSINI <span class="hlt">Radar</span> on-board Cassini spacecraft: the missions have been further extended . Three dedicated operational centers, namely SHOC, (Sharad Operating Centre), MOC (Marsis Operating Center) and CASSINI PAD are operating from the missions beginning to support all the scientific communities, institutional customers and experiment teams operation Each center is dedicated to a single instrument management and control, data processing and distribution and even if they had been conceived to operate autonomously and independently one from each other, synergies and overlaps have been envisaged leading to the suggestion of a unified center, the Planetary <span class="hlt">Radar</span> Processing Center (PROC). In order to harmonize operations either from logistics <span class="hlt">point</span> of view and from HW/SW capabilities <span class="hlt">point</span> of view PROC is designed and developed for offering improved functionalities to increase capabilities, mainly in terms of data exchange, comparison, interpretation and exploitation. PROC is, therefore, conceived as the Italian support facility to the scientific community for on-going and future Italian planetary exploration programs, such as Europa-Jupiter System Mission (EJSM) The paper describes how the new PROC is designed and developed, to allow SHOC, MOC and CASSINI PAD to operate as before, and to offer improved functionalities to increase capabilities, mainly in terms of data exchange, comparison, interpretation and exploitation aiding scientists to increase their knowledge in the field of surface</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720017712','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720017712"><span>probing the atmosphere with high power, high resolution <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hardy, K. R.; Katz, I.</p> <p>1969-01-01</p> <p>Observations of <span class="hlt">radar</span> echoes from the clear atmosphere are presented and the scattering mechanisms responsible for the two basic types of clear-air echoes are discussed. The commonly observed dot echo originates from a <span class="hlt">point</span> in space and usually shows little variation in echo intensity over periods of about 0.1 second. The second type of clear-air <span class="hlt">radar</span> echo appears diffuse in space, and signal intensities vary considerably over periods of less than 0.1 second. The echoes often occur in thin horizontal layers or as boundaries of convective activity; these are characterized by sharp gradients of refractive index. Some features of clear-air atmospheric structures as observed with <span class="hlt">radar</span> are presented. These structures include thin stable inversions, convective thermals, Benard convection cells, breaking gravity waves, and high tropospheric layers which are sufficiently turbulent to affect aircraft.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA550922','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA550922"><span>Solar <span class="hlt">Radar</span> Experiments</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1998-01-01</p> <p>communications satellites and electric power grids. RELATED PROJECTS Studies with the HAARP <span class="hlt">radar</span> facility being constructed in Alaska are conducted with...on wave-plasma interactions and also are assessing the possible use of HAARP as a solar <span class="hlt">radar</span>. REFERENCES James, J. C., <span class="hlt">Radar</span> studies of the sun, in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19840019081','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19840019081"><span>Design considerations for high-power VHF <span class="hlt">radar</span> transceivers: Phase matching long coaxial cables using a cable <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Johnson, P. E.; Ecklund, W. L.</p> <p>1983-01-01</p> <p>The Poker Flat 49.92-MHz MST <span class="hlt">radar</span> uses 64 phase-controlled transmitters in individual shelters distributed throughout the antenna array. Phase control is accomplished by sampling the transmitted pulse at the directional coupler of each transmitter and sending the sample pulse back to a phase-control unit. This method requires phase matching 64 long (256 meter) coaxial cables (RG-213) to within several electrical degrees. Tests with a time domain reflectometer showed that attenuation of high frequency components in the long RG-213 cable rounded the leading edge of the reflected pulse so that the cables could only be measured to within 50 cm (about 45 deg at 49.92 MHz). Another measurement technique using a vector voltmeter to compare forward and reflected phase required a directional coupler with unattainable directivity. Several other techniques were also found lacking, primarily because of loss in the long RG-213 cables. At this <span class="hlt">point</span> it was realized that what was needed was a simple version of the phase-coherent clear-air <span class="hlt">radar</span>, i.e., a cable <span class="hlt">radar</span>. The design and operation of this cable are described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27547484','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27547484"><span><span class="hlt">Vertical</span> load capacities of roof truss cross members.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gearhart, David F; Morsy, Mohamed Khaled</p> <p>2016-05-01</p> <p>Trusses used for roof support in coal mines are constructed of two grouted bolts installed at opposing forty-five degree angles into the roof and a cross member that ties the angled bolts together. The load on the cross member is <span class="hlt">vertical</span>, which is transverse to the longitudinal axis, and therefore the cross member is loaded in the weakest direction. Laboratory tests were conducted to determine the <span class="hlt">vertical</span> load capacity and deflection of three different types of cross members. Single-<span class="hlt">point</span> load tests, with the load applied in the center of the specimen and double-<span class="hlt">point</span> load tests, with a span of 2.4 m, were conducted. For the single-<span class="hlt">point</span> load configuration, the yield of the 25 mm solid bar cross member was nominally 98 kN of <span class="hlt">vertical</span> load, achieved at 42 cm of deflection. For cable cross members, yield was not achieved even after 45 cm of deflection. Peak <span class="hlt">vertical</span> loads were about 89 kN for 17 mm cables and 67 kN for the 15 mm cables. For the double-<span class="hlt">point</span> load configurations, the 25 mm solid bar cross members yielded at 150 kN of <span class="hlt">vertical</span> load and 25 cm of deflection. At 25 cm of deflection individual cable strands started breaking at 133 and 111 kN of <span class="hlt">vertical</span> load for the 17 and 15 mm cable cross members respectively.