[go: up one dir, main page]
More Web Proxy on the site http://driver.im/
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

Article Types

Countries / Regions

Search Results (40)

Search Parameters:
Keywords = ALOS 30 m

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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 513
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 ">
20 pages, 5259 KiB  
Article
Voltammetric Sensor Based on Titania Nanoparticles Synthesized with Aloe vera Extract for the Quantification of Dithiophosphates in Industrial and Environmental Samples
by Javier E. Vilasó-Cadre, Alondra Ramírez-Rodríguez, Juan Hidalgo, Iván A. Reyes-Domínguez, Roel Cruz, Mizraim U. Flores, Israel Rodríguez-Torres, Roberto Briones-Gallardo, Luis Hidalgo and Juan Jesús Piña Leyte-Vidal
Chemosensors 2024, 12(9), 195; https://doi.org/10.3390/chemosensors12090195 - 22 Sep 2024
Viewed by 1441
Abstract
In this work, TiO2 spherical nanoparticles with a mean diameter of 10.08 nm (SD = 4.54 nm) were synthesized using Aloe vera extract. Rutile, brookite, and anatase crystalline phases were identified. The surface morphology of a carbon paste electrode does not change [...] Read more.
In this work, TiO2 spherical nanoparticles with a mean diameter of 10.08 nm (SD = 4.54 nm) were synthesized using Aloe vera extract. Rutile, brookite, and anatase crystalline phases were identified. The surface morphology of a carbon paste electrode does not change in the presence of nanoparticles; however, the surface chemical composition does. The voltammetric response to dicresyl dithiophosphate was higher when the electrode was modified with TiO2 nanoparticles. After an electrochemical response study from pH 1.0 to 12.0, pH 7.0 was selected for the electroanalysis. The electroactive area of the modified sensor was 0.036 cm2, while it was 0.026 cm2 for the bare electrode. The oxidation process showed mixed adsorption-diffusion control. The charge transfer resistance of the modified sensor (530.1 Ω, SD = 4.08 Ω) was much lower than that of the bare electrode (4298 Ω, SD = 8.53 Ω). The linear quantitative range by square wave voltammetry was from 5 to 150 μmol/L, with a limit of detection of 1.89 μmol/L and a limit of quantification of 6.26 μmol/L under optimal pulse parameters of 50 Hz frequency, 1 mV step potential, and 25 mV pulse amplitude. The sensor response was repeatable and reproducible over 30 days. The results on real flotation and synthetically contaminated soil samples were statistically equivalent to those obtained by UV-vis spectrophotometry. A dithiocarbamate showed an interfering effect on the sensor response to dithiophosphate. Full article
(This article belongs to the Special Issue Advances in Electrochemical Sensing and Analysis)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>X-ray diffractogram of the TiO<sub>2</sub> nanoparticles synthetized using <span class="html-italic">Aloe vera</span> extract.</p>
Full article ">Figure 2
<p>(<b>a</b>) Transmission electron micrograph of TiO<sub>2</sub> nanoparticles obtained using <span class="html-italic">Aloe vera</span> extract, (<b>b</b>) Transmission micrograph at higher magnification, (<b>c</b>) Frequency histogram and size distribution curve of the nanoparticles, and (<b>d</b>) EDS spectrum of TiO<sub>2</sub> nanoparticles.</p>
Full article ">Figure 3
<p>SEM secondary electron micrograph: (<b>a</b>) CPE surface, (<b>b</b>) TiO<sub>2</sub>/CPE surface. SEM backscattered electron micrograph and EDS spectrum: (<b>c</b>) CPE surface, (<b>d</b>) TiO<sub>2</sub>/CPE surface; and (<b>e</b>) EDS mapping of a TiO<sub>2</sub>/CPE surface section.</p>
Full article ">Figure 4
<p>Cyclic voltammograms of a 5.4 mmol/L DCDTP solution at pH 12.0 using the CPE and the TiO<sub>2</sub>/CPE. Supporting electrolyte: 0.1 mol/L phosphate buffer and 0.1 mol/L KNO<sub>3</sub>. Scan rate: 50 mV/s.</p>
Full article ">Figure 5
<p>Current peak intensity using the TiO<sub>2</sub>/CPE versus the pH value of a 5.4 mmol/L DCDTP solution. Technique: Cyclic voltammetry. Supporting electrolyte: 0.1 mol/L KNO<sub>3</sub>. Scan rate: 50 mV/s.</p>
Full article ">Figure 6
<p>(<b>a</b>) Cyclic voltammograms of a 5.4 mmol/L DCDTP solution at different scan rates (supporting electrolyte: 0.1 mol/L phosphate buffer and 0.1 mol/L KNO<sub>3</sub>, pH 7.0), (<b>b</b>) Plot of the logarithm of the peak current versus the logarithm of the scan rate.</p>
Full article ">Figure 7
<p>(<b>a</b>) Cyclic voltammograms for a 10 mmol/L ferrocyanide solution in 1 mol/L KNO<sub>3</sub> at different scan rates, (<b>b</b>) Plot of peak current as a function of the square root of the scan rate.</p>
Full article ">Figure 8
<p>Nyquist plots for the CPE and the TiO<sub>2</sub>/CPE in 10 mmol/L potassium ferrocyanide dissolved in 1 mol/L KNO<sub>3</sub> as the supporting electrolyte. Inset: Equivalent circuit for both electrodes. Dots and triangles: Experimental data, Solid line: Simulated data from the equivalent circuit.</p>
Full article ">Figure 9
<p>Optimization of the peak current intensity for dithiophosphate by square wave voltammetry: (<b>a</b>) Effect of the pulse frequency keeping the potential step at 1 mV and the pulse amplitude at 5 mV, (<b>b</b>) effect of the potential step keeping the frequency at 50 Hz and the pulse amplitude at 5 mV, and (<b>c</b>) effect of the pulse amplitude keeping the frequency at 50 Hz and the potential step at 1 mV.</p>
Full article ">Figure 10
<p>Square wave voltammograms of solutions with different concentrations of DCDTP using the TiO<sub>2</sub>/CPE. Inset: Calibration curve (<span class="html-italic">N</span> = 3). Supporting electrolyte: 0.1 mol/L phosphate buffer and 0.1 mol/L KNO<sub>3</sub>, pH 7.0. Frequency: 50 Hz, potential step: 1 mV, pulse amplitude: 25 mV.</p>
Full article ">Figure 11
<p>Samples for the trueness test: (<b>a</b>) Flotation sample and (<b>b</b>) soil extract sample.</p>
Full article ">Figure 12
<p>(<b>a</b>) Cyclic voltammograms for three solutions containing 5.4 mmol/L DCDTP, 5.4 mmol/L DCDTP + 1 mmol/L NaIPX, and 5.4 mmol/L DCDTP + 0.05 mL DADTC, respectively (supporting electrolyte: 0.1 mol/L phosphate buffer and 0.1 mol/L KNO<sub>3</sub>, pH 7.0. Scan rate: 100 mV/s), (<b>b</b>) Plot of the peak current intensity versus the solution composition.</p>
Full article ">
22 pages, 4205 KiB  
Article
Sustainable Geoinformatic Approaches to Insurance for Small-Scale Farmers in Colombia
by Ahmad Abd Rabuh, Richard M. Teeuw, Doyle Ray Oakey, Athanasios V. Argyriou, Max Foxley-Marrable and Alan Wilkins
Sustainability 2024, 16(12), 5104; https://doi.org/10.3390/su16125104 - 15 Jun 2024
Cited by 1 | Viewed by 1368
Abstract
This article presents a low-cost insurance system developed for smallholder farms in disaster-prone regions, primarily using free Earth observation (EO) data and free open source software’s (FOSS), collectively termed “sustainable geoinformatics.” The study examined 30 farms in Risaralda Department, Colombia. A digital elevation [...] Read more.
This article presents a low-cost insurance system developed for smallholder farms in disaster-prone regions, primarily using free Earth observation (EO) data and free open source software’s (FOSS), collectively termed “sustainable geoinformatics.” The study examined 30 farms in Risaralda Department, Colombia. A digital elevation model (12.5 m pixels) from the ALOS PALSAR satellite sensor was used with a geographic information system (GIS) to map the terrain, drainage, and geohazards of each farming district. Google Earth Engine (GEE) was used to carry out time-series analysis of 15 EO and weather datasets for 1998 to 2020. This analysis enabled the levels of risk from hydrometeorological hazards to be determined for each farm of the study, providing key data for the setting of insurance premiums. A parametric insurance product was developed using a proprietary mobile phone app that collected GPS-tagged, time-stamped mobile phone photos to verify crop damage, with further verification of crop health also provided by daily near-real-time satellite imagery (e.g., PlanetScope with 3 m pixels). Machine learning was used for feature identification with the photos and the satellite imagery. Key features of this insurance system are its low operational cost and rapid damage verification relative to conventional approaches to farm insurance. This relatively fast, low-cost, and affordable approach to insurance for small-scale farming enhances sustainable development by enabling policyholder farmers to recover more quickly from disasters. Full article
Show Figures

