Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,682)

Search Parameters:
Keywords = radar remote sensing

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
20 pages, 3512 KiB  
Article
Enhanced Polarimetric Radar Vegetation Index and Integration with Optical Index for Biomass Estimation in Grazing Lands Across the Contiguous United States
by Jisung Geba Chang, Simon Kraatz, Martha Anderson and Feng Gao
Remote Sens. 2024, 16(23), 4476; https://doi.org/10.3390/rs16234476 - 28 Nov 2024
Abstract
Grazing lands are crucial for agricultural productivity, ecological stability, and carbon sequestration, underscoring the importance of monitoring vegetation biomass for the effective management of these ecosystems. Remote sensing data, including optical vegetation indices (VIs) like the Normalized Difference Vegetation Index (NDVI), are widely [...] Read more.
Grazing lands are crucial for agricultural productivity, ecological stability, and carbon sequestration, underscoring the importance of monitoring vegetation biomass for the effective management of these ecosystems. Remote sensing data, including optical vegetation indices (VIs) like the Normalized Difference Vegetation Index (NDVI), are widely used to monitor vegetation dynamics due to their simplicity and high sensitivity. In contrast, radar-based VIs, such as the Polarimetric Radar Vegetation Index (PRVI), offer additional advantages, including all-weather imaging capabilities, a wider saturation range, and sensitivity to the vegetation structure information. This study introduces an enhanced form of the PRVI, termed the Normalized PRVI (NPRVI), which is calibrated to a 0 to 1 range, constraining the minimum value to reduce the background effects. The calibration and range factor were derived from statistical analysis of PRVI components across vegetated regions in the Contiguous United States (CONUS), using dual-polarization C-band Sentinel-1 and L-band ALOS-PALSAR data on the Google Earth Engine (GEE) platform. Machine learning models using NPRVI and NDVI demonstrated their complementarity with annual herbaceous biomass data from the Rangeland Analysis Platform. The results showed that the Random Forest Model outperformed the other machine learning models tested, achieving R2 ≈ 0.51 and MAE ≈ 498 kg/ha (relative MAE ≈ 32.1%). Integrating NPRVI with NDVI improved biomass estimation accuracy by approximately 10% compared to using NDVI alone, highlighting the added value of incorporating radar-based vegetation indices. NPRVI may enhance the monitoring of grazing lands with relatively low biomass compared to other vegetation types, while also demonstrating applicability across a broad range of biomass levels and in diverse vegetation covers. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
25 pages, 41258 KiB  
Article
The Deformation Monitoring Capability of Fucheng-1 Time-Series InSAR
by Zhouhang Wu, Wenjun Zhang, Jialun Cai, Hongyao Xiang, Jing Fan and Xiaomeng Wang
Sensors 2024, 24(23), 7604; https://doi.org/10.3390/s24237604 - 28 Nov 2024
Abstract
The Fucheng-1 (FC-1) satellite has successfully transitioned from its initial operational phase and is now undergoing a detailed performance assessment for time-series deformation monitoring. This study evaluates the surface deformation monitoring capabilities of the newly launched FC-1 satellite using the interferometric synthetic aperture [...] Read more.
The Fucheng-1 (FC-1) satellite has successfully transitioned from its initial operational phase and is now undergoing a detailed performance assessment for time-series deformation monitoring. This study evaluates the surface deformation monitoring capabilities of the newly launched FC-1 satellite using the interferometric synthetic aperture radar (InSAR) technique, particularly in urban applications. By analyzing the observation data from 20 FC-1 scenes and 20 Sentinel-1 scenes, deformation velocity maps of a university in Mianyang city were obtained using persistent scatterer interferometry (PSI) and distributed scatterer interferometry (DSI) techniques. The results show that thanks to the high resolution of 3 × 3 m of the FC-1 satellite, significantly more PS points and DS points were detected than those detected by Sentinel-1, by 13.4 times and 17.9 times, respectively. The distribution of the major deformation areas detected by both satellites in the velocity maps is generally consistent. FC-1 performs better than Sentinel-1 in monitoring densely structured and vegetation-covered areas. Its deformation monitoring capability at the millimeter level was further validated through comparison with leveling measurements, with average errors and root mean square errors of 1.761 mm and 2.172 mm, respectively. Its high-resolution and high-precision interferometry capabilities make it particularly promising in the commercial remote sensing market. Full article
(This article belongs to the Special Issue Recent Advances in Synthetic Aperture Radar (SAR) Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Coverage areas of Sentinel-1 (purple) and FC-1 (brown), study area location marked by a five-pointed star, and COPDEM topographic map. (<b>b</b>) Google Maps image of the study area.</p>
Full article ">Figure 2
<p>Flow chart of DSI and PSI.</p>
Full article ">Figure 3
<p>(<b>a</b>) Spatio-temporal baseline map of FC-1 single master image. (<b>b</b>) Spatio-temporal baseline map of Sentinel-1 single master image.</p>
Full article ">Figure 4
<p>(<b>a</b>,<b>b</b>) Vertical deformation velocity maps from FC-1 using the DSI and PSI methods. (<b>c</b>,<b>d</b>) Vertical deformation velocity maps from Sentinel-1 using the DSI and PSI methods. (<b>e</b>) Drone orthophoto of the reference point.</p>
Full article ">Figure 5
<p>(<b>a</b>,<b>b</b>) Histograms of deformation velocity from FC-1 using the DSI and PSI methods. (<b>c</b>,<b>d</b>) Histograms of deformation velocity from Sentinel-1 using the DSI and PSI methods.</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic diagram of the research area on Google Earth. (<b>b</b>,<b>c</b>) Deformation rate maps of region R1 obtained by FC-1 and Sentinel-1 using the PSI method, with a drone image as the base map. (<b>d</b>–<b>g</b>) Deformation rate maps of regions R2 and R3 obtained by FC-1 and Sentinel-1 using the PSI method, with Google Earth as the base map. (<b>h</b>–<b>k</b>) Deformation rate maps of regions R4 and R5 obtained by FC-1 and Sentinel-1 using the DSI method, with Google Earth or a drone image as the base map.</p>
Full article ">Figure 7
<p>(<b>a</b>) Deformation velocity points obtained by FC-1 using the PSI method overlaid onto a drone image. (<b>b</b>) Deformation velocity points obtained by Sentinel-1 using the PSI method overlaid onto a drone image.</p>
Full article ">Figure 8
<p>(<b>a</b>,<b>d</b>) Deformation velocity maps from FC-1 and Sentinel-1 using the PSI method, with schematic maps of ZZ1 and ZZ2 locations. (<b>b</b>,<b>c</b>) PS deformation points from FC-1 overlaid onto drone oblique images of ZZ1 and ZZ2. (<b>e</b>,<b>f</b>) PS deformation points from Sentinel-1 overlaid onto drone oblique images of ZZ1 and ZZ2.</p>
Full article ">Figure 9
<p>(<b>a</b>,<b>b</b>) Deformation velocity points obtained by FC-1 using DSI and PSI methods overlaid onto Google imagery. (<b>c</b>,<b>d</b>) Deformation velocity points obtained by Sentinel-1 using DSI and PSI methods overlaid onto Google imagery.</p>
Full article ">Figure 10
<p>Diagram of road profile location.</p>
Full article ">Figure 11
<p>(<b>a</b>,<b>b</b>) Deformation velocity profile of FC-1 under the DSI and PSI methods. (<b>c</b>,<b>d</b>) Deformation velocity profile of Sentinel-1 under the DSI and PSI methods.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>) Deformation rate profiles of Sentinel-1 and FC-1 under the DSI method. (<b>c</b>) Diagram of position of vegetation section line. (<b>d</b>) UAV 3D model of vegetation area.</p>
Full article ">Figure 13
<p>(<b>a</b>,<b>b</b>) Coherence histograms and average coherence values for the PSI method with FC-1 and Sentinel-1. (<b>c</b>,<b>d</b>) Coherence histograms and average coherence values for the DSI method with FC-1 and Sentinel-1.</p>
Full article ">Figure 14
<p>(<b>a</b>,<b>b</b>) Standard deviation maps of deformation velocity for Sentinel-1 using PSI and DSI methods. (<b>c</b>,<b>d</b>) Standard deviation maps of deformation velocity for FC-1 using PSI and DSI methods.</p>
Full article ">Figure 15
<p>(<b>a</b>) Diagram of locations of four regions A, B, C and D. (<b>b</b>–<b>e</b>) Time-series settlement maps of FC-1 and Sentinel-1 using the DSI and PSI methods in four regions.</p>
Full article ">Figure 16
<p>(<b>a</b>–<b>d</b>) Spearman’s correlation matrix heatmaps of the time-series settlement amounts obtained by FC-1 and Sentinel-1 using the DSI and PSI methods in four regions.</p>
Full article ">Figure 17
<p>(<b>a</b>–<b>d</b>) Pearson’s correlation matrix plots of the time-series subsidence values between FC-1 and Sentinel-1 using the DSI and PSI methods in four regions.</p>
Full article ">Figure 18
<p>Illustrative Google Earth map showing the locations of level points.</p>
Full article ">Figure 19
<p>The subsidence measured by FC-1 using the DSI method compared to the subsidence measured by leveling.</p>
Full article ">Figure 20
<p>The subsidence measured by FC-1 and Sentinel-1 using the DSI method compared to the subsidence measured by leveling.</p>
Full article ">
19 pages, 2837 KiB  
Article
Thermal Optimization Design for a Small Flat-Panel Synthetic Aperture Radar Satellite
by Tian Bai, Yuanbo Zhang, Lin Kong, Hongrui Ao, Jisong Yu and Lei Zhang
Aerospace 2024, 11(12), 982; https://doi.org/10.3390/aerospace11120982 - 27 Nov 2024
Viewed by 214
Abstract
This article introduces a small microwave remote sensing satellite weighing 310 kg, operating in low earth orbit (LEO). It is equipped with an X-band synthetic aperture radar (SAR) antenna, capable of a maximum imaging resolution of 0.6 m. To achieve the objectives of [...] Read more.