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720017713','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720017713"><span>Furthur remarks on atmospheric probing by ultrasensitive <span class="hlt">radar</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Atlas, D.</p> <p>1969-01-01</p> <p>This paper is supplementary to that of Hardy and Katz. It emphasizes the meteorological value of the various capabilities of ultrasensitive <span class="hlt">radar</span>, highlights the <span class="hlt">points</span> of agreement and disagreement, and focuses upon the directions of promising research. The theory of backscatter from a refractively turbulent region is said to be confirmed by the <span class="hlt">radar</span> observations both with respect to magnitude and wavelength dependence. A reason for the apparent discrepancy between the results of some of the forwardscatter experiments and theory is suggested. Disagreement still exists with respect to the origin of clear air sea breeze echoes; the author does not agree with Hardy and Katz that they are due to insects. However, it is agreed that some unusually widespread echo displays on clear days are indeed due to insects. The meteorological value of ultrasensitive <span class="hlt">radars</span> demonstrated by Hardy and Katz, here, and by others is so profound as to demand their use in remote atmospheric probing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.A33Q..06K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.A33Q..06K"><span>Scanning Cloud <span class="hlt">Radar</span> Observations at the ARM sites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kollias, P.; Clothiaux, E. E.; Shupe, M.; Widener, K.; Bharadwaj, N.; Miller, M. A.; Verlinde, H.; Luke, E. P.; Johnson, K. L.; Jo, I.; Tatarevic, A.; Lamer, K.</p> <p>2012-12-01</p> <p>Recently, the DOE Atmospheric Radiation Measurement (ARM) program upgraded its fixed and mobile facilities with the acquisition of state-of-the-art scanning, dual-wavelength, polarimetric, Doppler cloud <span class="hlt">radars</span>. The scanning ARM cloud <span class="hlt">radars</span> (SACR's) are the most expensive and significant <span class="hlt">radar</span> systems at all ARM sites and eight SACR systems will be operational at ARM sites by the end of 2013. The SACR's are the primary instruments for the detection of 3D cloud properties (boundaries, volume cloud fractional coverage, liquid water content, dynamics, etc.) beyond the soda-straw (profiling) limited view. Having scanning capabilities with two frequencies and polarization allows more accurate probing of a variety of cloud systems (e.g., drizzle and shallow, warm rain), better correction for attenuation, use of attenuation for liquid water content retrievals, and polarimetric and dual-wavelength ratio characterization of non-spherical particles for improved ice crystal habit identification. Examples of SACR observations from four ARM sites are presented here: the fixed sites at Southern Great Plains (SGP) and North Slope of Alaska (NSA), and the mobile facility deployments at Graciosa Island, Azores and Cape Cod, Massachusetts. The 3D cloud structure is investigated both at the macro-scale (20-50 km) and cloud-scale (100-500 m). Doppler velocity measurements are corrected for velocity folding and are used either to describe the in-cloud horizontal wind profile or the 3D <span class="hlt">vertical</span> air motions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007PhDT........46B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007PhDT........46B"><span>Extending interferometric synthetic aperture <span class="hlt">radar</span> measurements from one to two dimensions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bechor, Noah</p> <p></p> <p>Interferometric synthetic aperture <span class="hlt">radar</span> (InSAR), a very effective technique for measuring crustal deformation, provides measurements in only one dimension, along the <span class="hlt">radar</span> line of sight. Imaging <span class="hlt">radar</span> measurements from satellite-based systems are sensitive to both <span class="hlt">vertical</span> and across-track displacements, but insensitive to along-track displacement. Multiple observations can resolve the first two components, but the along-track component remains elusive. The best existing method to obtain the along-track displacement involves pixel-level azimuth cross-correlation. The measurements are quite coarse (typically 15 cm precision), and they require large computation times. In contrast, across-track and <span class="hlt">vertical</span> InSAR measurements can reach centimeter-level precision and are readily derived. We present a new method to extract along-track displacements from InSAR data. The new method, multiple aperture InSAR (MAI), is based on split-beam processing of InSAR data to create forward- and backward-looking interferograms. The phase difference between the two modified interferograms provides the along-track displacement component. Thus, from each conventional InSAR pair we extract two components of the displacement vector: one along the line of sight, the other in the along-track direction. Multiple MAI observations, either at two look angles or from the ascending and descending <span class="hlt">radar</span> passes, then yield the three-dimensional displacement field. We analyze precision of our method by comparing our solution to GPS and offset-derived along-track displacements from interferograms of the M7.1 1999, Hector Mine earthquake. The RMS error between GPS displacements and our results ranges from 5 to 8.8cm. Our method is consistent with along-track displacements derived by pixel-offsets, themselves limited to 12-15cm precision. The theoretical MAI precision depends on SNR and coherence. For SNR=100 the expected precision is 3, 11cm for coherence of 0.8, 0.4, respectively. Finally, we</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940012280','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940012280"><span>External calibration of polarimetric <span class="hlt">radar</span> images using distributed targets</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yueh, Simon H.; Nghiem, S. V.; Kwok, R.