Figure 1

Figure 1
<p>Conceptual model of the insurance system and its geoinformatic components.</p>
Full article ">Figure 2
<p>Location of the study area in Riseralda Department, Colombia. Areas of forest are shown in dark green. The inset box indicates the area shown in detail within <a href="#sustainability-16-05104-f003" class="html-fig">Figure 3</a> (map source: OpenStreetMap).</p>
Full article ">Figure 3
<p>The Dosquebradas study area (corresponding to the inset box in <a href="#sustainability-16-05104-f002" class="html-fig">Figure 2</a>): locations and areal extents of the farms included in the testing of the insurance system.</p>
Full article ">Figure 4
<p>Satellite image of a studied farm, overlain with locations of GPS-tagged time-stamped photos from the InsurTech mobile phone app, taken at 50 m intervals around a field boundary, with inset example photo.</p>
Full article ">Figure 5
<p>An excerpt from the timeline chart of climate indices and MODIS NDVI for the Dosquebradas district from 2009 to 2015. Periods with flooding are highlighted in blue and periods of drought are shown in orange, with brown indicating drought with numerous wildfires.</p>
Full article ">Figure 6
<p>Long short-term memory (LSTM) layers.</p>
Full article ">Figure 7
<p>Parameter setting and loss value optimization to derive unbiased weights.</p>
Full article ">Figure 8
<p>Landform types of Dosquebradas district produced from DEM geomorphometrics. The inset box indicates the area examined in detail within <a href="#sustainability-16-05104-f009" class="html-fig">Figure 9</a>a,b.</p>
Full article ">Figure 9
<p>The Dosquebradas study area, with overlain outlines of the surveyed farms: (<b>a</b>) landslide hazard zones, with the arrow pointing at the eastern border of farms 17 and 18; (<b>b</b>) flood hazard zones, with the arrow pointing at parts of farms 17 and 18 at risk of flooding.</p>
Full article ">Figure 10
<p>Agriculture supply chain.</p>
Full article ">
24 pages, 19755 KiB  
Article
Vertical Accuracy Assessment and Improvement of Five High-Resolution Open-Source Digital Elevation Models Using ICESat-2 Data and Random Forest: Case Study on Chongqing, China
by Weifeng Xu, Jun Li, Dailiang Peng, Hongyue Yin, Jinge Jiang, Hongxuan Xia and Di Wen
Remote Sens. 2024, 16(11), 1903; https://doi.org/10.3390/rs16111903 - 25 May 2024
Cited by 3 | Viewed by 2054
Abstract
Digital elevation models (DEMs) are widely used in digital terrain analysis, global change research, digital Earth applications, and studies concerning natural disasters. In this investigation, a thorough examination and comparison of five open-source DEMs (ALOS PALSAR, SRTM1 DEM, SRTM3 DEM, NASADEM, and ASTER [...] Read more.
Digital elevation models (DEMs) are widely used in digital terrain analysis, global change research, digital Earth applications, and studies concerning natural disasters. In this investigation, a thorough examination and comparison of five open-source DEMs (ALOS PALSAR, SRTM1 DEM, SRTM3 DEM, NASADEM, and ASTER GDEM V3) was carried out, with a focus on the Chongqing region as a specific case study. By utilizing ICESat-2 ATL08 data for validation and employing a random forest model to refine terrain variables such as slope, aspect, land cover, and landform type, a study was undertaken to assess the precision of DEM data. Research indicates that spatial resolution significantly impacts the accuracy of DEMs. ALOS PALSAR demonstrated satisfactory performance, reducing the corrected root mean square error (RMSE) from 13.29 m to 9.15 m. The implementation of the random forest model resulted in a significant improvement in the accuracy of the 30 m resolution NASADEM product. This improvement was supported by a decrease in the RMSE from 38.24 m to 9.77 m, demonstrating a significant 74.45% enhancement in accuracy. Consequently, the ALOS PALSAR and NASADEM datasets are considered the preferred data sources for mountainous urban areas. Furthermore, the study established a clear relationship between the precision of DEMs and slope, demonstrating a consistent decline in precision as slope steepness increases. The influence of aspect on accuracy was considered to be relatively minor, while vegetated areas and medium-to-high-relief mountainous terrains were identified as the main challenges in attaining accuracy in the DEMs. This study offers valuable insights into selecting DEM datasets for complex terrains in mountainous urban areas, highlighting the critical importance of choosing the appropriate DEM data for scientific research. Full article
Show Figures

Figure 1

Figure 1
<p>Study area and ICESat-2 ATL08 tracks. (<b>a</b>) The location of the study area. (<b>b</b>) The overall tracks of ICESat-2 ATL08; the black dots represent the laser footprint of the ICESat-2 satellite on the ground. (<b>c</b>) Land-cover type. (<b>d</b>) Landform type: A–E represent, respectively, terrace, hills, small rolling hills, medium rolling hills, and large rolling hills.</p>
Full article ">Figure 2
<p>Accuracy evaluation and correction flow chart of five DEMs using ICESat-2.</p>
Full article ">Figure 3
<p>Histogram of the elevation error for ALOS PALSAR, SRTM1 DEM, SRTM3 DEM, NASADEM, and ASTER GDEM V3.</p>
Full article ">Figure 4
<p>ALOS PALSAR, SRTM1 DEM, SRTM3 DEM, NASADEM, and ASTER GDEM V3 elevation error histograms and scatter plots before and after the random forest model correction.</p>
Full article ">Figure 5
<p>Scatter density plots of the altimetric error before and after correction for ALOS PALSAR (<b>A</b>,<b>a</b>), SRTM1 DEM (<b>B</b>,<b>b</b>), SRTM3 DEM (<b>C</b>,<b>c</b>), NASADEM (<b>D</b>,<b>d</b>), and ASTER GDEM V3 (<b>E</b>,<b>e</b>) at different slopes.</p>
Full article ">Figure 6
<p>Differences between five DEMs and ICESat-2 ATL08 depicted by radial radar plots for different slope directions before and after correction (red diamonds in the figure indicate RMSE): (<b>A</b>,<b>a</b>) ALOS PALSAR, (<b>B</b>,<b>b</b>) SRTM1 DEM, (<b>C</b>,<b>c</b>) SRTM3 DEM, (<b>D</b>,<b>d</b>) NASADEM, and (<b>E</b>,<b>e</b>) ASTER GDEM V3.</p>
Full article ">Figure 7
<p>Error histograms of ALOS PALSAR, SRTM1 DEM, SRTM3 DEM, NASADEM and ASTER GDEM V3 before and after correction under different land cover types. CL: cropland, FR: forest, GL: grassland, SL: shrubland, WB: water body, AS: artificial surface.</p>
Full article ">Figure 8
<p>Normal curve and axis whisker plots of error distribution of five DEMs before and after correction for different landform types.</p>
Full article ">Figure 9
<p>Before and after comparison of five DEM corrections in Chongqing.</p>
Full article ">Figure 10
<p>Local features presented by the DEM data before and after the five corrections for different landforms.</p>
Full article ">
22 pages, 19794 KiB  
Article
Monitoring and Comparative Analysis of Hohhot Subway Subsidence Using StaMPS-PS Based on Two DEMS
by Sihai Zhao, Peixian Li, Hairui Li, Tao Zhang and Bing Wang
Remote Sens. 2023, 15(16), 4011; https://doi.org/10.3390/rs15164011 - 13 Aug 2023
Cited by 6 | Viewed by 1744
Abstract
The subway alleviates the traffic pressure in the city but also brings the potential risk of land subsidence. The land subsidence caused by the subway is a global problem that seriously affects the safety of subway operations and surrounding buildings. Therefore, it is [...] Read more.
The subway alleviates the traffic pressure in the city but also brings the potential risk of land subsidence. The land subsidence caused by the subway is a global problem that seriously affects the safety of subway operations and surrounding buildings. Therefore, it is very important to carry out long-term deformation monitoring on the subway system. StaMPS-PS is a time-series Interferometric Synthetic Aperture Radar (InSAR) technique that serves as an effective means for monitoring urban ground subsidence. However, the accuracy of external (Digital Elevation Models) DEM will affect the accuracy of StaMPS-PS monitoring, and previous studies have mostly used SRTM-1 arc DEM (30 m) as the external DEM. In this study, to obtain a more precise measurement of surface deformation caused by the excavation of the Hohhot subway, a total of 85 scenes of Sentinel-1A data from July 2015 to October 2021, as well as two different resolution digital elevation models (DEMs) (ALOS PALSAR DEM and SRTM-1 arc DEM), were used to calculate and analyze the subsidence along the subway line in Hohhot city. The StaMPS-PS monitoring results showed the ALOS PALSAR DEM, as an external DEM, had higher accuracy, and there was regional subsidence in both the construction processes of Line 1 and Line 2 of the Hohhot subway, with a maximum subsidence rate of −21.1 mm/year. The dynamic changes in subway subsidence were fitted using the Peck formula and the long short-term memory (LSTM) model. The Peck formula results showed the width and maximum subsidence of the settlement troughs gradually expanded during the construction of the subway. The predicted values of the Root Mean Square Error (RMSE) and Mean Absolute Percentage Error (MAPE) of the LSTM model were less than 4 mm and 10%, respectively, consistent with the measured results. Furthermore, we discussed the factors that affect settlement along the subway line and the impact of two external DEMs on StaMPS-PS. The study results provide a scientific method for DEM selection and subsidence analysis calculations in the StaMPS-PS monitoring of urban subway subsidence. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>SAR image coverage and study area range: (<b>a</b>) Study area and SAR image coverage; (<b>b</b>) Enlarged study area and metro line distribution map.</p>
Full article ">Figure 2
<p>Long short-term memory model (LSTM) structure diagram: (<b>a</b>) LSTM neuronal structure; (<b>b</b>) structure of the subsidence prediction model.</p>
Full article ">Figure 3
<p>Comparison of level results with StaMPS-PS results.</p>
Full article ">Figure 4
<p>StaMPS-PS results using DEM with 12.5 m ALOS PALSAR DEM.</p>
Full article ">Figure 5
<p>Deformation profile of subway line 1.</p>
Full article ">Figure 6
<p>Subway line 1 subsidence area map and statistical histogram of deformation rate distribution: (<b>a</b>) Subsidence area of subway line 1; (<b>b</b>) subsidence area A; (<b>c</b>) subsidence area B; (<b>d</b>) subsidence area C; (<b>e</b>) profile line of Hugangdonglu.</p>
Full article ">Figure 7
<p>Statistical histogram of deformation rate distribution: (<b>a</b>) Statistical histogram of deformation rate distribution in subsidence area A of subway line 1; (<b>b</b>) statistical histogram of deformation rate distribution in subsidence area B of subway line 1; (<b>c</b>) statistical histogram of deformation rate distribution in subsidence area C of subway line 1. (<b>d</b>) Statistical histogram of deformation rate distribution in subsidence area A of subway line 2; (<b>e</b>) statistical histogram of deformation rate distribution in subsidence area B of subway line 2; (<b>f</b>) statistical histogram of deformation rate distribution in subsidence area C of subway line 2.</p>
Full article ">Figure 8
<p>Typical PS points of subsidence trend in the subsidence area: (<b>a</b>) Typical PS point subsidence trend in subsidence area A of subway line 1; (<b>b</b>) Typical PS point subsidence trend in subsidence area B of subway line 1; (<b>c</b>) Typical PS point subsidence trend in subsidence area C of subway line 1; (<b>d</b>) Typical PS point subsidence trend in subsidence area A of subway line 2; (<b>e</b>) Typical PS point subsidence trend in subsidence area B of subway line 2; (<b>f</b>) Typical PS point subsidence trend in subsidence area C of subway line 2.</p>
Full article ">Figure 9
<p>Deformation profile of subway line 2.</p>
Full article ">Figure 10
<p>Subway line 2 subsidence area map and statistical histogram of deformation rate distribution: (<b>a</b>) Subsidence area of subway line 1; (<b>b</b>) subsidence area A; (<b>c</b>) subsidence area B; (<b>d</b>) subsidence area C; (<b>e</b>) profile line of Zhongshanlu.</p>
Full article ">Figure 11
<p>Peck formula: (<b>a</b>) Fitting results of the Peck formula for the settlement trough near Hugangdonglu; (<b>b</b>) fitting results of the Peck formula for the settlement trough near Zhongshanlu.</p>
Full article ">Figure 12
<p>Loss function and prediction chart: (<b>a</b>) Loss function chart of affiliated hospitals; (<b>b</b>) loss function chart of Hugangdonglu; (<b>c</b>) prediction chart of affiliated hospitals; (<b>d</b>) prediction chart of Hugangdonglu.</p>
Full article ">Figure 13
<p>Statistical chart of coherence under two external DEMs.</p>
Full article ">Figure 14
<p>Relationship between DEM differences and average deformation rate differences.</p>
Full article ">
37 pages, 11449 KiB  
Article
Polarimetric L-Band ALOS2-PALSAR2 for Discontinuous Permafrost Mapping in Peatland Regions
by Ridha Touzi, Steven M. Pawley, Paul Wilson, Xianfeng Jiao, Mehdi Hosseini and Masanobu Shimada
Remote Sens. 2023, 15(9), 2312; https://doi.org/10.3390/rs15092312 - 27 Apr 2023
Cited by 6 | Viewed by 2411
Abstract
Recently, it has been shown that the long penetrating polarimetric L-band ALOS is very promising for boreal and subarctic peatland mapping and monitoring. The unique information provided by the Touzi decomposition, and the dominant-scattering-type phase in particular, on peatland subsurface water flow permits [...] Read more.
Recently, it has been shown that the long penetrating polarimetric L-band ALOS is very promising for boreal and subarctic peatland mapping and monitoring. The unique information provided by the Touzi decomposition, and the dominant-scattering-type phase in particular, on peatland subsurface water flow permits an enhanced discrimination of bogs from fens, two peatland classes that can hardly be discriminated using conventional optical remote sensing sensors and C-band polarimetric SAR. In this study, the dominant and medium-scattering phases generated by the Touzi decomposition are investigated for discontinuous permafrost mapping in peatland regions. Polarimetric ALOS2, LiDAR, and field data were collected in the middle of August 2014, at the maximum permafrost thaw conditions, over discontinuous permafrost distributed within wooded palsa bogs and peat plateaus near the Namur Lake (Northern Alberta). The ALOS2 image, which was miscellaneously calibrated with antenna cross talk (−33 dB) much higher than the actual ones, was recalibrated. This led to a reduction of the residual calibration error (down to −43 dB) and permitted a significant improvement of the dominant and medium-scattering-type phase (20 to −30) over peatlands underlain by discontinuous permafrost. The Touzi decomposition, Cloude–Pottier α-H incoherent target scattering decomposition, and the HH-VV phase difference were investigated, in addition to the conventional multipolarization (HH, HV, and VV) channels, for discontinuous permafrost mapping using the recalibrated ALOS2 image. A LiDAR-based permafrost classification developed by the Alberta Geological Survey (AGS) was used in conjunction with the field data collected during the ALOS2 image acquisition for the validation of the results. It is shown that the dominant- and scattering-type phases are the only polarimetric parameters which can detect peatland subsurface discontinuous permafrost. The medium-scattering-type phase, ϕs2, performs better than the dominant-scattering-type phase, ϕs1, and permits a better detection of subsurface discontinuous permafrost in peatland regions. ϕs2 also allows for a better discrimination of areas underlain by permafrost from the nonpermafrost areas. The medium Huynen maximum polarization return (m2) and the minimum degree of polarization (DoP), pmin, can be used to remove the scattering-type phase ambiguities that might occur in areas with deep permafrost (more than 50 cm in depth). The excellent performance of polarimetric PALSAR2 in term of NESZ (−37 dB) permits the demonstration of the very promising L-band long-penetration SAR capabilities for enhanced detection and mapping of relatively deep (up to 50 cm) discontinuous permafrost in peatland regions. Full article
(This article belongs to the Special Issue SAR, Interferometry and Polarimetry Applications in Geoscience)
Show Figures