This article introduces a small microwave remote sensing satellite weighing 310 kg, operating in low earth orbit (LEO). It is equipped with an X-band synthetic aperture radar (SAR) antenna, capable of a maximum imaging resolution of 0.6 m. To achieve the objectives of lower cost, reduced weight, minimized power consumption, and enhanced temperature stability, an optimized thermal design method tailored for satellites has been developed, with a particular focus on SAR antennas. The thermal control method of the antenna is closely integrated with structural design, simplifying the thermal design and its assembly process, reducing the resource consumption of thermal control systems. The distribution of thermal interface material (TIM) in the antenna assembly has been carefully calculated, achieving a zero-consumption thermal design for the SAR antenna. And the temperature difference of the entire antennas when powered on and powered off would not exceed 17 °C, meeting the specification requirements. In addition, to ensure the accuracy of antenna pointing, the support plate of antennas requires stable temperature. The layout of the heaters on the board has been optimized, reducing the use of heaters by 30% while ensuring that the temperature variation of the support board remains within 5 °C. Then, an on-orbit thermal simulation analysis of the satellite was conducted to refine the design and verification. Finally, the thermal test of the SAR satellite under vacuum conditions was conducted, involving operating the high-power antenna, verifying that the peak temperature of T/RM is below 29 °C, the temperature fluctuation amplitude during a single imaging task is 10 °C, and the lowest temperature point of the support plate is 16 °C. The results of the thermal simulation and test are highly consistent, verifying the correctness and effectiveness of the thermal design. Full article
(This article belongs to the Section Astronautics & Space Science)
32 pages, 5846 KiB  
Article
Weather Radars Reveal Environmental Conditions for High Altitude Insect Movement Through the Aerosphere
by Samuel Hodges, Christopher Hassall and Ryan Neely
Remote Sens. 2024, 16(23), 4388; https://doi.org/10.3390/rs16234388 - 24 Nov 2024
Viewed by 278
Abstract
High-flying insects that exploit tropospheric winds can disperse over far greater distances in a single generation than species restricted to below-canopy flight. However, the ecological consequences of such long-range dispersal remain poorly understood. For example, high-altitude dispersal may facilitate more rapid range shifts [...] Read more.
High-flying insects that exploit tropospheric winds can disperse over far greater distances in a single generation than species restricted to below-canopy flight. However, the ecological consequences of such long-range dispersal remain poorly understood. For example, high-altitude dispersal may facilitate more rapid range shifts in these species and reduce their sensitivity to habitat fragmentation, in contrast to low-flying insects that rely more on terrestrial patch networks. Previous studies have primarily used surface-level variables with limited spatial coverage to explore dispersal timing and movement. In this study, we introduce a novel application of niche modelling to insect aeroecology by examining the relationship between a comprehensive set of atmospheric conditions and high-flying insect activity in the troposphere, as detected by weather surveillance radars (WSRs). We reveal correlations between large-scale dispersal events and atmospheric conditions, identifying key variables that influence dispersal behaviour. By incorporating high-altitude atmospheric conditions into niche models, we achieve significantly higher predictive accuracy compared with models based solely on surface-level conditions. Key predictive factors include the proportion of arable land, altitude, temperature, and relative humidity. Full article
Show Figures

Figure 1

Figure 1
<p>A high-level, generalised overview of our WSR-ENM procedure. The procedure can be considered to comprise two principal stages, radar filtering into insect presence–absence data and the pairing of 3D gridded atmospheric data with insect presence–absence [<a href="#B80-remotesensing-16-04388" class="html-bibr">80</a>]. This procedure produces Species with Data tables which can be used with a range of niche modelling approaches.</p>
Full article ">Figure 2
<p>Outcome of radar filtering applied to NXPol-1 observations (<a href="#sec2dot1dot1-remotesensing-16-04388" class="html-sec">Section 2.1.1</a>) on 10 May 2017 at ~12:00, demonstrated with plan position indicator (PPI) plots at 2.0° (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and 4.5° elevation (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) in the radar antenna. See <a href="#remotesensing-16-04388-t002" class="html-table">Table 2</a> for the list of classification rules per signal type. See <a href="#remotesensing-16-04388-f001" class="html-fig">Figure 1</a> of Lukach et al., (2022) [<a href="#B82-remotesensing-16-04388" class="html-bibr">82</a>] for a visual depiction of a PPI in real space. (<b>a</b>,<b>b</b>) Z<sub>H</sub>. (<b>c</b>,<b>d</b>) Z<sub>DR</sub>. (<b>e</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>H</mi> <mi>V</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>g</b>,<b>h</b>). Classifications based on DR (<a href="#remotesensing-16-04388-t002" class="html-table">Table 2</a>) (Kilambi et al., 2018) [<a href="#B80-remotesensing-16-04388" class="html-bibr">80</a>]. ‘Indeterminate’ scatter beyond the range of insect presence is due to a lack of Z<sub>V</sub> and consequently Z<sub>dr</sub>, resulting from attenuation in the vertical polarisation, which prevents classification by DR.</p>
Full article ">Figure 2 Cont.