</p> <p>1992-01-01</p> <p>A new technique is presented for calibrating polarimetric synthetic aperture <span class="hlt">radar</span> (SAR) images using only the responses from natural distributed targets. The model for polarimetric <span class="hlt">radars</span> is assumed to be X = cRST where X is the measured scattering matrix corresponding to the target scattering matrix S distorted by the system matrices T and R (in general T does not equal R(sup t)). To allow for the polarimetric calibration using only distributed targets and corner reflectors, van Zyl assumed a reciprocal polarimetric <span class="hlt">radar</span> model with T = R(sup t); when applied for JPL SAR data, a heuristic symmetrization procedure is used by POLCAL to compensate the phase difference between the measured HV and VH responses and then take the average of both. This heuristic approach causes some non-removable cross-polarization responses for corner reflectors, which can be avoided by a rigorous symmetrization method based on reciprocity. After the <span class="hlt">radar</span> is made reciprocal, a new algorithm based on the responses from distributed targets with reflection symmetry is developed to estimate the cross-talk parameters. The new algorithm never experiences problems in convergence and is also found to converge faster than the existing routines implemented for POLCAL. When the new technique is implemented for the JPL polarimetric data, symmetrization and cross-talk removal are performed on a line-by-line (azimuth) basis. After the cross-talks are removed from the entire image, phase and amplitude calibrations are carried out by selecting distributed targets either with azimuthal symmetry along the looking direction or with some well-known volume and surface scattering mechanisms to estimate the relative phases and amplitude responses of the horizontal and <span class="hlt">vertical</span> channels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21774418','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21774418"><span>Using Lidar and <span class="hlt">Radar</span> measurements to constrain predictions of forest ecosystem structure and function.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Antonarakis, Alexander S; Saatchi, Sassan S; Chazdon, Robin L; Moorcroft, Paul R</p> <p>2011-06-01</p> <p>Insights into vegetation and aboveground biomass dynamics within terrestrial ecosystems have come almost exclusively from ground-based forest inventories that are limited in their spatial extent. Lidar and synthetic-aperture <span class="hlt">Radar</span> are promising remote-sensing-based techniques for obtaining comprehensive measurements of forest structure at regional to global scales. In this study we investigate how Lidar-derived forest heights and <span class="hlt">Radar</span>-derived aboveground biomass can be used to constrain the dynamics of the ED2 terrestrial biosphere model. Four-year simulations initialized with Lidar and <span class="hlt">Radar</span> structure variables were compared against simulations initialized from forest-inventory data and output from a long-term potential-vegtation simulation. Both height and biomass initializations from Lidar and <span class="hlt">Radar</span> measurements significantly improved the representation of forest structure within the model, eliminating the bias of too many large trees that arose in the potential-vegtation-initialized simulation. The Lidar and <span class="hlt">Radar</span> initializations decreased the proportion of larger trees estimated by the potential vegetation by approximately 20-30%, matching the forest inventory. This resulted in improved predictions of ecosystem-scale carbon fluxes and structural dynamics compared to predictions from the potential-vegtation simulation. The <span class="hlt">Radar</span> initialization produced biomass values that were 75% closer to the forest inventory, with Lidar initializations producing canopy height values closest to the forest inventory. Net primary production values for the <span class="hlt">Radar</span> and Lidar initializations were around 6-8% closer to the forest inventory. Correcting the Lidar and <span class="hlt">Radar</span> initializations for forest composition resulted in improved biomass and basal-area dynamics as well as leaf-area index. Correcting the Lidar and <span class="hlt">Radar</span> initializations for forest composition and fine-scale structure by combining the remote-sensing measurements with ground-based inventory data further improved</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930009531','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930009531"><span>A comparison of airborne and ground-based <span class="hlt">radar</span> observations with rain gages during the CaPE experiment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Satake, Makoto; Short, David A.; Iguchi, Toshio</p> <p>1992-01-01</p> <p>The vicinity of KSC, where the primary ground truth site of the Tropical Rainfall Measuring Mission (TRMM) program is located, was the focal <span class="hlt">point</span> of the Convection and Precipitation/Electrification (CaPE) experiment in Jul. and Aug. 1991. In addition to several specialized <span class="hlt">radars</span>, local coverage was provided by the C-band (5 cm) <span class="hlt">radar</span> at Patrick AFB. <span class="hlt">Point</span> measurements of rain rate were provided by tipping bucket rain gage networks. Besides these ground-based activities, airborne <span class="hlt">radar</span> measurements with X- and Ka-band nadir-looking <span class="hlt">radars</span> on board an aircraft were also recorded. A unique combination data set of airborne <span class="hlt">radar</span> observations with ground-based observations was obtained in the summer convective rain regime of central Florida. We present a comparison of these data intending a preliminary validation. A convective rain event was observed simultaneously by all three instrument types on the evening of 27 Jul. 1991. The high resolution aircraft <span class="hlt">radar</span> was flown over convective cells with tops exceeding 10 km and observed reflectivities of 40 to 50 dBZ at 4 to 5 km altitude, while the low resolution surface <span class="hlt">radar</span> observed 35 to 55 dBZ echoes and a rain gage indicated maximum surface rain rates exceeding 100 mm/hr. The height profile of reflectivity measured with the airborne <span class="hlt">radar</span> show an attenuation of 6.5 dB/km (two way) for X-band, corresponding to a rainfall rate of 95 mm/hr.