Figure 1

Figure 1
<p>AGS Classification.</p>
Full article ">Figure 2
<p>AGCC classification; burned areas are presented in light pink.</p>
Full article ">Figure 3
<p>AWI classification.</p>
Full article ">Figure 4
<p>ALOS2 multipolarization image: HH (red), HV (green), and VV (Blue).</p>
Full article ">Figure 5
<p>HH original–recalibrated magnitude ratio.</p>
Full article ">Figure 6
<p>VV original–recalibrated magnitude ratio.</p>
Full article ">Figure 7
<p>HH-VV phase difference.</p>
Full article ">Figure 8
<p>HV original–recalibrated magnitude ratio.</p>
Full article ">Figure 9
<p>VH original–recalibrated magnitude ratio.</p>
Full article ">Figure 10
<p>VH intensity in dB.</p>
Full article ">Figure 11
<p>Scattering type Phis1o derived from the original ALOS2 image. The contours of permafrost areas generated from the AGS classification are included in the image.</p>
Full article ">Figure 12
<p>Scattering type phase Phis1o derived using the recalibrated ALOS2 image. The contours of permafrost areas generated from the AGS classification are included in the image.</p>
Full article ">Figure 13
<p>Phis1o recalibrated–original image error (absolute phase difference in degrees).</p>
Full article ">Figure 14
<p>Scattering-type phase Phis2 derived using the recalibrated ALOS2 image.</p>
Full article ">Figure 15
<p>Phis2 recalibrated–original image error (dB).</p>
Full article ">Figure 16
<p>Dominant-scattering-type magnitude alphas1 obtained using the recalibrated ALOS2 image.</p>
Full article ">Figure 17
<p>alphas1 calibration error: absolute difference (in degrees) between the scattering-type magnitude obtained from the original image and the one obtained from the recalibrated image.</p>
Full article ">Figure 18
<p>Study area: multi-pol ALOS2 image.</p>
Full article ">Figure 19
<p>AGS classification.</p>
Full article ">Figure 20
<p>AWI classification.</p>
Full article ">Figure 21
<p>AGCC classification: burned areas are presented in light pink.</p>
Full article ">Figure 22
<p>Touzi dominant-scattering-type phase Phis1o derived using the recalibrated ALOS2 image.</p>
Full article ">Figure 23
<p>Touzi medium-scattering-type phase Phis2 derived using the recalibrated ALOS2 image.</p>
Full article ">Figure 24
<p>HH-VV phase difference.</p>
Full article ">Figure 25
<p>Minimum DoP pmin.</p>
Full article ">Figure 26
<p>Maximum DoP pmax.</p>
Full article ">Figure 27
<p>Site A: AGS permafrost classification.</p>
Full article ">Figure 28
<p>Site A: AWI classification.</p>
Full article ">Figure 29
<p>Site A: AGCC classification.</p>
Full article ">Figure 30
<p>Site A: Touzi scattering-type phases: phis1o and phis2.</p>
Full article ">Figure 31
<p>Site A: minimum DoP pmin. The legend of the different colored dots assigned to the permafrost and non-permafrost sites is given in <a href="#remotesensing-15-02312-f027" class="html-fig">Figure 27</a>.</p>
Full article ">Figure 32
<p>Site A: medium scattering’s Huynen maximum polarization return (m<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>). The legend of the different colored dots assigned to the permafrost and non-permafrost sites is given in <a href="#remotesensing-15-02312-f027" class="html-fig">Figure 27</a>.</p>
Full article ">Figure 33
<p>Site A: HH-VV phase diffrence. The legend of the different colored dots assigned to the permafrost and non-permafrost sites is given in <a href="#remotesensing-15-02312-f027" class="html-fig">Figure 27</a>.</p>
Full article ">Figure 34
<p>Site A: multipolarization channel and span curves (in dB). Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost; CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 35
<p>Site A: Scattering type phase (Phis1, Phis2) and HH-VV phase difference curves (in degrees). Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost; CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 36
<p>Site A: pmax and pmin curves. Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost (1.8 m and more); CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 37
<p>Site A: Touzi scattering-type magnitudes and Cloude alpha (in degrees). Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost; CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 38
<p>Site A: ICTD eigenvalues and entropy. Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost; CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 39
<p>Site B: AGS classification.</p>
Full article ">Figure 40
<p>Site B: Touzi scattering-type phase: phis1o and Phis2.</p>
Full article ">Figure 41
<p>Site B: HH-VV phase difference.</p>
Full article ">Figure 42
<p>Site B: pmin.</p>
Full article ">Figure 43
<p>Site B: multipolarization and span curves (in dB). Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost (1.8 m and more); CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 44
<p>Site B: Scattering type phase (Phis1, Phis2) in degrees. Sample classes: BF: bog permafrost; DBF: deep bog permafrost; VDF: very deep permafrost; CS: collapse scar; FC: forest conifer.</p>
Full article ">Figure 45
<p>Site B: pmin and pmax curves.</p>
Full article ">Figure 46
<p>Site B: medium scattering’s Huynen maximum polarization return (m<math display="inline"><semantics> <msub> <mrow/> <mn>2</mn> </msub> </semantics></math>).</p>
Full article ">
20 pages, 8747 KiB  
Article
Impact of Topography on Rural Cycling Patterns: Case Study of Bugesera District, Rwanda
by Jean-Claude Baraka Munyaka, Jérôme Chenal, Alexis Gatoni Sebarenzi, Rim Mrani and Akuto Akpedze Konou
Urban Sci. 2023, 7(1), 8; https://doi.org/10.3390/urbansci7010008 - 13 Jan 2023
Viewed by 2741
Abstract
Rural mobility in Africa is an under-researched issue. Rural communities have often suffered from reduced mobility that has hampered their access to essential services and facilities such as education, health care, food, and clean water. In many rural communities, a more affordable mobility [...] Read more.
Rural mobility in Africa is an under-researched issue. Rural communities have often suffered from reduced mobility that has hampered their access to essential services and facilities such as education, health care, food, and clean water. In many rural communities, a more affordable mobility option, such as non-motorized mobility (cycling and walking), is the preferred way for people to travel. Apart from its well-known advantages, little is known about the impact of topography and routes on the mobility options adopted by rural communities. Therefore, this study aims to use Digital Elevation Models (DEMs) to analyze the impact of topography and routes on rural mobility patterns at the level of formal and informal cycle track networks in the Bugesera District, focusing on the Nyamata and Mayange sectors, Rwanda. This study used GPS devices given to 50 participants to collect mobility patterns in the two previously mentioned sectors. Then, the study imposed a 30-m buffer on the official road networks of Rwanda collected by the Rwanda Transport Development Agency (RTDA). These data were joined to GPS tracks to highlight official and unofficial roads (tracks that did not fall within the 30-m buffer). In addition, Digital Elevation Models were applied to analyze the SRTM (30 m resolution) and ALOS PALSAR (12.5 m resolution) elevation data of the Bugesera region. The findings revealed an elevation range of 1333 to 1535 for SRTM and 1323 to 1641 for ALOS PALSAR. The study calculated the slope to find the slope percentage (m) and length (m). The findings from the DEMs and the slope calculation revealed that Bugesera has a relatively flat surface, favorable for cycling. The slope percentage was further classified into five levels of slope ranging from steep to very steep. And the Van Zuidam classification results confirmed that Bugesera has “a flat or nearly flat surface, without significant denudation processes”. With a favorable topography and a higher bicycle ownership ratio, residents of the Bugesera district favor cycling in their daily activities more than any district in Rwanda. Cycling in Nyamata or Mayange links residents to areas with higher social, educational, administrative, and economic activities. Full article
Show Figures