<p>Outcome of radar filtering applied to NXPol-1 observations (<a href="#sec2dot1dot1-remotesensing-16-04388" class="html-sec">Section 2.1.1</a>) on 10 May 2017 at ~12:00, demonstrated with plan position indicator (PPI) plots at 2.0° (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) and 4.5° elevation (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) in the radar antenna. See <a href="#remotesensing-16-04388-t002" class="html-table">Table 2</a> for the list of classification rules per signal type. See <a href="#remotesensing-16-04388-f001" class="html-fig">Figure 1</a> of Lukach et al., (2022) [<a href="#B82-remotesensing-16-04388" class="html-bibr">82</a>] for a visual depiction of a PPI in real space. (<b>a</b>,<b>b</b>) Z<sub>H</sub>. (<b>c</b>,<b>d</b>) Z<sub>DR</sub>. (<b>e</b>,<b>f</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>ρ</mi> </mrow> <mrow> <mi>H</mi> <mi>V</mi> </mrow> </msub> </mrow> </semantics></math>. (<b>g</b>,<b>h</b>). Classifications based on DR (<a href="#remotesensing-16-04388-t002" class="html-table">Table 2</a>) (Kilambi et al., 2018) [<a href="#B80-remotesensing-16-04388" class="html-bibr">80</a>]. ‘Indeterminate’ scatter beyond the range of insect presence is due to a lack of Z<sub>V</sub> and consequently Z<sub>dr</sub>, resulting from attenuation in the vertical polarisation, which prevents classification by DR.</p>
Full article ">Figure 3
<p>Hourly mean counts of filtered insect presence (blue) over the diurnal and seasonal cycles. The grey area represents the 95% confidence interval built from daily data. Time is given in hours from midnight (00:00), in UTC. These lines were derived using the LOESS curve function in the ggplot2 package (version 3.4.4; Wickham, 2016 [<a href="#B86-remotesensing-16-04388" class="html-bibr">86</a>]) for R 4.2.2; R Core Team, 2022 [<a href="#B87-remotesensing-16-04388" class="html-bibr">87</a>].</p>
Full article ">Figure 4
<p>Boxplot comparison of models built with aerial variables (on pressure levels), terrestrial variables (surface only), and models combining the two. The plot is gridded into panels of model type (top), combining the variables used (upper text) and the subsampling factor (lower text). Each box-and-whisker represents the validation outcomes of 100 runs. Note the difference in scale between the Receiver Operator Curve (ROC, range 0–1) and True Skill Statistic (TSS, range −1–1). CTA—Classification Tree Analysis, GLM—Generalised Linear Model, RF—Random Forest.</p>
Full article ">Figure 5
<p>Biomod2 estimated variable importances for all variables used in this study, apart from vertical velocity, U-component of wind (zonal wind), potential vorticity, and divergence (which have an average contribution of &lt;0.15 for all models). Variables are sorted by averaged rank order of importance and taken from the aerial–terrestrial combined model with a subsampling factor of 0.1%. Variable importance (<span class="html-italic">y</span>-axis) is measured in 1—Pearson’s correlation coefficient (0–1); see <a href="#sec2dot2dot4-remotesensing-16-04388" class="html-sec">Section 2.2.4</a> for further details. Note the variability by model algorithm. CTA—Classification Tree Analysis, GLM—Generalised Linear Model, RF—Random Forest.</p>
Full article ">Figure 6
<p>Response curves for the top four contributing variables; curves are taken from ‘combined’ models with both aerial and terrestrial variables, for subsampling factor 0.01. The response curves are based on the model run with the best predictive skill (in terms of ROC and TSS) out of each set of 100. ‘Altitude band’ is given as a categorical variable where the number in km represents the median of the band (i.e., 1 km ± 0.5 km). CTA—Classification Tree Analysis, GLM—Generalised Linear Model, RF—Random Forest.</p>
Full article ">Figure 7
<p>Predictions of aerial habitat suitability (probability of presence) from the median predictive skill (in terms of TSS) GLM of insect activity (0.01 subsampling factor, aerial and terrestrial variables). Predictions were made using atmospheric data from 17 July 2017 at 00:00, 06:00, 12:00, and 18:00 h UTC (columns). Atmospheric data were taken from three altitude levels, 1000 m, 2000 m, and 3000 m above sea level (rows). This shows where similar atmospheric environments associated with insects occurred across the UK on this day, and how the suitable area developed over time.</p>
Full article ">
15 pages, 9270 KiB  
Communication
Effect of DEM Used for Terrain Correction on Forest Windthrow Detection Using COSMO SkyMed Data
by Michele Dalponte, Daniele Marinelli and Yady Tatiana Solano-Correa
Remote Sens. 2024, 16(22), 4309; https://doi.org/10.3390/rs16224309 - 19 Nov 2024
Viewed by 284
Abstract
Preprocessing Synthetic Aperture Radar (SAR) data is a crucial initial stage in leveraging SAR data for remote sensing applications. Terrain correction, both radiometric and geometric, and the detection of layover/shadow areas hold significant importance when SAR data are collected over mountainous regions. This [...] Read more.
Preprocessing Synthetic Aperture Radar (SAR) data is a crucial initial stage in leveraging SAR data for remote sensing applications. Terrain correction, both radiometric and geometric, and the detection of layover/shadow areas hold significant importance when SAR data are collected over mountainous regions. This study aims at investigating the impact of the Digital Elevation Model (DEM) used for terrain correction (radiometric and geometric) and for mapping layover/shadow areas on windthrow detection using COSMO SkyMed SAR images. The terrain correction was done using a radiometric and geometric terrain correction algorithm. Specifically, we evaluated five different DEMs: (i–ii) a digital terrain model and a digital surface model derived from airborne LiDAR flights; (iii) the ALOS Global Digital Surface Model; (iv) the Copernicus global DEM; and (v) the Shuttle Radar Topography Mission (SRTM) DEM. All five DEMs were resampled at 2 m and 30 m pixel spacing, obtaining a total of 10 DEMs. The terrain-corrected COSMO SkyMed SAR images were employed for windthrow detection in a forested area in the north of Italy. The findings revealed significant variations in windthrow detection across the ten corrections. The detailed LiDAR-derived terrain model (i.e., DTM at 2 m pixel spacing) emerged as the optimal choice for both pixel spacings considered. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The location of PAT in Italy and Europe (inset (<b>a</b>)), the location of the two reference sites inside the territory of PAT and the DTM of PAT (inset (<b>b</b>)), and the DTM of the two reference areas A and B (inset (<b>c</b>)).</p>
Full article ">Figure 2
<p>Architecture of the processing chain adopted in this study.</p>
Full article ">Figure 3
<p>(<b>A</b>) Difference images at 2 m pixel spacing between the local LiDAR DSM, the three global DEMs, and the local LiDAR DTM; (<b>B</b>) a zoom over a flat area (cropland); (<b>C</b>) zoom over a forest area; and (<b>D</b>) two vertical profiles of the five DEMs at 2 m pixel spacing.</p>
Full article ">Figure 4
<p>(<b>A</b>) Difference images at 30 m pixel spacing between the local LiDAR DSM, the three global DEMs, and the local LiDAR DTM; (<b>B</b>) a zoom over a flat area (cropland); (<b>C</b>) zoom over a forest area; and (<b>D</b>) two vertical profiles of the five DEMs at 30 m pixel spacing.</p>
Full article ">Figure 5
<p>Windthrow detection maps for a subset of the study area.</p>
Full article ">
18 pages, 13617 KiB  
Article
Observation and Numerical Simulation of Cross-Mountain Airflow at the Hong Kong International Airport from Range Height Indicator Scans of Radar and LIDAR
by Ying Wa Chan, Kai Wai Lo, Ping Cheung, Pak Wai Chan and Kai Kwong Lai
Atmosphere 2024, 15(11), 1391; https://doi.org/10.3390/atmos15111391 - 19 Nov 2024
Viewed by 263
Abstract
Apart from headwind changes, crosswind changes may be hazardous to aircraft operation. This paper presents two cases of recently observed crosswind changes from the range height indicator scans of ground-based remote sensing meteorological equipment, namely an X-band microwave radar and a short-range LIDAR. [...] Read more.
Apart from headwind changes, crosswind changes may be hazardous to aircraft operation. This paper presents two cases of recently observed crosswind changes from the range height indicator scans of ground-based remote sensing meteorological equipment, namely an X-band microwave radar and a short-range LIDAR. Both instruments have a range resolution down to around 30 m, allowing the study of fine-scale details of the vertical profiles of cross-mountain airflow at the Hong Kong International Airport. Rapidly evolving winds have been observed by the equipment in tropical cyclone situations, revealing high levels of turbulence and vertically propagating waves. The eddy dissipation rate derived from radar spectrum width indicated severe turbulence, with values exceeding 0.5 m2/3 s−1. In order to study the feasibility of predicting such disturbed airflow, a mesoscale meteorological model and a computational fluid dynamics model with high spatial resolution are used in this paper. It is found that the mesoscale meteorological model alone is sufficient to capture some rapidly evolving airflow features, including the turbulence level, the waves, and the rapidly changing wind speeds. However, the presence of reverse flow could only be reproduced with the use of a building-resolving computational fluid dynamics model. This paper aims at providing a reference for airports to consider the feasibility of performing high-resolution numerical simulations of rapidly evolving airflow to alert the pilots in advance for airports in complex terrains and the setup of buildings. Full article
(This article belongs to the Special Issue Tropical Cyclones: Observations and Prediction (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The surface synoptic chart at 0200 Hong Kong time on 2 September 2023 and (<b>b</b>) the surface synoptic chart at 0200 Hong Kong time on 7 September 2024.</p>
Full article ">Figure 2
<p>The locations of the X-band dual polarisation phased array weather radar (PAWR) and the wind profiler at Sha Lo Wan (SLW), as well as the short-range LIDAR at the Government Flying Service (GFS) headquarters at the Hong Kong International Airport (HKIA).</p>
Full article ">Figure 3
<p>(<b>a</b>) The RHI spectral width scan and (<b>b</b>) the associated cross-section of the eddy dissipation rate (EDR) as well as (<b>c</b>) the three-dimensional wind fields (left: horizontal wind field at height of around 34 m above sea level. Right: wind field projected on the cross-sectional plane A–B on the left panel) obtained/retrieved using the SLW radar data at around 0255 Hong Kong time on 2 September 2023.</p>
Full article ">Figure 3 Cont.