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199433p/','SCIGOV-HHH'); return false;" href="https://www.loc.gov/pictures/collection/hh/item/nd0078.photos.199433p/"><span>41. Perimeter acquisition <span class="hlt">radar</span> building <span class="hlt">radar</span> element and coaxial display, ...</span></a></p> <p><a target="_blank" href="http://www.loc.gov/pictures/collection/hh/">Library of Congress Historic Buildings Survey, Historic Engineering Record, Historic Landscapes Survey</a></p> <p></p> <p></p> <p>41. Perimeter acquisition <span class="hlt">radar</span> building <span class="hlt">radar</span> element and coaxial display, with drawing of typical antenna section. Drawing, from left to right, shows element, aluminum ground plane, cable connectors and hardware, cable, and back-up ring. Grey area is the concrete wall - Stanley R. Mickelsen Safeguard Complex, Perimeter Acquisition <span class="hlt">Radar</span> Building, Limited Access Area, between Limited Access Patrol Road & Service Road A, Nekoma, Cavalier County, ND</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860041465&hterms=digital+transformation&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Ddigital%2Btransformation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860041465&hterms=digital+transformation&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Ddigital%2Btransformation"><span>Digital image transformation and rectification of spacecraft and <span class="hlt">radar</span> images</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wu, S. S. C.</p> <p>1985-01-01</p> <p>The application of digital processing techniques to spacecraft television pictures and <span class="hlt">radar</span> images is discussed. The use of digital rectification to produce contour maps from spacecraft pictures is described; images with azimuth and elevation angles are converted into <span class="hlt">point</span>-perspective frame pictures. The digital correction of the slant angle of <span class="hlt">radar</span> images to ground scale is examined. The development of orthophoto and stereoscopic shaded relief maps from digital terrain and digital image data is analyzed. Digital image transformations and rectifications are utilized on Viking Orbiter and Lander pictures of Mars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004SPIE.5403..673S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004SPIE.5403..673S"><span>Maritime microwave <span class="hlt">radar</span> and electro-optical data fusion for homeland security</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seastrand, Mark J.</p> <p>2004-09-01</p> <p>US Customs is responsible for monitoring all incoming air and maritime traffic, including the island of Puerto Rico as a US territory. Puerto Rico offers potentially obscure <span class="hlt">points</span> of entry to drug smugglers. This environment sets forth a formula for an illegal drug trade - based relatively near the continental US. The US Customs Caribbean Air and Marine Operations Center (CAMOC), located in Puntas Salinas, has the charter to monitor maritime and Air Traffic Control (ATC) <span class="hlt">radars</span>. The CAMOC monitors ATC <span class="hlt">radars</span> and advises the Air and Marine Branch of US Customs of suspicious air activity. In turn, the US Coast Guard and/or US Customs will launch air and sea assets as necessary. The addition of a coastal <span class="hlt">radar</span> and camera system provides US Customs a maritime monitoring capability for the northwestern end of Puerto Rico (Figure 1). Command and Control of the <span class="hlt">radar</span> and camera is executed at the CAMOC, located 75 miles away. The Maritime Microwave Surveillance <span class="hlt">Radar</span> performs search, primary target acquisition and target tracking while the Midwave Infrared (MWIR) camera performs target identification. This wide area surveillance, using a combination of <span class="hlt">radar</span> and MWIR camera, offers the CAMOC a cost and manpower effective approach to monitor, track and identify maritime targets.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=PIA09096&hterms=mars+climate+orbiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3D%2527mars%2Bclimate%2Borbiter%2527','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=PIA09096&hterms=mars+climate+orbiter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3D%2527mars%2Bclimate%2Borbiter%2527"><span>Interpreting <span class="hlt">Radar</span> View near Mars' South Pole, Orbit 1334</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2006-01-01</p> <p><p/> A radargram from the Shallow Subsurface <span class="hlt">Radar</span> instrument (SHARAD) on NASA's Mars Reconnaissance Orbiter is shown in the upper-right panel and reveals detailed structure in the polar layered deposits of the south pole of Mars. <p/> The sounding <span class="hlt">radar</span> collected the data presented here during orbit 1334 of the mission, on Nov. 8, 2006. <p/> The horizontal scale in the radargram is distance along the ground track. It can be referenced to the ground track map shown in the lower right. The <span class="hlt">radar</span> traversed from about 75 to 85 degrees south latitude, or about 590 kilometers (370 miles). The ground track map shows elevation measured by the Mars Orbiter Laser Altimeter on NASA's Mars Global Surveyor orbiter. Green indicates low elevation; reddish-white indicates higher elevation. The traverse proceeds up onto a plateau formed by the layers. <p/> The <span class="hlt">vertical</span> scale on the radargram is time delay of the <span class="hlt">radar</span> signals reflected back to Mars Reconnaissance Orbiter from the surface and subsurface. For reference, using an assumed velocity of the <span class="hlt">radar</span> waves in the subsurface, time is converted to depth below the surface at one place: about 1,500 meters (5,000 feet) to one of the deeper subsurface reflectors. The color scale varies from black for weak reflections to white for strong reflections. <p/> The middle panel shows mapping of the major subsurface reflectors, some of which can be traced for a distance of 100 kilometers (60 miles) or more. The layers are not all horizontal and the reflectors are not always parallel to one another. Some of this is due to variations in surface elevation, which produce differing velocity path lengths for different reflector depths. However, some of this behavior is due to spatial variations in the deposition and removal of material in the layered deposits, a result of the recent climate history of Mars. <p/> The Shallow Subsurface <span class="hlt">Radar</span> was provided by the Italian Space Agency (ASI). Its operations are led by the University of Rome and its</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhDT.......135S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhDT.......135S"><span>Waveform-Diverse Multiple-Input Multiple-Output <span class="hlt">Radar</span> Imaging Measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stewart, Kyle B.