Figure 1

Figure 1
<p>Maps of Nyamata and Mayange sectors, Bugesera district, Eastern Province of Rwanda (Software: QGIS).</p>
Full article ">Figure 2
<p>Flowchart of extracting SRTM and ALOS PALSAR DEMS and derived DEMs from the handheld GPS tracks and points for the Study Area.</p>
Full article ">Figure 3
<p>Slope Line Segment Length (L) expressed in a graphical format.</p>
Full article ">Figure 4
<p>Bugesera <b>Income</b> Category in relation to Household Bicycle ownership (%) [<a href="#B27-urbansci-07-00008" class="html-bibr">27</a>] (Modelling method: histogram, Software: Stata).</p>
Full article ">Figure 5
<p>Bugesera Income in relation to available Work category (modelling method: Cross-Tabulation, Software: Stata) [<a href="#B27-urbansci-07-00008" class="html-bibr">27</a>].</p>
Full article ">Figure 6
<p>Map (<b>1</b>) represents the mobility patterns of GPS Tracks collected, Map (<b>2</b>) the mobility patterns of GPS Tracks collected in Nyamata and Mayange, while Map (<b>3</b>) represents the Overlaying Nya-mata and Mayange Mobility patterns in merged GPS Trackers, OSM and Rwanda road Network.</p>
Full article ">Figure 7
<p>Area of concentration of bicycle mobility patterns and activities in Nyamata and Mayange.</p>
Full article ">Figure 8
<p>Map (<b>1</b>) Rwanda road network data in the Bugesera, Map (<b>2</b>) Buffer analysis of Rwanda road networks, Map (<b>3</b>) Joint of Buffered Rwanda Roads Networks and merged GPS tracks.</p>
Full article ">Figure 9
<p>Spatial Joint of Nyamata GPS Tracks and Rwanda Road networks with buffer.</p>
Full article ">Figure 10
<p>Spatial Joint of Mayange GPS Tracks and Rwanda Road networks with buffer.</p>
Full article ">Figure 11
<p>Nyamata’s official roads and non-identified roads (non-official roads).</p>
Full article ">Figure 12
<p>Mayange’s four official roads and no identified roads (no official roads).</p>
Full article ">Figure 13
<p>ALOS PALSAR vs. SRTM elevation ranges of the GDEMs in the study.</p>
Full article ">Figure 14
<p>Slope percentages and Slope length of collected bicycle routes in Bugesera.</p>
Full article ">
20 pages, 6039 KiB  
Article
Synthesis and Characterization of α-Al2O3/Ba-β-Al2O3 Spheres for Cadmium Ions Removal from Aqueous Solutions
by Pamela Nair Silva-Holguín, Álvaro de Jesús Ruíz-Baltazar, Nahum Andrés Medellín-Castillo, Gladis Judith Labrada-Delgado and Simón Yobanny Reyes-López
Materials 2022, 15(19), 6809; https://doi.org/10.3390/ma15196809 - 30 Sep 2022
Cited by 4 | Viewed by 2262
Abstract
The search for adsorbent materials with a certain chemical inertness, mechanical resistance, and high adsorption capacity, as is the case with alumina, is carried out with structural or surface modifications with the addition of additives or metallic salts. This research shows the synthesis, [...] Read more.
The search for adsorbent materials with a certain chemical inertness, mechanical resistance, and high adsorption capacity, as is the case with alumina, is carried out with structural or surface modifications with the addition of additives or metallic salts. This research shows the synthesis, characterization, phase evolution and Cd(II) adsorbent capacity of α-Al2O3/Ba-β-Al2O3 spheres obtained from α-Al2O3 nanopowders by the ion encapsulation method. The formation of the Ba-β-Al2O3 phase is manifested at 1500 °C according to the infrared spectrum by the appearance of bands corresponding to AlO4 bonds and the appearance of peaks corresponding to Ba-O bonds in Raman spectroscopy. XRD determined the presence of BaO·Al2O3 at 1000 °C and the formation of Ba-β-Al2O3 at 1600 °C. Scanning electron microscopy revealed the presence of spherical grains corresponding to α-Al2O3 and hexagonal plates corresponding to β-Al2O3 in the spheres treated at 1600 °C. The spheres obtained have dimensions of 4.65 ± 0.30 mm in diameter, weight of 43 ± 2 mg and a surface area of 0.66 m2/g. According to the curve of pH vs. zeta potential, the spheres have an acid character and a negative surface charge of −30 mV at pH 5. Through adsorption studies, an adsorbent capacity of Cd(II) of 59.97 mg/g (87 ppm Cd(II)) was determined at pH 5, and the data were fitted to the pseudo first order, pseudo second order and Freundlich models, with correlation factors of 0.993, 0.987 and 0.998, respectively. Full article
(This article belongs to the Special Issue Surface Modification to Improve Properties of Material)
Show Figures