<p>(<b>a</b>) The RHI spectral width scan and (<b>b</b>) the associated cross-section of the eddy dissipation rate (EDR) as well as (<b>c</b>) the three-dimensional wind fields (left: horizontal wind field at height of around 34 m above sea level. Right: wind field projected on the cross-sectional plane A–B on the left panel) obtained/retrieved using the SLW radar data at around 0255 Hong Kong time on 2 September 2023.</p>
Full article ">Figure 4
<p>(<b>a</b>) The RHI spectral width scan and (<b>b</b>) the associated cross-section of the eddy dissipation rate (EDR) as well as (<b>c</b>) the three-dimensional wind fields (left: horizontal wind field at height of around 34 m above sea level. Right: wind field projected on the cross-sectional plane A–B on the left panel) obtained/retrieved using the SLW radar data at around 0305 Hong Kong time on 2 September 2023.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) The RHI spectral width scan and (<b>b</b>) the associated cross-section of the eddy dissipation rate (EDR) as well as (<b>c</b>) the three-dimensional wind fields (left: horizontal wind field at height of around 34 m above sea level. Right: wind field projected on the cross-sectional plane A–B on the left panel) obtained/retrieved using the SLW radar data at around 0305 Hong Kong time on 2 September 2023.</p>
Full article ">Figure 5
<p>(<b>a</b>) The 1st, 2nd, and 3rd nested domains of the RAMS simulation. (<b>b</b>) The 3rd, 4th, and 5th nested domains of the RAMS simulation.</p>
Full article ">Figure 6
<p>(<b>a</b>) Observations of horizontal wind profiles from the SLW wind profiler from 18:00 Hong Kong time on 1 September 2023 to 06:00 Hong Kong Time on 2 September 2023 (+8 UTC). (<b>b</b>) Simulated horizontal wind profiles (wind barbs) and vertical velocities (background colour) from the RAMS on 1 September 2023 from 15UTC to 22UTC.</p>
Full article ">Figure 7
<p>Simulation of RHI scans of (<b>a</b>) the EDR and (<b>b</b>) wind field from the RAMS on 18:54:10 UTC, 1 September 2023.</p>
Full article ">Figure 8
<p>(<b>a</b>) The domain of the PALM simulation denoted by a red rectangle where the boundary is the 5th nested domain of the RAMS simulation. The blue line indicates the location of the RHI scan of the GFS LIDAR. (<b>b</b>) The PALM simulation domain with building heights indicated by the colour bar. The blue cross symbol indicates the location of the GFS LIDAR, and the blue line shows the location of its RHI scan.</p>
Full article ">Figure 9
<p>(<b>a</b>) RHI scans of radial wind velocity from the GFS LIDAR at 00:40:18 Hong Kong time (+8 UTC) on 7 September 2024. (<b>b</b>) RHI scans of radial wind velocity from the GFS LIDAR at 00:42:36 Hong Kong time (+8 UTC) on 7 September 2024.</p>
Full article ">Figure 10
<p>(<b>a</b>) The RAMS simulation for RHI scans of radial wind velocity from the GFS LIDAR at 16:40:10 UTC on 6 September 2024. (<b>b</b>) The RAMS simulation for RHI scans of radial wind velocity from the GFS LIDAR at 16:43:40 UTC on 6 September 2024.</p>
Full article ">Figure 11
<p>The RAMS simulation of radial wind velocity from the GFS LIDAR with an extended range at 16:40:10 UTC on 6 September 2024.</p>
Full article ">Figure 12
<p>The PALM simulation for RHI scans of radial wind velocity from the GFS LIDAR at four different instances, namely (<b>a</b>) 16:40UTC, (<b>b</b>) 16:42 UTC, (<b>c</b>) 16:43 UTC, and (<b>d</b>) 16:45 UTC on 6 September 2024.</p>
Full article ">
19 pages, 21578 KiB  
Article
A Gradual Adversarial Training Method for Semantic Segmentation
by Yinkai Zan, Pingping Lu and Tingyu Meng
Remote Sens. 2024, 16(22), 4277; https://doi.org/10.3390/rs16224277 - 16 Nov 2024
Viewed by 577
Abstract
Deep neural networks (DNNs) have achieved great success in various computer vision tasks. However, they are susceptible to artificially designed adversarial perturbations, which limit their deployment in security-critical applications. In this paper, we propose a gradual adversarial training (GAT) method for remote sensing [...] Read more.
Deep neural networks (DNNs) have achieved great success in various computer vision tasks. However, they are susceptible to artificially designed adversarial perturbations, which limit their deployment in security-critical applications. In this paper, we propose a gradual adversarial training (GAT) method for remote sensing image segmentation. Our method incorporates a domain-adaptive mechanism that dynamically modulates input data, effectively reducing adversarial perturbations. GAT not only improves segmentation accuracy on clean images but also significantly enhances robustness against adversarial attacks, all without necessitating changes to the network architecture. The experimental results demonstrate that GAT consistently outperforms conventional standard adversarial training (SAT), showing increased resilience to adversarial attacks of varying intensities on both optical and Synthetic Aperture Radar (SAR) images. Compared to the SAT defense method, GAT achieves a notable defense performance improvement of 1% to 12%. Full article
(This article belongs to the Special Issue SAR-Based Signal Processing and Target Recognition (Second Edition))
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of no defense (ND), active defense (AD), and passive defense (PD). Active defense is robust by adjusting the network, while passive defense is defended by preprocessing operations outside the network.</p>
Full article ">Figure 2
<p>Schematic diagram of the manifold hypothesis. Natural images lie on a low-dimensional manifold, while images with adversarial perturbations added to them lie outside the low-dimensional manifold.</p>
Full article ">Figure 3
<p>GAT training flowchart. The GAT method proposed in this paper can be divided into two modules: intermediate domain data generation and standard DNN training process. The intermediate domain data generation module generates intermediate domain data based on clean images and uses them as input to the latter. The standard DNN training process trains the model based on the input data and provides the former with parameters for the generation of adversarial perturbations.</p>
Full article ">Figure 4
<p>A presentation of data from San Francisco. (<b>a</b>) Pauli decomposition result and (<b>b</b>) Ground truth. Blue is water, green is vegetation, red is high-density urban, yellow is low-density urban, and purple is development areas, black is unlabeled background.</p>
Full article ">Figure 5
<p>A presentation of data from Vaihingen. The blue box area in the left figure is the selected typical area with 4 different types of ground objects, which are used for analysis in the subsequent presentation of the experimental results. (<b>a</b>) ISPRS-Vaihingen dataset example and (<b>b</b>) Ground truth. Blue is buildings, light blue is low vegetation, green is trees, yellow is cars, red is the background, and white is imperviouos surfaces.</p>
Full article ">Figure 6
<p>Metric curves of segmentation results on the SF-RS2 dataset facing adversarial attacks with different attack intensities. The first row shows the Acc evaluation index curve, and the second row shows the F1 score evaluation index curve. From the first column to the fourth column, the attack algorithms using FGSM, DAG, PGD, and segPGD are shown. The horizontal axis of each graph is the attack intensity, which ranges from 0.00 to 0.01, and the vertical axis is the evaluation index.</p>
Full article ">Figure 7
<p>SF-RS2 dataset segmentation results in the face of FGSM, DAG, PGD, and segPGD attacks. The attack intensity ranges from 0 to 0.01. For each adversarial attack algorithm, the segmentation results of no defense, SAT, and GAT are compared in turn. Taking the yellow circle as an example, the feature type of the area is high-density city, and it can clearly be seen that the segmentation accuracy of GAT is better than that of SAT and no defense. The dotted gray lines correspond to 90% accuracy and the dotted yellow lines correspond to 75% accuracy. When accuracy is reduced to the same level, the GAT method can withstand a stronger attack intensity.</p>
Full article ">Figure 8
<p>Metric curves of segmentation results on the ISPRS-Vaihingen dataset facing adversarial attacks with different attack intensities. The first row shows the Acc evaluation index curve, and the second row shows the F1 score evaluation index curve. From the first column to the fourth column, the attack algorithms using FGSM, DAG, PGD, and segPGD are shown. The horizontal axis of each graph is the attack intensity, which ranges from 0.00 to 0.0157, and the vertical axis is the evaluation index.</p>
Full article ">Figure 9
<p>ISPRS-Vaihingen dataset segmentation results in the face of FGSM, DAG, PGD, and segPGD attacks. The attack intensity ranges from 0 to 0.0196. For each adversarial attack algorithm, the segmentation results of no defense, SAT, and GAT are compared in turn. Taking the red box area as an example, the feature type of this area is building, and it can clearly be seen that the segmentation accuracy of GAT is better than that of SAT and no defense. The dotted gray lines correspond to 50% accuracy. When accuracy is reduced to the same level, the GAT method can withstand a stronger attack intensity.</p>
Full article ">
17 pages, 8238 KiB  
Article
Mapping Polylepis Forest Using Sentinel, PlanetScope Images, and Topographical Features with Machine Learning
by Diego Pacheco-Prado, Esteban Bravo-López and Luis Á. Ruiz
Remote Sens. 2024, 16(22), 4271; https://doi.org/10.3390/rs16224271 - 16 Nov 2024
Viewed by 407
Abstract
Globally, there is a significant trend in the loss of native forests, including those of the Polylepis genus, which are essential for soil conservation across the Andes Mountain range. These forests play a critical role in regulating water flow, promoting soil regeneration, and [...] Read more.