</p> <p></p> <p>Multiple-input multiple-output (MIMO) <span class="hlt">radar</span> is an emerging set of technologies designed to extend the capabilities of multi-channel <span class="hlt">radar</span> systems. While conventional <span class="hlt">radar</span> architectures emphasize the use of antenna array beamforming to maximize real-time power on target, MIMO <span class="hlt">radar</span> systems instead attempt to preserve some degree of independence between their received signals and to exploit this expanded matrix of target measurements in the signal-processing domain. Specifically the use of sparse “virtual” antenna arrays may allow MIMO <span class="hlt">radars</span> to achieve gains over traditional multi-channel systems by post-processing diverse received signals to implement both transmit and receive beamforming at all <span class="hlt">points</span> of interest within a given scene. MIMO architectures have been widely examined for use in <span class="hlt">radar</span> target detection, but these systems may yet be ideally suited to real and synthetic aperture <span class="hlt">radar</span> imaging applications where their proposed benefits include improved resolutions, expanded area coverage, novel modes of operation, and a reduction in hardware size, weight, and cost. While MIMO <span class="hlt">radar</span>'s theoretical benefits have been well established in the literature, its practical limitations have not received great attention thus far. The effective use of MIMO <span class="hlt">radar</span> techniques requires a diversity of signals, and to date almost all MIMO system demonstrations have made use of time-staggered transmission to satisfy this requirement. Doing so is reliable but can be prohibitively slow. Waveform-diverse systems have been proposed as an alternative in which multiple, independent waveforms are broadcast simultaneously over a common bandwidth and separated on receive using signal processing. Operating in this way is much faster than its time-diverse equivalent, but finding a set of suitable waveforms for this technique has proven to be a difficult problem. In light of this, many have questioned the practicality of MIMO <span class="hlt">radar</span> imaging and whether or not its theoretical benefits</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=gain+AND+function&id=EJ1019617','ERIC'); return false;" href="https://eric.ed.gov/?q=gain+AND+function&id=EJ1019617"><span>The Gains from <span class="hlt">Vertical</span> Scaling</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Briggs, Derek C.; Domingue, Ben</p> <p>2013-01-01</p> <p>It is often assumed that a <span class="hlt">vertical</span> scale is necessary when value-added models depend upon the gain scores of students across two or more <span class="hlt">points</span> in time. This article examines the conditions under which the scale transformations associated with the <span class="hlt">vertical</span> scaling process would be expected to have a significant impact on normative interpretations…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3678487','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3678487"><span>An in situ approach to detect tree root ecology: linking ground-penetrating <span class="hlt">radar</span> imaging to isotope-derived water acquisition zones</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Isaac, Marney E; Anglaaere, Luke C N</p> <p>2013-01-01</p> <p>Tree root distribution and activity are determinants of belowground competition. However, studying root response to environmental and management conditions remains logistically challenging. Methodologically, nondestructive in situ tree root ecology analysis has lagged. In this study, we tested a nondestructive approach to determine tree coarse root architecture and function of a perennial tree crop, Theobroma cacao L., at two edaphically contrasting sites (sandstone and phyllite–granite derived soils) in Ghana, West Africa. We detected coarse root <span class="hlt">vertical</span> distribution using ground-penetrating <span class="hlt">radar</span> and root activity via soil water acquisition using isotopic matching of δ18O plant and soil signatures. Coarse roots were detected to a depth of 50 cm, however, intraspecifc coarse root <span class="hlt">vertical</span> distribution was modified by edaphic conditions. Soil δ18O isotopic signature declined with depth, providing conditions for plant–soil δ18O isotopic matching. This pattern held only under sandstone conditions where water acquisition zones were identifiably narrow in the 10–20 cm depth but broader under phyllite–granite conditions, presumably due to resource patchiness. Detected coarse root count by depth and measured fine root density were strongly correlated as were detected coarse root count and identified water acquisition zones, thus validating root detection capability of ground-penetrating <span class="hlt">radar</span>, but exclusively on sandstone soils. This approach was able to characterize trends between intraspecific root architecture and edaphic-dependent resource availability, however, limited by site conditions. This study successfully demonstrates a new approach for in situ root studies that moves beyond invasive <span class="hlt">point</span> sampling to nondestructive detection of root architecture and function. We discuss the transfer of such an approach to answer root ecology questions in various tree-based landscapes. PMID:23762519</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011AGUFM.H41G1138D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011AGUFM.H41G1138D"><span>Validation of attenuation, beam blockage, and calibration estimation methods using two dual polarization X band weather <span class="hlt">radars</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Diederich, M.; Ryzhkov, A.; Simmer, C.; Mühlbauer, K.