Figure 1

Figure 1
<p>Photographs of alumina spheres: (<b>A</b>) without thermal treatment, (<b>B</b>) with treatment at 1600 °C and (<b>C</b>) several spheres at 1600 °C in petri dish.</p>
Full article ">Figure 2
<p>Infrared spectra of the alumina powders and sphere treated at different temperatures: (<b>A</b>) powdered sample (P-Al<sub>2</sub>O<sub>3</sub>) and (<b>B</b>) sphere (S-Al<sub>2</sub>O<sub>3</sub>).</p>
Full article ">Figure 3
<p>Raman spectra of the alumina powders and sphere treated at different temperatures.</p>
Full article ">Figure 4
<p>DRX of the alumina powders and sphere treated at different temperatures.</p>
Full article ">Figure 5
<p>Scheme of synthesis mechanism of spinel Ba-β-Al<sub>2</sub>O<sub>3</sub> in the spheres.</p>
Full article ">Figure 6
<p>XRD diffractogram of powders and alumina sphere at 1600 °C.</p>
Full article ">Figure 7
<p>Crystalline systems and network parameters of the analyzed phases.</p>
Full article ">Figure 8
<p>SEM micrographs of alumina spheres treated at 1000 °C (<b>A</b>,<b>B</b>) and 1600 °C (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 9
<p>N<sub>2</sub> adsorption–desorption isotherm: (<b>A</b>) Alumina sphere treated at 1000 °C and (<b>B</b>) Alumina sphere at 1600 °C.</p>
Full article ">Figure 10
<p>SEM micrographs of non-fragmented (<b>A</b>–<b>C</b>) and fragmented (<b>D</b>–<b>F</b>) alumina spheres treated at 1600 °C.</p>
Full article ">Figure 11
<p>SEM micrographs of surface of alumina spheres treated at 1600 °C: (<b>A</b>) 500×, (<b>B</b>) 1000×, (<b>C</b>) 3000× and (<b>D</b>) 6000×.</p>
Full article ">Figure 12
<p>Microelemental surface analysis of alumina spheres treated at 1600 °C.</p>
Full article ">Figure 13
<p>Microelemental mapping analysis surface of alumina spheres treated at 1600 °C. (<b>A</b>) Micrography, (<b>B</b>) Oxygen, (<b>C</b>) Aluminum and (<b>D</b>) Barium.</p>
Full article ">Figure 14
<p>Curves of pH vs. zeta potential of powders and alumina sphere. (<b>A</b>) P-Al<sub>2</sub>O<sub>3</sub> at 100 °C, (<b>B</b>) P-Al<sub>2</sub>O<sub>3</sub> at 1000 °C, (<b>C</b>) P-Al<sub>2</sub>O<sub>3</sub> 1600 °C and (<b>D</b>) S-Al<sub>2</sub>O<sub>3</sub> at 1600 °C.</p>
Full article ">Figure 15
<p>Adsorption of Cd(II) on alumina spheres at 1600 °C: (<b>A</b>) Adsorption kinetic models, (<b>B</b>) Adsorption equilibrium isotherm.</p>
Full article ">Figure 16
<p>SEM surface micrographs of alumina spheres (<b>A</b>,<b>B</b>) before adsorption and (<b>C</b>,<b>D</b>) after Cd(II) adsorption.</p>
Full article ">Figure 17
<p>Infrared spectrum of S-Al<sub>2</sub>O<sub>3</sub> 1600 °C and spheres after adsorbing cadmium.</p>
Full article ">
27 pages, 15649 KiB  
Article
Detection, Morphometric Analysis and Digital Surveying of Archaeological Mounds in Southern Iraq with CartoSat-1 and COSMO-SkyMed DEMs
by Deodato Tapete and Francesca Cigna
Land 2022, 11(9), 1406; https://doi.org/10.3390/land11091406 - 27 Aug 2022
Cited by 4 | Viewed by 2823
Abstract
In Near and Middle Eastern archaeology, satellite-derived digital elevation models (DEM) of medium spatial resolution (≥30 m) are mostly used to locate and map archaeological mounds (namely ‘tells’), whereas high resolution DEMs (≤10 m) are still poorly exploited. To fill this gap, the [...] Read more.
In Near and Middle Eastern archaeology, satellite-derived digital elevation models (DEM) of medium spatial resolution (≥30 m) are mostly used to locate and map archaeological mounds (namely ‘tells’), whereas high resolution DEMs (≤10 m) are still poorly exploited. To fill this gap, the 5 m resolution CartoSat-1 Euro-Maps 3D Digital Surface Model (DSM) is assessed vs. the 30 m Shuttle Radar Topography Mission (SRTM) global DEM, the Advanced Land Observing Satellite (ALOS) World 3D–30 m (AW3D30) and a 10 m COSMO-SkyMed DEM, on a test area in Wasit, southern Iraq, where the high density of tells is yet to be exhaustively documented. A total of 344 sites was mapped, with one order of magnitude improvement compared to previous mapping exercises, existing databases and historical maps. The morphometric analysis not only highlights the reliability of CartoSat-1 DSM height and volume estimates, but also suggests that, in the test area, the volume of a tell can robustly be calculated based on the simple knowledge of its basal area, following a quadratic function. Morphology and elevation of at least 53% irregularly shaped tells were impacted by anthropogenic disturbances. Morphometric indices (e.g., Topographic Position Index, DEViation from mean elevation) are a viable automated method to ease tells detection. When integrated with other satellite datasets (e.g., CORONA, Google Earth, Sentinel-2 imagery), the CartoSat-1 DSM can unveil morphological changes and support condition assessment. In Wasit, agriculture and modern development are among the major threats for tells preservation, alongside looting. Full article
(This article belongs to the Special Issue Landscape Archaeology by Using Remote Sensing Data)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of Al-Ahrar, Al-Nu’maniya, Wasit (Iraq) and indication of the zoomed area in subfigures (<b>b</b>,<b>c</b>). (<b>b</b>) Land cover according to the WorldCover 10 m dataset (©ESA WorldCover project 2020/Contains modified Copernicus Sentinel data 2020, processed by ESA WorldCover consortium [<a href="#B20-land-11-01406" class="html-bibr">20</a>]). (<b>c</b>) Surface elevation (<span class="html-italic">h<sub>C</sub></span><sub>1</sub>) above reference WGS84 ellipsoid, derived from CartoSat-1 Euro-Maps 3D Digital Surface Model (DSM) product (©GAF AG, Munich, Germany. Includes material ©Antrix, Bangalore, India, distributed by GAF AG), overlapped onto high resolution satellite imagery (contains Copernicus Sentinel-2 data, 2019).</p>
Full article ">Figure 2
<p>Sketch of archaeological tells and identification of their main dimensional parameters: planar view of (<b>a</b>) irregular and (<b>b</b>) elliptical tells, and (<b>c</b>) cross-section. Notation: <span class="html-italic">F<sub>max</sub></span> and <span class="html-italic">F<sub>min</sub></span>, maximum and minimum Feret diameters; <span class="html-italic">H</span>, height; <span class="html-italic">h<sub>max</sub></span> and <span class="html-italic">h<sub>min</sub></span>, top and base elevation, respectively.</p>
Full article ">Figure 3
<p>Zoomed view of the high resolution CartoSat-1 DSM product (©GAF AG. Includes material ©Antrix, distributed by GAF AG) enhancing the anthropogenic topographic relieves of Tūlūl al-Baqarat tells from the surrounding agricultural fields in the eastern sector of the test area (see location in <a href="#land-11-01406-f001" class="html-fig">Figure 1</a>c).</p>
Full article ">Figure 4
<p>Distribution of the observed elevation differences (Δ<span class="html-italic">h<sub>i</sub></span>) between the CartoSat-1 DSM and the three reference DEMs: (<b>a</b>) COSMO-SkyMed, (<b>b</b>) SRTM 30 m and (<b>c</b>) AW3D30.</p>
Full article ">Figure 5
<p>Spatial distribution of the mapped tells within the test area, onto high resolution satellite imagery (contains Copernicus Sentinel-2 data, 2019), compared to (<b>a</b>) Ancient Near East (ANE) sites as mapped by [<a href="#B23-land-11-01406" class="html-bibr">23</a>] and (<b>b</b>) tells investigated and reported by [<a href="#B49-land-11-01406" class="html-bibr">49</a>].</p>
Full article ">Figure 6
<p>Relationship between (<b>a</b>) maximum diameter (<span class="html-italic">d<sub>max</sub></span>) and area (<span class="html-italic">A</span>), and (<b>b</b>) minimum (<span class="html-italic">F<sub>min</sub></span>) and maximum (<span class="html-italic">F<sub>max</sub></span>) Feret diameters of the 344 tells mapped within the test area.</p>
Full article ">Figure 7
<p>Distribution of height (<span class="html-italic">H</span>) values estimated from CartoSat-1 DSM dataset vs. the reference datasets: COSMO-SkyMed, SRTM 30 m and AW3D30.</p>
Full article ">Figure 8