Globally, there is a significant trend in the loss of native forests, including those of the Polylepis genus, which are essential for soil conservation across the Andes Mountain range. These forests play a critical role in regulating water flow, promoting soil regeneration, and retaining essential nutrients and sediments, thereby contributing to the soil conservation of the region. In Ecuador, these forests are often fragmented and isolated in areas of high cloud cover, making it difficult to use remote sensing and spectral vegetation indices to detect this forest species. This study developed twelve scenarios using medium- and high-resolution satellite data, integrating datasets such as Sentinel-2 and PlanetScope (optical), Sentinel-1 (radar), and the Sigtierras project topographic data. The scenarios were categorized into two groups: SC1–SC6, combining 5 m resolution data, and SC7–SC12, combining 10 m resolution data. Additionally, each scenario was tested with two target types: multiclass (distinguishing Polylepis stands, native forest, Pine, Shrub vegetation, and other classes) and binary (distinguishing Polylepis from non-Polylepis). The Recursive Feature Elimination technique was employed to identify the most effective variables for each scenario. This process reduced the number of variables by selecting those with high importance according to a Random Forest model, using accuracy and Kappa values as criteria. Finally, the scenario that presented the highest reliability was SC10 (Sentinel-2 and Topography) with a pixel size of 10 m in a multiclass target, achieving an accuracy of 0.91 and a Kappa coefficient of 0.80. For the Polylepis class, the User Accuracy and Producer Accuracy were 0.90 and 0.89, respectively. The findings confirm that, despite the limited area of the Polylepis stands, integrating topographic and spectral variables at a 10 m pixel resolution improves detection accuracy. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sentinel-2A image mosaic covering tiles 17MPS, 17MPT, 17MQS, and 17MQT (24 August 2020) of the study area and location of <span class="html-italic">Polylepis</span> stands.</p>
Full article ">Figure 2
<p><span class="html-italic">Polylepis reticulata</span> tree (<b>a</b>) and stand (<b>b</b>) in Toreadora lagoon zone.</p>
Full article ">Figure 3
<p>Workflow diagram.</p>
Full article ">Figure 4
<p>Behavior of Accuracy (blue line), Kappa (orange line), UA (gray line), and PA (yellow line) metrics during the RFE process in the SC6 scenario. Peak values are obtained with 11 variables.</p>
Full article ">Figure 5
<p>Variables that were repeated (occurrences) in the best eight scenarios. S2 (Sentinel-2), S1 (Sentinel-1), TOPO (Topography), and PS (PlanetScope).</p>
Full article ">Figure 6
<p>Detailed view of the eight best classification scenarios in the Toreadora Lagoon region.</p>
Full article ">Figure 7
<p>SC10 multiclass map was selected as the best scenario. Known regions such as Toreadora Lagoon (purple polygon) and Llaviuco Lagoon (red polygon) helped to visually interpret the quality of the classification models. Other land uses and coverages in the area are shown in white color.</p>
Full article ">Figure 8
<p>Feature importance of SC10 multiclass scenario.</p>
Full article ">
21 pages, 6345 KiB  
Article
Integration of Optical and Synthetic Aperture Radar Data with Different Synthetic Aperture Radar Image Processing Techniques and Development Stages to Improve Soybean Yield Prediction
by Isabella A. Cunha, Gustavo M. M. Baptista, Victor Hugo R. Prudente, Derlei D. Melo and Lucas R. Amaral
Agriculture 2024, 14(11), 2032; https://doi.org/10.3390/agriculture14112032 - 12 Nov 2024
Viewed by 645
Abstract
Predicting crop yield throughout its development cycle is crucial for planning storage, processing, and distribution. Optical remote sensing has been used for yield prediction but has limitations, such as cloud interference and only capturing canopy-level data. Synthetic Aperture Radar (SAR) complements optical data [...] Read more.
Predicting crop yield throughout its development cycle is crucial for planning storage, processing, and distribution. Optical remote sensing has been used for yield prediction but has limitations, such as cloud interference and only capturing canopy-level data. Synthetic Aperture Radar (SAR) complements optical data by capturing information even in cloudy conditions and providing additional plant insights. This study aimed to explore the correlation of SAR variables with soybean yield at different crop stages, testing if SAR data enhances predictions compared to optical data alone. Data from three growing seasons were collected from an area of 106 hectares, using eight SAR variables (Alpha, Entropy, DPSVI, RFDI, Pol, RVI, VH, and VV) and four speckle noise filters. The Random Forest algorithm was applied, combining SAR variables with the EVI optical index. Although none of the SAR variables showed strong correlations with yield (r < |0.35|), predictions improved when SAR data were included. The best performance was achieved using DPSVI with the Boxcar filter, combined with EVI during the maturation stage (with EVI:RMSE = 0.43, 0.49, and 0.60, respectively, for each season; while EVI + DPSVI:RMSE = 0.39, 0.49, and 0.42). Despite improving predictions, the computational demands of SAR processing must be considered, especially when optical data are limited due to cloud cover. Full article
(This article belongs to the Special Issue Applications of Remote Sensing in Agricultural Soil and Crop Mapping)
Show Figures

Figure 1

Figure 1
<p>Experimental area with field boundaries marked in red and soybean yield data points in each harvest.</p>
Full article ">Figure 2
<p>Temporal profiles of SAR data in <span class="html-italic">VV</span> and <span class="html-italic">VH</span> backscatter coefficient (<b>a</b>) and optical data considering EVI (<b>b</b>). The red circle represents the selected image dates based on the EVI.</p>
Full article ">Figure 3
<p>SAR data workflow for obtaining (<b>a</b>) backscatter coefficients and (<b>b</b>) polarimetric decomposition.</p>
Full article ">Figure 4
<p>Prediction scenarios performed. Input data corresponding to each tested scenario (in red): (<b>a</b>) using all stages and SAR variables together, (<b>b</b>) using stages separately and all SAR variables together, (<b>c</b>) using the stage that previously performed best with the variables separated.</p>
Full article ">Figure 5
<p>Spearman correlation coefficient between SAR data and soybean yield, including harvest, growth stages, speckle noise reduction filters, and SAR variables. Significant correlations at 5%.</p>
Full article ">Figure 6
<p>R<sup>2</sup> and RMSE values of predictions for each harvest individually with all stages of image collection, using only optical data (EVI) compared to using optical data together with all SAR variables.</p>
Full article ">Figure 7
<p>DPSVI index map for distinct growth stages and soybean harvests. The highlighted area in black shows the difference in cultivar in harvest 3.</p>
Full article ">Figure 8
<p>Percentage difference in R<sup>2</sup> of predictions with EVI and adding SAR variables in models using each stage individually compared to all stages combined.</p>
Full article ">Figure 9
<p>Percentage difference in RMSE of predictions with EVI and adding SAR variables in models using each stage individually compared to all stages combined.</p>
Full article ">Figure 10
<p>R<sup>2</sup> values for predictions using all growth stages with only optical data and using optical data in conjunction with all SAR variables.</p>
Full article ">Figure 11
<p>RMSE values for predictions using all growth stages with only optical data and using optical data in conjunction with all SAR variables.</p>
Full article ">Figure 12
<p>R<sup>2</sup> values obtained for Stage 3 using scenarios with separate SAR variables in conjunction with EVI, compared to using all SAR variables combined with EVI and using only optical data (EVI).</p>
Full article ">Figure 13
<p>Visual comparison between actual yield maps and predicted yield using DPSVI in conjunction with EVI for Stage 3, using the Boxcar filter. The actual yield data were interpolated using ordinary kriging. The error maps represent the difference between the actual and predicted maps, showing positive and negative variations.</p>
Full article ">
24 pages, 2680 KiB  
Review
Remote Sensing Techniques for Assessing Snow Avalanche Formation Factors and Building Hazard Monitoring Systems
by Natalya Denissova, Serik Nurakynov, Olga Petrova, Daniker Chepashev, Gulzhan Daumova and Alena Yelisseyeva
Atmosphere 2024, 15(11), 1343; https://doi.org/10.3390/atmos15111343 - 9 Nov 2024
Viewed by 691
Abstract
Snow avalanches, one of the most severe natural hazards in mountainous regions, pose significant risks to human lives, infrastructure, and ecosystems. As climate change accelerates shifts in snowfall and temperature patterns, it is increasingly important to improve our ability to monitor and predict [...] Read more.