</p> <p>2011-12-01</p> <p>The amplitude a of <span class="hlt">radar</span> wave reflected by meteorological targets can be misjudged due to several factors. At X band wavelength, attenuation of the <span class="hlt">radar</span> beam by hydro meteors reduces the signal strength enough to be a significant source of error for quantitative precipitation estimation. Depending on the surrounding orography, the <span class="hlt">radar</span> beam may be partially blocked when scanning at low elevation angles, and the knowledge of the exact amount of signal loss through beam blockage becomes necessary. The phase shift between the <span class="hlt">radar</span> signals at horizontal and <span class="hlt">vertical</span> polarizations is affected by the hydrometeors that the beam travels through, but remains unaffected by variations in signal strength. This has allowed for several ways of compensating for the attenuation of the signal, and for consistency checks between these variables. In this study, we make use of several weather <span class="hlt">radars</span> and gauge network measuring in the same area to examine the effectiveness of several methods of attenuation and beam blockage corrections. The methods include consistency checks of <span class="hlt">radar</span> reflectivity and specific differential phase, calculation of beam blockage using a topography map, estimating attenuation using differential propagation phase, and the ZPHI method proposed by Testud et al. in 2000. Results show the high effectiveness of differential phase in estimating attenuation, and potential of the ZPHI method to compensate attenuation, beam blockage, and calibration errors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JGRA..123.3183Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JGRA..123.3183Y"><span>Estimation of Mesospheric Densities at Low Latitudes Using the Kunming Meteor <span class="hlt">Radar</span> Together With SABER Temperatures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yi, Wen; Xue, Xianghui; Reid, Iain M.; Younger, Joel P.; Chen, Jinsong; Chen, Tingdi; Li, Na</p> <p>2018-04-01</p> <p>Neutral mesospheric densities at a low latitude have been derived during April 2011 to December 2014 using data from the Kunming meteor <span class="hlt">radar</span> in China (25.6°N, 103.8°E). The daily mean density at 90 km was estimated using the ambipolar diffusion coefficients from the meteor <span class="hlt">radar</span> and temperatures from the Sounding of the Atmosphere using Broadband Emission Radiometry (SABER) instrument. The seasonal variations of the meteor <span class="hlt">radar</span>-derived density are consistent with the density from the Mass Spectrometer and Incoherent Scatter (MSIS) model, show a dominant annual variation, with a maximum during winter, and a minimum during summer. A simple linear model was used to separate the effects of atmospheric density and the meteor velocity on the meteor <span class="hlt">radar</span> peak detection height. We find that a 1 km/s difference in the <span class="hlt">vertical</span> meteor velocity yields a change of approximately 0.42 km in peak height. The strong correlation between the meteor <span class="hlt">radar</span> density and the velocity-corrected peak height indicates that the meteor <span class="hlt">radar</span> density estimates accurately reflect changes in neutral atmospheric density and that meteor peak detection heights, when adjusted for meteoroid velocity, can serve as a convenient tool for measuring density variations around the mesopause. A comparison of the ambipolar diffusion coefficient and peak height observed simultaneously by two co-located meteor <span class="hlt">radars</span> indicates that the relative errors of the daily mean ambipolar diffusion coefficient and peak height should be less than 5% and 6%, respectively, and that the absolute error of the peak height is less than 0.2 km.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120015537','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120015537"><span>Integrating Wind Profiling <span class="hlt">Radars</span> and Radiosonde Observations with Model <span class="hlt">Point</span> Data to Develop a Decision Support Tool to Assess Upper-Level Winds for Space Launch</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bauman, William H., III; Flinn, Clay</p> <p>2013-01-01</p> <p>On the day-of-launch, the 45th Weather Squadron (45 WS) Launch Weather Officers (LWOs) monitor the upper-level winds for their launch customers to include NASA's Launch Services Program and NASA's Ground Systems Development and Operations Program. They currently do not have the capability to display and overlay profiles of upper-level observations and numerical weather prediction model forecasts. The LWOs requested the Applied Meteorology Unit (AMU) develop a tool in the form of a graphical user interface (GUI) that will allow them to plot upper-level wind speed and direction observations from the Kennedy Space Center (KSC) 50 MHz tropospheric wind profiling <span class="hlt">radar</span>, KSC Shuttle Landing Facility 915 MHz boundary layer wind profiling <span class="hlt">radar</span> and Cape Canaveral Air Force Station (CCAFS) Automated Meteorological Processing System (AMPS) radiosondes, and then overlay forecast wind profiles from the model <span class="hlt">point</span> data including the North American Mesoscale (NAM) model, Rapid Refresh (RAP) model and Global Forecast System (GFS) model to assess the performance of these models. The AMU developed an Excel-based tool that provides an objective method for the LWOs to compare the model-forecast upper-level winds to the KSC wind profiling <span class="hlt">radars</span> and CCAFS AMPS observations to assess the model potential to accurately forecast changes in the upperlevel profile through the launch count. The AMU wrote Excel Visual Basic for Applications (VBA) scripts to automatically retrieve model <span class="hlt">point</span> data for CCAFS (XMR) from the Iowa State University Archive Data Server (http://mtarchive.qeol.iastate.edu) and the 50 MHz, 915 MHz and AMPS observations from the NASA/KSC Spaceport Weather Data Archive web site (http://trmm.ksc.nasa.gov). The AMU then developed code in Excel VBA to automatically ingest and format the observations and model <span class="hlt">point</span> data in Excel to ready the data for generating Excel charts for the LWO's. The resulting charts allow the LWOs to independently initialize the three models 0</p> </li> <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 <span class="hlt">radar</span> observations: A guide for Magellan <span class="hlt">radar</span>-image 