<p>Comparison of (<b>a</b>–<b>c</b>) height (<span class="html-italic">H<sub>i</sub></span>) and (<b>d</b>–<b>f</b>) volume (<span class="html-italic">V<sub>i</sub></span>) estimates between CartoSat-1 DSM (<span class="html-italic">i</span> = C1) and the three reference DEM datasets: COSMO-SkyMed (<span class="html-italic">i</span> = CSK), SRTM 30 m (<span class="html-italic">i</span> = SRTM) and AW3D30 (<span class="html-italic">i</span> = AW3D30).</p>
Full article ">Figure 9
<p>Relationships between height (<span class="html-italic">H<sub>i</sub></span>), area (<span class="html-italic">A</span>) and volume (<span class="html-italic">V<sub>i</sub></span>), for: (<b>a</b>–<b>c</b>) CartoSat-1 DSM, and (<b>d</b>–<b>f</b>) COSMO-SkyMed, (<b>g</b>–<b>i</b>) SRTM and (<b>j</b>–<b>l</b>) AW3D30 DEMs.</p>
Full article ">Figure 10
<p>Relationship between maximum diameter (<span class="html-italic">d<sub>max</sub></span>) and height (<span class="html-italic">H<sub>C</sub></span><sub>1</sub>) estimated from CartoSat-1 DSM, for the whole sample of tells mapped within the test area.</p>
Full article ">Figure 11
<p>Spatial distribution of the mapped tells by the mostly impacting threats due to anthropogenic actions: (<b>a</b>) reworking and reshaping, (<b>b</b>) leveling and flattening, (<b>c</b>) looting, (<b>d</b>) agriculture, (<b>e</b>) building construction, (<b>f</b>) fire, (<b>g</b>) modern canals, (<b>h</b>) road, (<b>i</b>) disturbance by heavy vehicles, and (<b>j</b>) change of use.</p>
Full article ">Figure 12
<p>Example of tell (precise location undisclosed for security concerns) affected by ‘reworking and reshaping’ and ‘looting’ incidents, as documented by multi-temporal analysis of: (<b>a</b>) CORONA imagery, (<b>b</b>) CartoSat-1 DSM (©GAF AG. Includes material ©Antrix, distributed by GAF AG) and (<b>c</b>) COSMO-SkyMed DEM (COSMO-SkyMed<sup>®®</sup> Products ©ASI, Italian Space Agency, Rome, Italy, 2018. All rights reserved) products, as well as contextual information from Google Earth images (©2021 Maxar Technologies, Westminster, CO, USA) collected on: (<b>d</b>) 18 January 2015, (<b>e</b>) 7 January 2018 and (<b>f</b>) 26 July 2018. (<b>g</b>,<b>h</b>) Insets highlight the presence of looting pits. White and red arrows, respectively, point to trenches and the southern portion of the tell that was ultimately leveled and converted to crop field.</p>
Full article ">Figure 13
<p>Example of tells (precise location undisclosed for security concerns) where ‘leveling and flattening’ led to conversion to crop fields and further damage due to ‘agriculture’, as documented by multi-temporal analysis of (<b>a</b>) CORONA imagery and (<b>b</b>) CartoSat-1 DSM (©GAF AG. Includes material ©Antrix, distributed by GAF AG), and contextual information from (<b>c</b>) Google Earth image (©2021 Maxar Technologies) collected on 9 April 2002.</p>
Full article ">Figure 14
<p>Example of tells (labeled “i” and “ii”; precise location undisclosed for security concerns) affected by ‘change of use’ and ‘modern canals’, respectively, as documented by multi-temporal analysis of: (<b>a</b>) CORONA imagery and (<b>b</b>) CartoSat-1 DSM (©GAF AG. Includes material ©Antrix, distributed by GAF AG), as well as contextual information from Google Earth images (©2021 Maxar Technologies) collected on: (<b>c</b>) 9 April 2002, (<b>d</b>) 5 January 2016 and (<b>e</b>) 22 December 2020, with (<b>f</b>–<b>h</b>) zoomed views highlighting the transformation of tell “i” and its change of use in connection to very recently built farm dwellings.</p>
Full article ">Figure 15
<p>(<b>a</b>) Overview of the tells located in the eastern part of the test area (see also <a href="#land-11-01406-f011" class="html-fig">Figure 11</a>f; precise location undisclosed for security concerns) that were affected by fire incidents, with indication of (<b>b</b>–<b>e</b>) zoomed views highlighting the burning patterns and scars, as captured by Google Earth image (©2021 Maxar Technologies) collected on 7 October 2013.</p>
Full article ">Figure 16
<p>Zoomed view of the northernmost part of the test area as depicted in (<b>a</b>) CORONA imagery, (<b>b</b>) Sentinel-2 false-colored image (contains Copernicus Sentinel-2 data, 2019), (<b>c</b>) CartoSat-1 DSM (©GAF AG. Includes material ©Antrix, distributed by GAF AG), and its derivatives: (<b>d</b>) Topographic Position Index (<span class="html-italic">TPI</span>) and (<b>e</b>) DEViation from mean elevation (<span class="html-italic">DEV</span>).</p>
Full article ">Figure 17
<p>Zoomed views of tells (<b>a</b>,<b>b</b>) Bismaya H882 and (<b>c</b>–<b>e</b>) Tūlūl al-Baqarat TB1 and TB6 showcasing the benefits of ground-truth and archaeological interpretation to verify, and even revise/refine, tell boundaries that are drawn via desk-based assessment of features and topographic relieves visible in CORONA imagery and a high resolution DEM product. Pictures (<b>b</b>,<b>d</b>) contain the CartoSat-1 DSM product (©GAF AG. Includes material ©Antrix, distributed by GAF AG); picture (<b>e</b>) contains Google Earth image (©2021 Maxar Technologies) collected on 3 October 2020.</p>
Full article ">Figure 18
<p>Validation of height (<span class="html-italic">H<sub>i</sub></span>) values derived from (<b>a</b>) CartoSat-1 DSM and (<b>b</b>) COSMO-SkyMed DEM against in situ measurements. The latter are according to the values in [<a href="#B49-land-11-01406" class="html-bibr">49</a>].</p>
Full article ">
19 pages, 8105 KiB  
Article
Molecular Docking and Efficacy of Aloe vera Gel Based on Chitosan Nanoparticles against Helicobacter pylori and Its Antioxidant and Anti-Inflammatory Activities
by Reham Yahya, Aisha M. H. Al-Rajhi, Saleh Zaid Alzaid, Mohamed A. Al Abboud, Mohammed S. Almuhayawi, Soad K. Al Jaouni, Samy Selim, Khatib Sayeed Ismail and Tarek M. Abdelghany
Polymers 2022, 14(15), 2994; https://doi.org/10.3390/polym14152994 - 24 Jul 2022
Cited by 61 | Viewed by 5561
Abstract
The medicinal administration of Aloe vera gel has become promising in pharmaceutical and cosmetic applications particularly with the development of the nanotechnology concept. Nowadays, effective H. pylori treatment is a global problem; therefore, the development of natural products with nanopolymers such as chitosan [...] Read more.
The medicinal administration of Aloe vera gel has become promising in pharmaceutical and cosmetic applications particularly with the development of the nanotechnology concept. Nowadays, effective H. pylori treatment is a global problem; therefore, the development of natural products with nanopolymers such as chitosan nanoparticles (CSNPs) could represent a novel strategy for the treatment of gastric infection of H. pylori. HPLC analysis of A. vera gel indicated the presence of chlorogenic acid as the main constituent (1637.09 µg/mL) with other compounds pyrocatechol (1637.09 µg/mL), catechin (1552.92 µg/mL), naringenin (528.78 µg/mL), rutin (194.39 µg/mL), quercetin (295.25 µg/mL), and cinnamic acid (37.50 µg/mL). CSNPs and A. vera gel incorporated with CSNPs were examined via TEM, indicating mean sizes of 83.46 nm and 36.54 nm, respectively. FTIR spectra showed various and different functional groups in CSNPs, A. vera gel, and A. vera gel incorporated with CSNPs. Two strains of H. pylori were inhibited using A. vera gel with inhibition zones of 16 and 16.5 mm, while A. vera gel incorporated with CSNPs exhibited the highest inhibition zones of 28 and 30 nm with resistant and sensitive strains, respectively. The minimal inhibitory concentration (MIC) was 15.62 and 3.9 µg/mL, while the minimal bactericidal concentration (MBC) was 15.60 and 7.8 µg/mL with MBC/MIC 1 and 2 indexes using A. vera gel and A. vera gel incorporated with CSNPs, respectively, against the resistance strain. DPPH Scavenging (%) of the antioxidant activity exhibited an IC50 of 138.82 μg/mL using A.vera gel extract, and 81.7 μg/mL when A.vera gel was incorporated with CSNPs. A.vera gel incorporated with CSNPs enhanced the hemolysis inhibition (%) compared to using A.vera gel alone. Molecular docking studies through the interaction of chlorogenic acid and pyrocatechol as the main components of A. vera gel and CSNPs with the crystal structure of the H. pylori (4HI0) protein supported the results of anti-H. pylori activity. Full article
(This article belongs to the Special Issue Biomedical Applications of Polymer-Based Nanomaterials)
Show Figures