Snow avalanches, one of the most severe natural hazards in mountainous regions, pose significant risks to human lives, infrastructure, and ecosystems. As climate change accelerates shifts in snowfall and temperature patterns, it is increasingly important to improve our ability to monitor and predict avalanches. This review explores the use of remote sensing technologies in understanding key geomorphological, geobotanical, and meteorological factors that contribute to avalanche formation. The primary objective is to assess how remote sensing can enhance avalanche risk assessment and monitoring systems. A systematic literature review was conducted, focusing on studies published between 2010 and 2025. The analysis involved screening relevant studies on remote sensing, avalanche dynamics, and data processing techniques. Key data sources included satellite platforms such as Sentinel-1, Sentinel-2, TerraSAR-X, and Landsat-8, combined with machine learning, data fusion, and change detection algorithms to process and interpret the data. The review found that remote sensing significantly improves avalanche monitoring by providing continuous, large-scale coverage of snowpack stability and terrain features. Optical and radar imagery enable the detection of crucial parameters like snow cover, slope, and vegetation that influence avalanche risks. However, challenges such as limitations in spatial and temporal resolution and real-time monitoring were identified. Emerging technologies, including microsatellites and hyperspectral imaging, offer potential solutions to these issues. The practical implications of these findings underscore the importance of integrating remote sensing data with ground-based observations for more robust avalanche forecasting. Enhanced real-time monitoring and data fusion techniques will improve disaster management, allowing for quicker response times and more effective policymaking to mitigate risks in avalanche-prone regions. Full article
Show Figures

Figure 1

Figure 1
<p>Flow chart of the literature search strategy.</p>
Full article ">Figure 2
<p>Geographic distribution of study areas where relevant literature was found.</p>
Full article ">Figure 3
<p>Number of publications per year.</p>
Full article ">Figure 4
<p>Word cloud illustrating the frequency of terms in titles of reviewed articles.</p>
Full article ">Figure 5
<p>Clustered co-occurrence map of most relevant terms from titles of the compiled articles.</p>
Full article ">
16 pages, 32403 KiB  
Article
Integrated Analysis of Rockfalls and Floods in the Jiului Gorge, Romania: Impacts on Road and Rail Traffic
by Marian Puie and Bogdan-Andrei Mihai
Appl. Sci. 2024, 14(22), 10270; https://doi.org/10.3390/app142210270 - 8 Nov 2024
Viewed by 664
Abstract
This study examines the impact of rockfalls and floods on road and rail traffic in the Jiului Gorge, Romania, a critical transportation corridor. Using Sentinel-1 radar imagery processed through ESA SNAP and ArcGIS Pro, alongside traffic detection facilitated by YOLO models, we assessed [...] Read more.
This study examines the impact of rockfalls and floods on road and rail traffic in the Jiului Gorge, Romania, a critical transportation corridor. Using Sentinel-1 radar imagery processed through ESA SNAP and ArcGIS Pro, alongside traffic detection facilitated by YOLO models, we assessed susceptibility to both rockfalls and floods. The primary aim was to enhance public safety for traffic participants by providing accurate hazard mapping. Our study focuses on the area from Bumbești-Jiu to Petroșani, traversing the Southern Carpathians. The results demonstrate the utility of integrating remote sensing with machine learning to improve hazard management and inform more effective traffic planning. These findings contribute to safer, more resilient infrastructure in areas vulnerable to natural hazards. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area.</p>
Full article ">Figure 2
<p>Sentinel 1 GRD images of study area from descending orbit (left, 20 January 2023), from ascending orbit (middle, 20 January 2023), RGB interferogram and processing software workflow.</p>
Full article ">Figure 3
<p>ESA SNAP software workflow image samples for rockfall detection from Sentinel-1 SLC product.</p>
Full article ">Figure 4
<p>Flood map of the Jiului Gorge region, illustrating the extent and severity of flood events based on Sentinel-1 GRD images.</p>
Full article ">Figure 5
<p>Rockfall map displaying incidents along National Road 66 and surrounding slopes for specified dates.</p>
Full article ">Figure 6
<p>Rockfall susceptibility map showing areas highly susceptible to rockfall, with a notable prevalence in the upper section of the gorge.</p>
Full article ">Figure 7
<p>Rockfall susceptibility map combined with affected areas from radar images, highlighting the upper part of the gorge with high susceptibility.</p>
Full article ">Figure 8
<p>Floods susceptibility map combining DEM-derived slope classifications, land cover types, rainfall, and proximity to water bodies, showing higher susceptibility in wider parts of the gorge.</p>
Full article ">Figure 9
<p>Floods susceptibility map combined with radar-detected flood areas, illustrating increased susceptibility in the central and southern parts of the gorge.</p>
Full article ">Figure 10
<p>Train detection and recognition using YOLO models, illustrating detection from a significant distance with reduced visibility.</p>
Full article ">Figure 11
<p>Road traffic element detection with greater precision due to closer camera proximity.</p>
Full article ">Figure 12
<p>Training results from YOLOv9 model, showcasing classes obtained after training.</p>
Full article ">Figure 13
<p>Detection results including several classes, highlighting various rockfall types.</p>
Full article ">Figure 14
<p>Detection results focusing on a single class, illustrating detailed rockfall identification.</p>
Full article ">
18 pages, 982 KiB  
Review
Remote Sensing and GIS in Natural Resource Management: Comparing Tools and Emphasizing the Importance of In-Situ Data
by Sanjeev Sharma, Justin O. Beslity, Lindsey Rustad, Lacy J. Shelby, Peter T. Manos, Puskar Khanal, Andrew B. Reinmann and Churamani Khanal
Remote Sens. 2024, 16(22), 4161; https://doi.org/10.3390/rs16224161 - 8 Nov 2024
Viewed by 1263
Abstract
Remote sensing (RS) and Geographic Information Systems (GISs) provide significant opportunities for monitoring and managing natural resources across various temporal, spectral, and spatial resolutions. There is a critical need for natural resource managers to understand the expanding capabilities of image sources, analysis techniques, [...] Read more.
Remote sensing (RS) and Geographic Information Systems (GISs) provide significant opportunities for monitoring and managing natural resources across various temporal, spectral, and spatial resolutions. There is a critical need for natural resource managers to understand the expanding capabilities of image sources, analysis techniques, and in situ validation methods. This article reviews key image analysis tools in natural resource management, highlighting their unique strengths across diverse applications such as agriculture, forestry, water resources, soil management, and natural hazard monitoring. Google Earth Engine (GEE), a cloud-based platform introduced in 2010, stands out for its vast geospatial data catalog and scalability, making it ideal for global-scale analysis and algorithm development. ENVI, known for advanced multi- and hyperspectral image processing, excels in vegetation monitoring, environmental analysis, and feature extraction. ERDAS IMAGINE specializes in radar data analysis and LiDAR processing, offering robust classification and terrain analysis capabilities. Global Mapper is recognized for its versatility, supporting over 300 data formats and excelling in 3D visualization and point cloud processing, especially in UAV applications. eCognition leverages object-based image analysis (OBIA) to enhance classification accuracy by grouping pixels into meaningful objects, making it effective in environmental monitoring and urban planning. Lastly, QGIS integrates these remote sensing tools with powerful spatial analysis functions, supporting decision-making in sustainable resource management. Together, these tools when paired with in situ data provide comprehensive solutions for managing and analyzing natural resources across scales. Full article
Show Figures

Figure 1

Figure 1
<p>Articles published using different image analysis tools in different time intervals.</p>
Full article ">Figure 2
<p>Map of sites identified and included in database.</p>
Full article ">
17 pages, 2746 KiB  
Article
Deterministic Sea Wave Reconstruction and Prediction Based on Coherent S-Band Radar Using Condition Number Regularized Least Squares
by Zhongqian Hu, Zezong Chen, Chen Zhao and Xi Chen
Remote Sens. 2024, 16(22), 4147; https://doi.org/10.3390/rs16224147 - 7 Nov 2024
Viewed by 367
Abstract
Coherent S-band radar is a remote sensing observation device with high spatial-temporal resolution and can be used to achieve deterministic sea wave reconstruction and prediction (DSWRP) technology. However, coherent S-band radar can observe nonlinear details of the sea surface due to its high [...] Read more.