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 <span class="hlt">radar</span> images 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 <span class="hlt">radar</span> and shuttle-imaging-<span class="hlt">radar</span> images. Analogies are drawn for the potential interpretation of <span class="hlt">radar</span> images 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 imaging geometry, spatial resolution, and wavelength of the imaging <span class="hlt">radar</span> 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 observed in images. <span class="hlt">Radar</span> responses that are governed by small-scale surface roughness may serve to distinguish flow types, but do not provide unambiguous information. Imaging 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 <span class="hlt">radar</span> wavelength. With a single look angle, conditions that enable shallow subsurface imaging to occur do not provide the information necessary to determine whether the <span class="hlt">radar</span> has recorded surface or subsurface features. The topographic linearity of many tectonic landforms is enhanced on images 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('https://ntrs.nasa.gov/search.jsp?R=20150008578&hterms=interpolation+processing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dinterpolation%2Bprocessing','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20150008578&hterms=interpolation+processing&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dinterpolation%2Bprocessing"><span>Spotlight-Mode Synthetic Aperture <span class="hlt">Radar</span> Processing for High-Resolution Lunar Mapping</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Harcke, Leif; Weintraub, Lawrence; Yun, Sang-Ho; Dickinson, Richard; Gurrola, Eric; Hensley, Scott; Marechal, Nicholas</p> <p>2010-01-01</p> <p>During the 2008-2009 year, the Goldstone Solar System <span class="hlt">Radar</span> was upgraded to support <span class="hlt">radar</span> mapping of the lunar poles at 4 m resolution. The finer resolution of the new system and the accompanying migration through resolution cells called for spotlight, rather than delay-Doppler, imaging techniques. A new pre-processing system supports fast-time Doppler removal and motion compensation to a <span class="hlt">point</span>. Two spotlight imaging techniques which compensate for phase errors due to i) out of focus-plane motion of the <span class="hlt">radar</span> and ii) local topography, have been implemented and tested. One is based on the polar format algorithm followed by a unique autofocus technique, the other is a full bistatic time-domain backprojection technique. The processing system yields imagery of the specified resolution. Products enabled by this new system include topographic mapping through <span class="hlt">radar</span> interferometry, and change detection techniques (amplitude and coherent change) for geolocation of the NASA LCROSS mission impact site.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940033279','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940033279"><span>NASA ER-2 Doppler <span class="hlt">radar</span> reflectivity calibration for the CAMEX project</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Caylor, I. J.; Heymsfield, G. M.; Bidwell, S. W.; Ameen, S.</p> <p>1994-01-01</p> <p>The NASA ER-2 Doppler <span class="hlt">radar</span> (EDOP) was flown aboard the ER-2 high-altitude aircraft in September and October 1993 for the Convection and Moisture Experiment. During these flights, the first reliable reflectivity observations were performed with the EDOP instrument. This report details the procedure used to convert real-time engineering data into calibrated <span class="hlt">radar</span> reflectivity. Application of the calibration results produces good agreement between the EDOP nadir <span class="hlt">pointing</span> reflectivity and ground truth provided by a National Weather Service WSR-88D <span class="hlt">radar</span>. The rms deviation between WSR-88D and EDOP is 6.9 dB, while measurements of the ocean surface backscatter coefficient are less than 3 dB from reported scatterometer coefficients. After an initial 30-minute period required for the instrument to reach thermal equilibrium, the <span class="hlt">radar</span> is stable to better than 0.25 dB during flight. The range performance of EDOP shows excellent agreement with aircraft altimeter and meteorological sounding data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19980227532&hterms=Hawaii+Kilauea+volcano&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DHawaii%2BKilauea%2Bvolcano','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19980227532&hterms=Hawaii+Kilauea+volcano&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DHawaii%2BKilauea%2Bvolcano"><span>Surface Deformation and Coherence Measurements of Kilauea Volcano, Hawaii, from SIR-C <span class="hlt">Radar</span> Interferometry</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rosen, P. A.; Hensley, S.; Zebker, H. A.; Webb, F. H.; Fielding, E. J.</p> <p>1996-01-01</p> <p>The shuttle imaging <span class="hlt">radar</span> C/X synthetic aperture <span class="hlt">radar</span> (SIR-C/X-SAR) <span class="hlt">radar</span> on board the space shuttle Endeavor imaged Kilauea Volcano, Hawaii, in April and October 1994 for the purpose of measuring active surface deformation by the methods of repeat-pass differential <span class="hlt">radar</span> interferometry. Observations at 24 cm (L band) and 5.6 cm (C band) wavelengths were reduced to interferograms showing apparent surface deformation over the 6-month interval and over a succession of 1-day intervals in October. A statistically significant local phase signature in the 6-month interferogram is coincident with the Pu'u O'o lava vent. Interpreted as deformation, the signal implies centimeter-scale deflation in an area several kilometers wide surrounding the vent. Peak deflation is roughly 14 cm if the deformation is purely <span class="hlt">vertical</span>, centered southward of the Pu'u O'o caldera. Delays in the <span class="hlt">radar</span> signal phase induced by atmospheric refractivity anomalies introduce spurious apparent deformation signatures, at the level of 12 cm peak-to-peak in the <span class="hlt">radar</span> line-of-sight direction. Though the phase observations are suggestive of the wide-area deformation measured by Global Positioning System (GPS) methods, the atmospheric effects are large enough to limit the interpretation of the result. It is difficult to characterize centimeter-scale deformations spatially distributed over tens of kilometers using differential interferometry without supporting simultaneous, spatially distributed measurements of reactivity along the <span class="hlt">radar</span> line of sight. Studies of the interferometric correlation of images acquired at different times show that L band is far superior to C band in the vegetated areas, even when the observations are separated by only 1 day. These results imply longer wavelength instruments are more appropriate for studying surfaces by repeat-pass observations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A21I3155E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A21I3155E"><span>Application of a Snow Growth Model to <span class="hlt">Radar</span> Remote Sensing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Erfani, E.; Mitchell, D. L.