Figure 1

Figure 1
<p><span class="html-italic">A. vera</span> leaf showing outer green rind and viscous clear liquid (gel).</p>
Full article ">Figure 2
<p>HPLC chromatograms of detected flavonoids and phenolic acids content of <span class="html-italic">A. vera</span> gel extract.</p>
Full article ">Figure 3
<p>UV–vis spectra of CSNPs and <span class="html-italic">A. vera</span> gel incorporated with CSNPs.</p>
Full article ">Figure 4
<p>TEM of synthesized <span class="html-italic">A. vera</span> gel incorporated with CSNPs (<b>A</b>), CSNPs (<b>B</b>), and diameter of some particles (<b>C</b>).</p>
Full article ">Figure 5
<p>FTIR spectra of <span class="html-italic">A. vera</span> gel.</p>
Full article ">Figure 6
<p>FTIR spectra of CSNPs.</p>
Full article ">Figure 7
<p>FTIR spectra of <span class="html-italic">A. vera</span> gel incorporated to CSNPs.</p>
Full article ">Figure 8
<p>Inhibitory action of <span class="html-italic">A. vera</span> gel incorporated with CSNPs (1), <span class="html-italic">A. vera</span> gel (2), positive control (3), CSNPs (4), and acetic acid (5) against resistant strain (<b>A</b>) and sensitive strain (<b>B</b>) of <span class="html-italic">Helicobacter pylori</span>.</p>
Full article ">Figure 9
<p>Antioxidant activity of <span class="html-italic">A. vera</span> gel and <span class="html-italic">A. vera</span> gel incorporated with CSNPs.</p>
Full article ">Figure 10
<p>Hemolysis inhibition (%) of <span class="html-italic">A. vera</span> gel, <span class="html-italic">A. vera</span> gel incorporated with CSNPs, and indomethacin.</p>
Full article ">Figure 11
<p>Docking interactions of certain compounds of <span class="html-italic">A. vera</span> gel extract.2D and 3D diagrams show the interaction between chlorogenic acid and active sites of 4HI0 protein (<b>A</b>); 2D and 3D diagrams show the interaction between chitosan and active sites of 4HI0 protein (<b>B</b>); 2D and 3D diagrams show the interaction between pyrocatechol and active sites of 4HI0 protein (<b>C</b>).</p>
Full article ">Figure 12
<p>The representative key for the types of interaction between chitosan and pyrocatechol.</p>
Full article ">
18 pages, 3266 KiB  
Article
Spatial Estimates of Flood Damage and Risk Are Influenced by the Underpinning DEM Resolution: A Case Study in Kuala Lumpur, Malaysia
by Eva Fatdillah, Balqis M. Rehan, Ponnambalam Rameshwaran, Victoria A. Bell, Zed Zulkafli, Badronnisa Yusuf and Paul Sayers
Water 2022, 14(14), 2208; https://doi.org/10.3390/w14142208 - 13 Jul 2022
Cited by 12 | Viewed by 3693
Abstract
The sensitivity of simulated flood depth and area to DEM resolution are acknowledged, but their effects on flood damage and risk estimates are less well understood. This study sought to analyse the relative benefits of using global DEMs of different resolution sizes, 5 [...] Read more.
The sensitivity of simulated flood depth and area to DEM resolution are acknowledged, but their effects on flood damage and risk estimates are less well understood. This study sought to analyse the relative benefits of using global DEMs of different resolution sizes, 5 m AW3D Standard, 12.5 m ALOS PALSAR and 30 m SRTM, to simulate flood inundation, damage and risk. The HEC-RAS 2D model was adopted for flood simulations, and the Toba River in the Klang River Basin in Malaysia was chosen for the case study. Simulated inundation areas from AW3D coincide the most with reported flooded areas, but the coarser-resolution DEMs did capture some of the reported flooded areas. The inundation area increased as the resolution got finer. As a result, AW3D returned almost double flood damage and risk estimates compared to ALOS PALSAR, and almost quadruple compared to SRTM for building-level damage and risk analysis. The findings indicate that a finer-resolution DEM improves inundation modelling and could provide greater flood damage and risk estimates compared to a coarser DEM. However, DEMs of coarser resolution remain useful in data-scarce regions or for large-scale assessments in efforts to manage flood risk. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>The Toba Catchment. (<b>a</b>) Reported flood locations with some locations having multiple occurrences, and (<b>b</b>) land use map.</p>
Full article ">Figure 2
<p>Conceptual workflow for this study. The workflow was repeated for three DEMs with different spatial resolutions.</p>
Full article ">Figure 3
<p>Maps of before and after modification of DEM terrain in Toba Catchment: (<b>a</b>) AW3D Standard, (<b>b</b>) ALOS PALSAR and (<b>c</b>) SRTM.</p>
Full article ">Figure 4
<p>Peak flow–return period relationship of the Toba River.</p>
Full article ">Figure 5
<p>Distribution of flood depths across the floodplain areas of AW3D Standard, ALOS PALSAR and SRTM. Results from all flood events are shown. The <span class="html-italic">y</span>-axis is the percentage of inundation area over the total inundation area for each considered flood depth.</p>
Full article ">Figure 6
<p>River cross-sections at three different locations: D = upstream, E = midstream and F = downstream. Black dots in the cross-sections are referring to the riverbank. Flood maps of simulated flooded area and depths from 100-year return period event for (<b>a</b>) AW3D Standard (<b>b</b>) ALOS PALSAR and (<b>c</b>) SRTM. Reported flooded locations are in red dots.</p>
Full article ">Figure 7
<p>Flood damage estimates for AW3D Standard, ALOS PALSAR and SRTM with interpolation lines. Bullet dots refer to the total damage for each flood event.</p>
Full article ">Figure 8
<p>The output comparisons of AW3D Standard, ALOS PALSAR and SRTM.</p>
Full article ">
28 pages, 9175 KiB  
Article
Improved k-NN Mapping of Forest Attributes in Northern Canada Using Spaceborne L-Band SAR, Multispectral and LiDAR Data
by André Beaudoin, Ronald J. Hall, Guillermo Castilla, Michelle Filiatrault, Philippe Villemaire, Rob Skakun and Luc Guindon
Remote Sens. 2022, 14(5), 1181; https://doi.org/10.3390/rs14051181 - 27 Feb 2022
Cited by 13 | Viewed by 3315
Abstract
Satellite forest inventories are the only feasible way to map Canada’s vast, remote forest regions, such as those in the Northwest Territories (NWT). A method used to create such inventories is the k-nearest neighbour (k-NN) algorithm, which spatially extends information [...] Read more.
Satellite forest inventories are the only feasible way to map Canada’s vast, remote forest regions, such as those in the Northwest Territories (NWT). A method used to create such inventories is the k-nearest neighbour (k-NN) algorithm, which spatially extends information from forest inventory (FI) plots to the entire forest land base using wall-to-wall features typically derived from Landsat data. However, the benefits of integrating L-band synthetic aperture radar (SAR) data, strongly correlated to forest biomass, have not been assessed for Canadian northern boreal forests. Here we describe an optimized multivariate k-NN implementation of a 151,700 km2 area in southern NWT that included ca. 2007 Landsat and dual-polarized Phased Array type L-band SAR (PALSAR) data on board the Advanced Land Observing Satellite (ALOS). Five forest attributes were mapped at 30 m cells: stand height, crown closure, stand/total volume and aboveground biomass (AGB). We assessed accuracy gains compared to Landsat-based maps. To circumvent the scarcity of FI plots, we used 3600 footprints from the Geoscience Laser Altimeter System (GLAS) as surrogate FI plots, where forest attributes were estimated using Light Detection and Ranging (LiDAR) metrics as predictors. After optimization, k-NN predicted forest attribute values for each pixel as the average of the 4 nearest (k = 4) surrogate FI plots within the Euclidian space of 9 best features (selected among 6 PALSAR, 10 Landsat, and 6 environmental features). Accuracy comparisons were based on 31 National Forest Inventory ground plots and over 1 million airborne LiDAR plots. Maps that included PALSAR HV backscatter resulted in forest attribute predictions with higher goodness of fit (adj. R2), lower percent mean error (ME%), and percent root mean square error (RMSE%), and lower underestimation for larger attribute values. Predictions were most accurate for conifer stand height (RMSE% = 32.1%, adj. R2 = 0.58) and AGB (RMSE% = 47.8%, adj. R2 = 0.74), which is much more abundant in the area than mixedwood or broadleaf. Our study demonstrates that optimizing k-NN parameters and feature space, including PALSAR, Landsat, and environmental variables, is a viable approach for inventory mapping of the northern boreal forest regions of Canada. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>MVI phase 1 study area (red outline) within a broader area across 2 provinces and 2 territories (separated by thin black outline) that has as a backdrop a ca. 2007 landcover map that includes forest cover types (C: conifer; M: mixedwood; B: broadleaf) with 3 density classes (sparse, open, dense) along with the Geoscience Laser Altimeter System (GLAS) reference dataset of surrogate forest inventory (FI) plots and 2 validation sample sets. The top right zoomed-in inset shows a single GLAS FI plot surrounded by BT−ALS plots in a 500 m by 500 m area corresponding to an intersection between the BT−ALS transect and an ICESat track. Map is in Albers equal area conic projection.</p>
Full article ">Figure 2
<p>Percent occurrence of forest cover types (WT: wetland treed; C: conifer; M: mixedwood; B: broadleaf) with cover density classes (sparse, open, dense) across all forested pixels of the study area and the initial and final GLAS samples of surrogate forest inventory FI plots, respectively, according to the ca. 2007 landcover map (<a href="#remotesensing-14-01181-f001" class="html-fig">Figure 1</a>).</p>
Full article ">Figure 3
<p><span class="html-italic">k</span>-NN optimization and mapping workflow to generate the Satellite Vegetation Inventory (SVI) raster maps of five forest attributes and SVI map comparative accuracy assessment using Landsat-based map version (SVI_L) and published (PUB) maps. Numbers in brackets refer to related sections in the article.</p>
Full article ">Figure 4
<p>Multivariate global root mean square difference (GRMSD) metric (left Y axis) and the number of times a particular feature was selected in univariate feature selection across five attributes (right Y axis) for (<b>a</b>) initial selection features among the 20 candidate features and (<b>b</b>) the adjusted final selection of nine features (marked as * in panel (<b>a</b>)).</p>
Full article ">Figure 5
<p>Percent statistic values relative to their optimal values (100%) across a range of <span class="html-italic">k</span> values (rel<sub>stat_opt</sub>) for pseudo-R<sup>2</sup> (T<sup>2</sup>, Equation (2)), root mean square difference (RMSD, Equation (3)), and mean difference using the lower and upper 5% of distribution (MD<sub>5</sub>, MD<sub>95,</sub> Equations (5) and (6)), supporting the selection of the optimal <span class="html-italic">k</span> value of 4.</p>
Full article ">Figure 6
<p>SVI raster maps from <span class="html-italic">k</span>-NN predictions of (<b>a</b>) stand height and (<b>b</b>) AGB for the Phase 1 area. White pixels are non-forested lands, whereas light blue pixels are water bodies. Low and high attribute values are the 5% and 95% percentile, respectively. SVI maps for all five forest attributes are found in <a href="#app1-remotesensing-14-01181" class="html-app">Figure S2</a>.</p>
Full article ">Figure 6 Cont.
<p>SVI raster maps from <span class="html-italic">k</span>-NN predictions of (<b>a</b>) stand height and (<b>b</b>) AGB for the Phase 1 area. White pixels are non-forested lands, whereas light blue pixels are water bodies. Low and high attribute values are the 5% and 95% percentile, respectively. SVI maps for all five forest attributes are found in <a href="#app1-remotesensing-14-01181" class="html-app">Figure S2</a>.</p>
Full article ">Figure 7
<p>Comparison of scatterplots of observations versus predictions of stand height (left column) and AGB (right column) from (<b>a</b>) SVI maps, (<b>b</b>) Landsat-based SVI_L maps and (<b>c</b>) previously published Landsat-based maps using all NFI plots (blue dots) and BT−ALS LiDAR plots (density scatterplot). Dashed blue and black lines are regression lines along with equations and adj. R<sup>2</sup> values, respectively, based on NFI plots and BT−ALS LiDAR plots (see <a href="#app1-remotesensing-14-01181" class="html-app">Table S3</a> for linear regression statistics of all attributes). Distinctive symbols for the NFI plots (blue dots) distinguish the three forest cover types.</p>
Full article ">Figure 8
<p>Goodness of fit (adj. R<sup>2</sup>) for (<b>a</b>) stand height and (<b>b</b>) AGB relative to the validation datasets comprising NFI plots and BT−ALS LiDAR plots using all samples (ALL) then partitioned by forest cover type (C: conifer, M: mixedwood, B: broadleaf) for SVI maps compared to Landsat-based SVI_L maps and published (PUB) maps.</p>
Full article ">Figure 9
<p>Percent mean error (ME%) for (<b>a</b>) stand height and (<b>b</b>) AGB relative to the validation datasets comprising NFI plots and BT−ALS LiDAR plots using all samples (ALL) then partitioned by forest cover type (C: conifer, M: mixedwood, B: broadleaf) for SVI maps compared to Landsat-based SVI_L maps and published (PUB) maps.</p>
Full article ">Figure 10
<p>Percent root mean square error (RMSE%) for (<b>a</b>) stand height and (<b>b</b>) AGB relative to the validation datasets comprising NFI plots and BT−ALS LiDAR plots using all samples (ALL) then partitioned by forest cover type (C: conifer, M: mixedwood, B: broadleaf) for SVI maps compared to Landsat-based SVI_L maps and published (PUB) maps.</p>
Full article ">Figure 11
<p>Plot of mean prediction error (predicted minus observed) for (<b>a</b>) stand height and (<b>b</b>) AGB using all NFI plots (horizontal lines) and NFI plots grouped by quartiles Q1 to Q4 (mean error ± one standard deviation, dots) for SVI maps, Landsat-based SVI_L maps and published (PUB) maps. Dotted lines are added to highlight trends across four quartiles.</p>
Full article ">
11 pages, 371 KiB  
Article
In Vitro Antileishmanial and Antitrypanosomal Activities of Plicataloside Isolated from the Leaf Latex of Aloe rugosifolia Gilbert & Sebsebe (Asphodelaceae)
by Gete Chemeda, Daniel Bisrat, Mariamawit Y. Yeshak and Kaleab Asres
Molecules 2022, 27(4), 1400; https://doi.org/10.3390/molecules27041400 - 18 Feb 2022
Cited by 8 | Viewed by 1884
Abstract
Trypanosomiasis and leishmaniasis are among the major neglected diseases that affect poor people, mainly in developing countries. In Ethiopia, the latex of Aloe rugosifolia Gilbert & Sebsebe is traditionally used for the treatment of protozoal diseases, among others. In this study, the in [...] Read more.
Trypanosomiasis and leishmaniasis are among the major neglected diseases that affect poor people, mainly in developing countries. In Ethiopia, the latex of Aloe rugosifolia Gilbert & Sebsebe is traditionally used for the treatment of protozoal diseases, among others. In this study, the in vitro antitrypanosomal activity of the leaf latex of A. rugosifolia was evaluated against Trypanosoma congolense field isolate using in vitro motility and in vivo infectivity tests. The latex was also tested against the promastigotes of Leishmania aethiopica and L. donovani clinical isolates using alamar blue assay. Preparative thin-layer chromatography of the latex afforded a naphthalene derivative identified as plicataloside (2,8-O,O-di-(β-D-glucopyranosyl)-1,2,8-trihydroxy-3-methyl-naphthalene) by means of spectroscopic techniques (HRESI-MS, 1H, 13C-NMR). Results of the study demonstrated that at 4.0 mg/mL concentration plicataloside arrested mobility of trypanosomes within 30 min of incubation period. Furthermore, plicataloside completely eliminated subsequent infectivity in mice for 30 days at concentrations of 4.0 and 2.0 mg/mL. Plicataloside also displayed antileishmanial activity against the promastigotes of L. aethopica and L. donovani with IC50 values 14.22 ± 0.41 µg/mL (27.66 ± 0.80 µM) and 18.86 ± 0.03 µg/mL (36.69 ± 0.06 µM), respectively. Thus, plicataloside may be used as a scaffold for the development of novel drugs effective against trypanosomiasis and leishmaniasis. Full article
(This article belongs to the Section Natural Products Chemistry)
Show Figures