Coherent S-band radar is a remote sensing observation device with high spatial-temporal resolution and can be used to achieve deterministic sea wave reconstruction and prediction (DSWRP) technology. However, coherent S-band radar can observe nonlinear details of the sea surface due to its high resolution, which makes the propagation operator matrix an ill-conditioned overdetermined matrix. To solve this problem, this paper proposes a DSWRP scheme using condition number regularized least squares (CN-RLS) for coherent S-band radar. First, the space-time velocity information was obtained from the radar echo. Second, the CN-RLS method solved the phase-resolved model coefficients. Finally, the deterministic wave field was predicted according to the solved model coefficients. The proposed scheme was verified by simulation data and the real radar dataset observed by the coherent S-band wave-measuring radar onboard the ship XIANGYANGHONG-18 in the East China Sea in April 2024. The predicted wave elevation of the proposed method was compared with the wave elevation observed based on the X-band wave-measuring radar, and the root mean square error (RMSE) and correlation coefficient (CC) were 0.22 m and 0.76, respectively, which show that the proposed method could effectively implement the DSWRP technology. Full article
Show Figures

Figure 1

Figure 1
<p>Illustration of surface wave velocity measured by radar.</p>
Full article ">Figure 2
<p>The CN-RLS and L-curve RLS methods were used to estimate <math display="inline"><semantics> <mover accent="true"> <mi mathvariant="bold-italic">y</mi> <mo stretchy="false">^</mo> </mover> </semantics></math>.</p>
Full article ">Figure 3
<p>The distribution of CC depended on the <span class="html-italic">K</span> value under different condition numbers (<math display="inline"><semantics> <mrow> <mi>c</mi> <mi>o</mi> <mi>n</mi> <mi>d</mi> <mo>(</mo> <msup> <mi mathvariant="bold-italic">P</mi> <mi>T</mi> </msup> <mi mathvariant="bold">P</mi> <mo>)</mo> </mrow> </semantics></math>).</p>
Full article ">Figure 4
<p>Flow chart of DSWRP.</p>
Full article ">Figure 5
<p>Spatial−temporal velocity series: (<b>a</b>) without broken waves; (<b>b</b>) with broken waves; (<b>c</b>) velocity time series of a single range cell.</p>
Full article ">Figure 6
<p>Simulated and radar-observed wave spectrum.</p>
Full article ">Figure 7
<p>The estimated model coefficients and amplitudes: (<b>a</b>) model coefficient: <math display="inline"><semantics> <msub> <mi>a</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics></math>; (<b>b</b>) model coefficient: <math display="inline"><semantics> <msub> <mi>b</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics></math>; (<b>c</b>) amplitude: <math display="inline"><semantics> <msub> <mi>A</mi> <mrow> <mi>i</mi> <mi>j</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 8
<p>Absolute error of the DSWRP: (<b>a</b>) spatial-temporal absolute error based on the L-curve RLS method; (<b>b</b>) spatial-temporal absolute error based on the proposed method; (<b>c</b>) sea wave surface elevation at the 140th range cell; (<b>d</b>) absolute error at the 140th range cell.</p>
Full article ">Figure 9
<p>(<b>a</b>) The experiment location map (the red five-pointed star indicates the experimental site); (<b>b</b>) the location of the radar installation on XIANGYANGHONG-18 (blue oval).</p>
Full article ">Figure 10
<p>The illuminated region of antenna 4 and wave direction.</p>
Full article ">Figure 11
<p>Echo data from the coherent S-band radar observed by antenna 4 on 8 April 2024 from 20:08 to 20:11: (<b>a</b>) the time–Doppler spectra at the 30th range bin; (<b>b</b>) the space-time radial velocity series; (<b>c</b>) the wavenumber–frequency spectrum of (<b>b</b>).</p>
Full article ">Figure 12
<p>(<b>a</b>) The results of the DSWRP from the velocities in <a href="#remotesensing-16-04147-f011" class="html-fig">Figure 11</a>b after adopting the L-curve RLS method; (<b>b</b>) the results of DSWRP from the velocities in <a href="#remotesensing-16-04147-f011" class="html-fig">Figure 11</a>b after adopting the CN-RLS method.</p>
Full article ">Figure 13
<p>Sea surface elevation predicted by S-band coherent radar and wave elevation observed by X-band radar on 8 April 2024.</p>
Full article ">Figure 14
<p>Scatterplot of the predicted wave elevation (S-band radar) versus the observed wave elevation (X-band).</p>
Full article ">Figure 15
<p>(<b>a</b>) Correlation coefficient plot of DSWRP using the L-curve RLS method and the proposed method under different wind speeds; (<b>b</b>) root mean square error plot of DSWRP using the L-curve RLS method and the proposed method under different wind speeds.</p>
Full article ">
14 pages, 4324 KiB  
Article
Mapping Soil Surface Moisture of an Agrophytocenosis via a Neural Network Based on Synchronized Radar and Multispectral Optoelectronic Data of SENTINEL-1,2—Case Study on Test Sites in the Lower Volga Region
by Anatoly Zeyliger, Konstantin Muzalevskiy, Olga Ermolaeva, Anastasia Grecheneva, Ekaterina Zinchenko and Jasmina Gerts
Sustainability 2024, 16(21), 9606; https://doi.org/10.3390/su16219606 - 4 Nov 2024
Viewed by 608
Abstract
In this article, the authors developed a novel method for the moisture mapping of the soil surface of agrophytocenosis using a neural network based on synchronized radar and multispectral optoelectronic data from Sentinel-1,2. The significance of this research lies in its potential to [...] Read more.
In this article, the authors developed a novel method for the moisture mapping of the soil surface of agrophytocenosis using a neural network based on synchronized radar and multispectral optoelectronic data from Sentinel-1,2. The significance of this research lies in its potential to enhance precision farming practices, which are increasingly vital in addressing global agricultural challenges such as water scarcity and the need for sustainable resource management. To verify the developed method, data from two experimental plots were utilized. These plots were located on irrigated soybean crops, with the first plot situated on the right bank (plot No. 1) and the second on the left bank (plot No. 2) of the lower Volga River. Two experimental soil moisture geodatasets were created through measurements and geo-referencing points using the gravimetric method (for plot No. 1) and the proximal sensing method (for plot No. 2) employing the Soil Moisture Sensor ML3-KIT (THETAKIT, Delta). The soil moisture retrieval algorithm was based on the use of a neural network to predict the reflection coefficient of an electro-magnetic wave from the soil surface, followed by inversion into soil moisture using a dielectric model that takes into account the soil texture. The input parameter of the neural network was the ratio of the microwave radar vegetation index (calculated based on Sentinel-1 data) to the index (calculated based on the data of multispectral optoelectronic channels 8 and 11 of Sentinel-2). The retrieved soil moisture values were compared with in situ measurements, showing a determination coefficient of 0.44–0.65 and a standard deviation of 2.4–4.2% for plot No. 1 and similar metrics for plot No. 2. The conducted research laid the groundwork for developing a new technology for remote sensing of soil moisture content in agrophytocenosis, serving as a crucial component of precision farming systems and agroecology. The integration of this technology promotes sustainable agricultural practices by minimizing water consumption while maximizing crop productivity. This aligns with broader environmental goals of conserving natural resources and reducing agricultural runoff. On a larger scale, data derived from such studies can inform policy decisions related to water resource management, guiding regulations that promote efficient water use in agriculture. Full article
(This article belongs to the Special Issue Biotechnology on Sustainable Agriculture)
Show Figures

Figure 1

Figure 1
<p>Test plot No. 1, southwest of the Volgograd city region (<b>a</b>) and test plot No. 2, southeast of the Saratov city (see black dash lines) region (<b>b</b>). Images obtained from Google Maps and the Sentinel-2 satellite in QGIS on 11 July 2020 and 22 August 2022, respectively.</p>
Full article ">Figure 2
<p>Location of soil and plant sampling/measurement points in test plot No. 1, 11 July 2020 (<b>a</b>,<b>c</b>), and test plot No. 2, 22 August 2022 (<b>b</b>,<b>d</b>). Soil moisture interpolation map calculated via soil sampling of test plot No. 1 (<b>a</b>) and test plot No. 2 (<b>b</b>) NDVI map calculated based on Sentinel-2 data, 11 July 2020 (<b>c</b>) and 22 August 2022 (<b>d</b>). The dots in both figures mark the places where samplings/measurements were taken out.</p>
Full article ">Figure 3
<p>The RBC value calculated based on data of Sentinel-1 at VV and VH polarizations as a function of soil volumetric moisture (<b>a</b>) and the relationship between the NDVI and plant height (<b>b</b>) obtained in test plot No. 1.</p>
Full article ">Figure 4
<p>Dependence of the multispectral index I0 calculated based on Sentinel-2 measurements on plant height (<b>a</b>) and dependence of the microwave plant index calculated based on Sentinel-1 measurements on plant height (<b>b</b>), obtained for test plot No. 1.</p>
Full article ">Figure 5
<p>Ratio of multispectral optical index I<sub>0</sub> to microwave index of vegetation versus volumetric soil moisture in test plot No. 1 (<b>a</b>) and ratio of multispectral optical index I<sub>0</sub> to the microwave index of vegetation versus plant height in test plot No. 1 (<b>b</b>).</p>
Full article ">Figure 6
<p>Simple NN with one hidden L1N layer containing N neurons.</p>
Full article ">Figure 7
<p>Coefficient of determination (<b>a</b>) and RMSE (<b>b</b>) between true and predicted reflectance coefficient NN values depending on the number of neurons.</p>
Full article ">Figure 8
<p>Values of volumetric soil moisture reconstructed from Sentinel-1,2 satellite data and NN model depending on soil moisture measured in test No. 1 (sampling plot, see <a href="#sustainability-16-09606-f002" class="html-fig">Figure 2</a>a).</p>
Full article ">Figure 9
<p>Maps of soil surface moisture predicted via NN at test plot No. 1, (<b>a</b>) 9 July 2020 and (<b>b</b>) 9 July 2020. Absolute difference between the soil moisture values predicted via the NN and measured using the gravimetric method at test plot No. 1, (<b>c</b>) 21 July and (<b>d</b>) 9 July 2020.</p>
Full article ">Figure 10
<p>Soil moisture maps of test plot No. 2 built based on training the NN with the input parameters I0 (<b>a</b>), RVI (<b>b</b>), and the pre-trained NN model using the entire data set with the input parameter NN RVI/I0 (<b>c</b>). The maps are built on the same interpolation grid (Sentinel-2, channel 11) with a step of 20 m.</p>
Full article ">Figure 11
<p>Correlation between soil moisture values measured in test plot No. 2 on 22 August 2022, with moisture reconstructed using various input parameters in pre-trained NN: I0 (<b>a</b>), RVI (<b>b</b>), and RVI\I0 (<b>c</b>) (measurement locations, see <a href="#sustainability-16-09606-f002" class="html-fig">Figure 2</a>b).</p>
Full article ">
20 pages, 22822 KiB  
Article
Monitoring Aeolian Erosion from Surface Coal Mines in the Mongolian Gobi Using InSAR Time Series Analysis
by Jungrack Kim, Bayasgalan Amgalan and Amanjol Bulkhbai
Remote Sens. 2024, 16(21), 4111; https://doi.org/10.3390/rs16214111 - 3 Nov 2024
Viewed by 1015
Abstract
Surface mining in the southeastern Gobi Desert has significant environmental impacts, primarily due to the creation of large coal piles that are highly susceptible to aeolian processes. Using spaceborne remote sensing and numerical simulations, we investigated erosional processes and their environmental impacts. Our [...] Read more.