</p> <p>2014-12-01</p> <p>Microphysical growth processes of diffusion, aggregation and riming are incorporated analytically in a steady-state snow growth model (SGM) to solve the zeroth- and second- moment conservation equations with respect to mass. The SGM is initiated by <span class="hlt">radar</span> reflectivity (Zw), supersaturation, temperature, and a <span class="hlt">vertical</span> profile of the liquid water content (LWC), and it uses a gamma size distribution (SD) to predict the <span class="hlt">vertical</span> evolution of size spectra. Aggregation seems to play an important role in the evolution of snowfall rates and the snowfall rates produced by aggregation, diffusion and riming are considerably greater than those produced by diffusion and riming alone, demonstrating the strong interaction between aggregation and riming. The impact of ice particle shape on particle growth rates and fall speeds is represented in the SGM in terms of ice particle mass-dimension (m-D) power laws (m = αDβ). These growth rates are qualitatively consistent with empirical growth rates, with slower (faster) growth rates predicted for higher (lower) β values. In most models, β is treated constant for a given ice particle habit, but it is well known that β is larger for the smaller crystals. Our recent work quantitatively calculates β and α for cirrus clouds as a function of D where the m-D expression is a second-order polynomial in log-log space. By adapting this method to the SGM, the ice particle growth rates and fall speeds are predicted more accurately. Moreover, the size spectra predicted by the SGM are in good agreement with those from aircraft measurements during Lagrangian spiral descents through frontal clouds, indicating the successful modeling of microphysical processes. Since the lowest Zw over complex topography is often significantly above cloud base, the precipitation is often underestimated by <span class="hlt">radar</span> quantitative precipitation estimates (QPE). Our SGM is capable of being initialized with Zw at the lowest reliable <span class="hlt">radar</span> echo and consequently improves</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://images.nasa.gov/#/details-PIA01778.html','SCIGOVIMAGE-NASA'); return false;" href="https://images.nasa.gov/#/details-PIA01778.html"><span>Space <span class="hlt">Radar</span> Image of Patagonian Ice Fields</span></a></p> <p><a target="_blank" href="https://images.nasa.gov/">NASA Image and Video Library</a></p> <p></p> <p>1999-04-15</p> <p>This pair of images illustrates the ability of multi-parameter <span class="hlt">radar</span> imaging sensors such as the Spaceborne Imaging <span class="hlt">Radar</span>-C/X-band Synthetic Aperture <span class="hlt">radar</span> to detect climate-related changes on the Patagonian ice fields in the Andes Mountains of Chile and Argentina. The images show nearly the same area of the south Patagonian ice field as it was imaged during two space shuttle flights in 1994 that were conducted five-and-a-half months apart. The images, centered at 49.0 degrees south latitude and 73.5degrees west longitude, include several large outlet glaciers. The images were acquired by SIR-C/X-SAR on board the space shuttle Endeavour during April and October 1994. The top image was acquired on April 14, 1994, at 10:46 p.m. local time, while the bottom image was acquired on October 5,1994, at 10:57 p.m. local time. Both were acquired during the 77th orbit of the space shuttle. The area shown is approximately 100 kilometers by 58 kilometers (62 miles by 36 miles) with north toward the upper right. The colors in the images were obtained using the following <span class="hlt">radar</span> channels: red represents the C-band (horizontally transmitted and received); green represents the L-band (horizontally transmitted and received); blue represents the L-band (horizontally transmitted and <span class="hlt">vertically</span> received). The overall dark tone of the colors in the central portion of the April image indicates that the interior of the ice field is covered with thick wet snow. The outlet glaciers, consisting of rough bare ice, are the brightly colored yellow and purple lobes which terminate at calving fronts into the dark waters of lakes and fiords. During the second mission the temperatures were colder and the corresponding change in snow and ice conditions is readily apparent by comparing the images. The interior of the ice field is brighter because of increased <span class="hlt">radar</span> return from the dryer snow. The distinct green/orange boundary on the ice field indicates an abrupt change in the structure of the snowcap</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090008689','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090008689"><span>Temporal Stability of Surface Roughness Effects on <span class="hlt">Radar</span> Based Soil Moisture Retrieval During the Corn Growth Cycle</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Joseph, A.T.; Lang, R.; O'Neill, P.E.; van der Velde, R.; Gish, T.</p> <p>2008-01-01</p> <p> uncertainty depending on the sensing configuration. The effects of surface roughness variations are found to be smallest for observations acquired at a view angle of 55 degrees and HH polarization. A possible explanation for this result is that at 55 degrees and HH polarization the effect of <span class="hlt">vertical</span> surface height changes on the observed <span class="hlt">radar</span> response are limited because the microwaves travel parallel to the incident plane and as a result will not interact directly with <span class="hlt">vertically</span> oriented soil structures.</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>