Figure 1

Figure 1
<p>Structural formula of plicataloside.</p>
Full article ">
20 pages, 392 KiB  
Article
Improvement of Selected Morphological, Physiological, and Biochemical Parameters of Roselle (Hibiscus sabdariffa L.) Grown under Different Salinity Levels Using Potassium Silicate and Aloe saponaria Extract
by Alaa Idris Badawy Abou-Sreea, Mohamed H. H. Roby, Hayam A. A. Mahdy, Nasr M. Abdou, Amira M. El-Tahan, Mohamed T. El-Saadony, Khaled A. El-Tarabily and Fathy M. A. El-Saadony
Plants 2022, 11(4), 497; https://doi.org/10.3390/plants11040497 - 11 Feb 2022
Cited by 15 | Viewed by 3603
Abstract
Two successive field trials were carried out at the experimental farm of the Agriculture Department of Fayoum University, Fayoum, Egypt, to investigate the sole or dual interaction effect of applying a foliar spray of Aloe saponaria extract (Ae) or potassium silicate (KSi) on [...] Read more.
Two successive field trials were carried out at the experimental farm of the Agriculture Department of Fayoum University, Fayoum, Egypt, to investigate the sole or dual interaction effect of applying a foliar spray of Aloe saponaria extract (Ae) or potassium silicate (KSi) on reducing the stressful salinity impacts on the development, yield, and features of roselle (Hibiscus sabdariffa L.) plants. Both Ae or KSi were used at three rates: 0% (0 cm3 L−1), 0.5% (5 cm3 L−1), and 1% (10 cm3 L−1) and 0, 30, and 60 g L−1, respectively. Three rates of salinity, measured by the electrical conductivity of a saturated soil extract (ECe), were also used: normal soil (ECe < 4 dS/m) (S1); moderately-saline soil (ECe: 4–8 dS/m) (S2); and highly-saline soil (ECe: 8–16 dS/m) (S3). The lowest level of salinity yielded the highest levels of all traits except for pH, chloride, and sodium. Ae at 0.5% increased the values of total soluble sugars, total free amino acids, potassium, anthocyanin, a single-photon avalanche diode, stem diameter, fruit number, and fresh weight, whereas 1% of Ae resulted in the highest plant height, chlorophyll fluorescence (Fv/Fm), performance index, relative water content, membrane stability index, proline, total soluble sugars, and acidity. KSi either at 30 or 60 g L−1 greatly increased these abovementioned attributes. Fruit number and fruit fresh weight per plant also increased significantly with the combination of Ae at 1% and KSi at 30 g L−1 under normal soil conditions. Full article
29 pages, 25750 KiB  
Article
Regional-Scale Systematic Mapping of Archaeological Mounds and Detection of Looting Using COSMO-SkyMed High Resolution DEM and Satellite Imagery
by Deodato Tapete, Arianna Traviglia, Eleonora Delpozzo and Francesca Cigna
Remote Sens. 2021, 13(16), 3106; https://doi.org/10.3390/rs13163106 - 6 Aug 2021
Cited by 19 | Viewed by 5333
Abstract
“Tells” are archaeological mounds formed by deposition of large amounts of anthropogenic material and sediments over thousands of years and are the most important and prominent features in Near and Middle Eastern archaeological landscapes. In the last decade, archaeologists have exploited free-access global [...] Read more.
“Tells” are archaeological mounds formed by deposition of large amounts of anthropogenic material and sediments over thousands of years and are the most important and prominent features in Near and Middle Eastern archaeological landscapes. In the last decade, archaeologists have exploited free-access global digital elevation model (DEM) datasets at medium resolution (i.e., up to 30 m) to map tells on a supra-regional scale and pinpoint tentative tell sites. Instead, the potential of satellite DEMs at higher resolution for this task was yet to be demonstrated. To this purpose, the 3 m resolution imaging capability allowed by the Italian Space Agency’s COSMO-SkyMed Synthetic Aperture Radar (SAR) constellation in StripMap HIMAGE mode was used in this study to generate DEM products of enhanced resolution to undertake, for the first time, a systematic mapping of tells and archaeological deposits. The demonstration is run at regional scale in the Governorate of Wasit in central Iraq, where the literature suggested a high density of sites, despite knowledge gaps about their location and spatial distribution. Accuracy assessment of the COSMO-SkyMed DEM is provided with respect to the most commonly used SRTM and ALOS World 3D DEMs. Owing to the 10 m posting and the consequent enhanced observation capability, the COSMO-SkyMed DEM proves capable to detect both well preserved and levelled or disturbed tells, standing out for more than 4 m from the surrounding landscape. Through the integration with CORONA KH-4B tiles, 1950s Soviet maps and recent Sentinel-2 multispectral images, the expert-led visual identification and manual mapping in the GIS environment led to localization of tens of sites that were not previously mapped, alongside the computation of a figure as up-to-date as February 2019 of the survived tells, with those affected by looting. Finally, this evidence is used to recognize hot-spot areas of potential concern for the conservation of tells. To this purpose, we upgraded the spatial resolution of the observations up to 1 m by using the Enhanced Spotlight mode to collect a bespoke time series. The change detection tests undertaken on selected clusters of disturbed tells prove how a dedicated monitoring activity may allow a regular observation of the impacts due to anthropogenic disturbance (e.g., road and canal constructions or ploughing). Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of Wasit study area in central Iraq and spatial extent of: the bespoke collection of COSMO-SkyMed StripMap (SM) HIMAGE acquisitions from which the digital elevation model (DEM) was generated; CORONA imagery; and Soviet maps (see <a href="#sec2dot2-remotesensing-13-03106" class="html-sec">Section 2.2</a>, <a href="#sec2dot3-remotesensing-13-03106" class="html-sec">Section 2.3</a> and <a href="#sec2dot4-remotesensing-13-03106" class="html-sec">Section 2.4</a>, respectively). Iraq roads and waterways from OpenStreetMap.org are displayed onto ArcGIS basemap image.</p>
Full article ">Figure 2
<p>(<b>a</b>) Spatial extent and coverage of CORONA KH-4B tiles accessed from [<a href="#B61-remotesensing-13-03106" class="html-bibr">61</a>] and Soviet maps [<a href="#B66-remotesensing-13-03106" class="html-bibr">66</a>,<a href="#B67-remotesensing-13-03106" class="html-bibr">67</a>], overlapped onto 10 m resolution Copernicus Sentinel-2 data 2019. The yellow polygon indicates the location of the zoomed views in (<b>b</b>,<b>c</b>). (<b>b</b>) CORONA tile compared with (<b>c</b>) the Soviet map. Archaeological and natural features documented in these two historical cartographic and imagery sources are mapped and jointly displayed. (<b>d</b>) Excerpt of the Soviet topographic map symbols indicating “62. Burial mound (height indicated in meters)”, “456. Triangulation point on burial mound”, “472. Burial mound”, and “473. Tailings pile” from the map legend translated by the US Department of the Army (1958) [<a href="#B68-remotesensing-13-03106" class="html-bibr">68</a>]. The cyan polygons in picture (<b>c</b>) highlight surveyors’ annotation of tells.</p>
Full article ">Figure 3
<p>Spatial distribution of placemarks of Ancient Near East (ANE) sites as mapped by [<a href="#B70-remotesensing-13-03106" class="html-bibr">70</a>]. The density of mapped features decreases going toward the east from the Dejmej reservoir. Iraq administrative boundaries, roads, and waterways from OpenStreetMap.org, onto ArcGIS basemap image.</p>
Full article ">Figure 4
<p>(<b>a</b>) Overview of high-resolution COSMO-SkyMed SM DEM tiles #1-#3 with (<b>b</b>) zoomed view on a sector with a number of tells and archaeological deposits. COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2018. All rights reserved.</p>
Full article ">Figure 5
<p>(<b>a</b>) Overview of high-resolution COSMO-SkyMed SM DEM tiles #4 and #5 with (<b>b</b>) zoomed view on a sector with a number of tells and archaeological deposits. COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2019. All rights reserved.</p>
Full article ">Figure 6
<p>Tell detection through combination of (<b>a</b>) very high resolution (VHR) optical imagery, (<b>b</b>) Sentinel-2 multispectral false-colored composite (R: Band 8—NIR; G: Band 4—red; B: Band 3—green), COSMO-SkyMed (<b>c</b>) SAR radar backscatter image, and (<b>d</b>) DEM with (<b>e</b>) height profile drawn along A-A’ section. Products in (<b>b</b>–<b>d</b>) were all collected in April 2018. Example of clear visibility of a single mound with mud-brick architecture (precise location undisclosed for security concerns). COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2018. All rights reserved. Contains Copernicus Sentinel-2 data, 2018.</p>
Full article ">Figure 7
<p>Tell detection through combination of (<b>a</b>) VHR optical imagery, (<b>b</b>) Sentinel-2 multispectral false-colored composite (R: Band 8—NIR; G: Band 4—red; B: Band 3—green), COSMO-SkyMed (<b>c</b>) SAR radar backscatter image, and (<b>d</b>) DEM with (<b>e</b>) height profile drawn along A-A’ section. Products in (<b>b</b>–<b>d</b>) were all collected in April 2018, as per <a href="#remotesensing-13-03106-f006" class="html-fig">Figure 6</a>. Example of small flat mound (Neo-Babylonian to Islamic periods; precise location undisclosed for security concerns) that is detected thanks to the DEM only. COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2018. All rights reserved. Contains Copernicus Sentinel-2 data, 2018.</p>
Full article ">Figure 8
<p>Spatial distribution of tells (indicated by archaeological “priority” score from 1 “very low” to 5 “very high”) in the north-western region of the study area (Al-Zubaidiya and Al-Shehamiya sub-districts, administrative codes IQG18Q04N01 and IQG18Q04N02, respectively) vs. the location of artificial and natural mounds as recorded in historic maps, and paleo-channels mapped from CORONA and more recent satellite data. Contains Copernicus Sentinel-2 data, 2019.</p>
Full article ">Figure 9
<p>Dense spatial distribution of tells at “medium” to “very high” archaeological priority (i.e., scores from 3 to 5) in the south-eastern sector of the study area, north-east of the Delmej reservoir (Al-Ahrar sub-district, code IQG18Q03N01), onto COSMO-SkyMed<sup>®</sup> DEM Product ©ASI, Italian Space Agency, 2018-2019. All rights reserved.</p>
Full article ">Figure 10
<p>Spatial distribution of mapped tells (<b>a</b>) by priority and their diameter (expressed in meters); and (<b>b</b>) by condition as assessed through visual interpretation of evidence of looting in satellite imagery.</p>
Full article ">Figure 11
<p>Statistics of the mapped tells by: (<b>a</b>) size (i.e., maximum diameter) vs. archaeological priority, (<b>b</b>) size class, (<b>c</b>) condition (i.e., as assessed through visual interpretation of satellite imagery), and (<b>d</b>) looting evidence vs. archaeological priority.</p>
Full article ">Figure 12
<p>Selection of tells (precise location undisclosed for security concerns) with evidence of looting (yellow arrows) in areas where crop- and soil-marks highlight the presence of buried walls (white arrows), as observed from Google Earth imagery collected on (<b>a</b>) 1 January 2005 versus (<b>b</b>) 13 March 2010, respectively, and (<b>c</b>) ESRI basemap imagery and (<b>d</b>) Google Earth imagery, the latter collected on 21 November 2004. Google Earth image © 2021 Maxar Technologies.</p>
Full article ">Figure 13
<p>(<b>a</b>) Location of the area covered by (<b>b</b>) the bespoke time series of COSMO-SkyMed Enhanced Spotlight images collected in the period 27 July 2019–24 March 2020 to monitor the condition of (<b>c</b>) a cluster of high to very high priority tells (precise location undisclosed for security concerns). COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2019–2020. All rights reserved.</p>
Full article ">Figure 14
<p>(<b>a</b>) VHR optical (21 November 2004) and (<b>b</b>) multi-temporal average COSMO-SkyMed Enhanced Spotlight (27 July 2019–24 March 2020) images showing clear evidence of looting at Tulul el-Barakat TB1 site (id.57, known from [<a href="#B70-remotesensing-13-03106" class="html-bibr">70</a>]). Impact and extent of 2019–2020 anthropogenic disturbance are highlighted through (<b>c</b>) COSMO-SkyMed amplitude change detection map. Google Earth image © 2021 Maxar Technologies; COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2019–2020.</p>
Full article ">Figure 15
<p>(<b>a</b>) Multi-temporal average COSMO-SkyMed Enhanced Spotlight image (27 July 2019–24 March 2020) of detected tells in which Google Earth imagery showed evidence of looting. (<b>b</b>) COSMO-SkyMed amplitude change detection map (2019–2020) suggests a differential impact of anthropogenic disturbance between the sites. COSMO-SkyMed<sup>®</sup> Products ©ASI, Italian Space Agency, 2019–2020.</p>
Full article ">
Back to TopTop