Surface mining in the southeastern Gobi Desert has significant environmental impacts, primarily due to the creation of large coal piles that are highly susceptible to aeolian processes. Using spaceborne remote sensing and numerical simulations, we investigated erosional processes and their environmental impacts. Our primary tool was Interferometric Synthetic Aperture Radar (InSAR) data from Sentinel-1 imagery collected between 2017 and 2022. We analyzed these data using phase angle information from the Small Baseline InSAR time series framework. The time series analyses revealed intensive aeolian erosion in the coal piles, represented as thin deformation patterns along the potential pathways of aerodynamic transportation. Further analysis of multispectral data, combined with correlations between wind patterns and trajectory simulations, highlighted the detrimental impact of coal dust on the surrounding environment and the mechanism of aeolian erosion. The lack of mitigation measures, such as water spray, appeared to exacerbate erosion and dust generation. This study demonstrates the feasibility of using publicly available remote sensing data to monitor coal mining activities and their environmental hazards. Our findings contribute to a better understanding of coal dust generation processes in surface mining operations as well as the aeolian erosion mechanism in desert environments. Full article
(This article belongs to the Special Issue Remote Sensing and Geophysics Methods for Geomorphology Research)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location of study areas, (<b>b</b>) geographical and topographic context of Ail Bayan and Tavan Tolgoi coal mines, (<b>c</b>) surrounding hydrological contexts (source: <a href="https://eic.mn/" target="_blank">https://eic.mn/</a>, accessed on 1 July 2024), (<b>d</b>) coal production and transportation at Tavan Tolgoi (43.625N, 105.474E) and (<b>e</b>) coal dust generation at Ail Bayan (43.717N, 108.946E) (images were taken in March 2019). Note that the transportation vehicles in Tavan Tolgoi are well-confined so as not to produce coal dust.</p>
Full article ">Figure 2
<p>Acquisition times and connection graphs of employed ascending/descending Sentinel-1 InSAR pairs over Ail Bayan (<b>a</b>,<b>b</b>) and Tavan Tolgoi (<b>c</b>–<b>e</b>). Note that the phase coherences of InSAR pairs are always higher than 0.7. In Ail Bayan, the ascending mode InSAR observations were interpolated into the descending mode time domain for decomposition. Similarly, over Tavan Tolgoi, the ascending and descending mode observations in path 62 were interpolated into the descending mode time domain of path 135. While the thresholds for perpendicular and temporal baselines were set to 150 m and 25 days, respectively, some InSAR pairs exceeding these thresholds were included to enhance interferometric coverage.</p>
Full article ">Figure 3
<p>Processing workflow for InSAR time series data, including integration with other satellite and spatial datasets.</p>
Full article ">Figure 4
<p>(<b>a</b>) Google map, which reveals the details of Ail Bayan, (<b>b</b>) topography presented in Copernicus 30 m DEM, (<b>c</b>) ascending LOS velocity and (<b>d</b>) descending LOS deformation velocity.</p>
Full article ">Figure 5
<p>(<b>a</b>) Decomposed horizontal velocity and (<b>b</b>) vertical velocity in Ail Bayan. The overlaid average wind velocities were extracted using GEE and interpolated to a 1 km resolution from the original 11.132 km ERA5-Land data using kriging.</p>
Full article ">Figure 6
<p>(<b>a</b>) Google map which reveals the details of Tavan Tolgoi, (<b>b</b>) topography presented in Copernicus 30 m DEM, (<b>c</b>) ascending LOS velocity, (<b>d</b>) descending LOS deformation velocity of path 135 coverage, and (<b>e</b>) descending LOS deformation velocity of path 62 coverage.</p>
Full article ">Figure 7
<p>Decomposed velocities in Tavan Tolgoi: (<b>a</b>) horizontal velocity and (<b>b</b>) vertical velocity. Note that the wind directions are similar to those in the Ail Bayan area, blowing from west to east.</p>
Full article ">Figure 8
<p>The behavior of seven RoIs along with mean wind velocities in different modes: (<b>a</b>) ascending mode, (<b>b</b>) descending mode, (<b>c</b>) decomposed horizontal deformation velocities, and (<b>d</b>) decomposed vertical deformation velocities.</p>
Full article ">Figure 9
<p>Correlation maps between InSAR deformation velocities and average wind velocities for corresponding periods in (<b>a</b>) ascending mode, (<b>b</b>) descending mode, (<b>c</b>) horizontal component of decomposed InSAR velocities, and (<b>d</b>) vertical component of decomposed InSAR velocities.</p>
Full article ">Figure 10
<p>Spectral signature analyses using Sentinel-2 time series images on (<b>a</b>) 23 March 2018, (<b>b</b>) 17 April 2018, (<b>c</b>) 2 May 2018, and (<b>d</b>) 22 May 2018. Note that a lower SID value indicates greater spectral similarity. (<b>e</b>) Visual band view of Sentinel-2 image (2 May 2018), (<b>f</b>) spectral signatures in Down1 and Down3/Up3 areas representing coal mine and major FMP aeolian sites.</p>
Full article ">Figure 11
<p>Wind factors influencing coal mine dust generation: (<b>a</b>) monthly wind velocity at an altitude of 10 m, (<b>b</b>) friction velocities for different 10 m wind velocities and roughness lengths, (<b>c</b>) trajectory simulations originating from the coal mine during the sand dust season from March to May 2018, (<b>d</b>) trajectory simulations during the summer season of 2018.</p>
Full article ">Figure 12
<p>Average NMDI maps for (<b>a</b>) Ail Bayan from 24 September 2017 to 14 August 2018, (<b>b</b>) the same region from 23 January 2017 to 9 May 2022, (<b>c</b>) the Tavan Tolgoi region from 11 September 2017 to 10 March 2018, and (<b>d</b>) the same region from 4 January 2017 to 26 May 2022.</p>
Full article ">Figure 13
<p>Environmental consequences of coal mine dust generation: (<b>a</b>) HYSPLIT trajectory simulations originating from the coal mine in 8 March 2018, (<b>b</b>) HYSPLIT trajectory simulations originating from the coal mine in 18 April 2018, using an ensemble HYSPLIT model with 150-h forward trajectory options. (<b>c</b>) Ground photos from Down3 area in Ail Bayan showing contaminated soil and vegetation by blown FMP (images were taken in March 2019).</p>
Full article ">
Back to TopTop