[go: up one dir, main page]

 
 
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

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (348)

Search Parameters:
Keywords = ALOS PALSAR

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
22 pages, 12863 KiB  
Article
Remote and Proximal Sensors Data Fusion: Digital Twins in Irrigation Management Zoning
by Hugo Rodrigues, Marcos B. Ceddia, Wagner Tassinari, Gustavo M. Vasques, Ziany N. Brandão, João P. S. Morais, Ronaldo P. Oliveira, Matheus L. Neves and Sílvio R. L. Tavares
Sensors 2024, 24(17), 5742; https://doi.org/10.3390/s24175742 - 4 Sep 2024
Viewed by 276
Abstract
The scientific field of precision agriculture employs increasingly innovative techniques to optimize inputs, maximize profitability, and reduce environmental impact. However, obtaining a high number of soil samples is challenging in order to make precision agriculture viable. There is a trade-off between the amount [...] Read more.
The scientific field of precision agriculture employs increasingly innovative techniques to optimize inputs, maximize profitability, and reduce environmental impact. However, obtaining a high number of soil samples is challenging in order to make precision agriculture viable. There is a trade-off between the amount of data needed and the time and resources spent to obtain these data compared to the accuracy of the maps produced with more or fewer points. In the present study, the research was based on an exhaustive dataset of apparent electrical conductivity (aEC) containing 3906 points distributed along 26 transects with spacing between each of up to 40 m, measured by the proximal soil sensor EM38-MK2, for a grain-producing area of 72 ha in São Paulo, Brazil. A second sparse dataset was simulated, showing only four transects with a 400 m distance and, in the end, only 162 aEC points. The aEC map via ordinary kriging (OK) from the grid with 26 transects was considered the reference, and two other mapping approaches were used to map aEC via sparse grid: kriging with external drift (KED) and geographically weighted regression (GWR). These last two methods allow the increment of auxiliary variables, such as those obtained by remote sensors that present spatial resolution compatible with the pivot scale, such as data from the Landsat-8, Aster, and Sentinel-2 satellites, as well as ten terrain covariates derived from the Alos Palsar digital elevation model. The KED method, when used with the sparse dataset, showed a relatively good fit to the aEC data (R2 = 0.78), with moderate prediction accuracy (MAE = 1.26, RMSE = 1.62) and reasonable predictability (RPD = 1.76), outperforming the GWR method, which had the weakest performance (R2 = 0.57, MAE = 1.78, RMSE = 2.30, RPD = 0.81). The reference aEC map using the exhaustive dataset and OK showed the highest accuracy with an R2 of 0.97, no systematic bias (ME = 0), and excellent precision (RMSE = 0.56, RPD = 5.86). Management zones (MZs) derived from these maps were validated using soil texture data from clay samples measured at 0–10 cm depth in a grid of 72 points. The KED method demonstrated the highest potential for accurately defining MZs for irrigation, producing a map that closely resembled the reference MZ map, thereby providing reliable guidance for irrigation management. Full article
(This article belongs to the Special Issue Sensors and Artificial Intelligence in Smart Agriculture)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the methodology of simulation of the aEC dataset with sparse sampling and the mapping methods followed by the management zones approach.</p>
Full article ">Figure 2
<p>Location of the study area with the height gradient and digital elevation model.</p>
Full article ">Figure 3
<p>(<b>A</b>) EM38-MK2 being calibrated to the specific magnetic scenario of the field; (<b>B</b>) the sensor is paired with the handheld controller to set the timing acquisition.</p>
Full article ">Figure 4
<p>Study area with digital elevation model in the background; Exhaustive Grid and Sparse Grid showing the distance between sampling lines; external validation dataset.</p>
Full article ">Figure 5
<p>Remote sensing covariates used as predictors in the kriging with external drift and geographically weighted regression.</p>
Full article ">Figure 6
<p>(<b>A</b>) Soil sampling of 0–10 cm using soil sampler ring; (<b>B</b>) planting area covered by beans and irrigated by a central pivot on the background.</p>
Full article ">Figure 7
<p>Electrical conductivity data in mS/m (millisiemens per meter); (<b>A</b>) original format; (<b>B</b>) transformed to Neperian logarithm.</p>
Full article ">Figure 8
<p>Predicted versus experimental aEC values of proximal sensor data as a function of remote sensor data using the training dataset. The continuous black lines adjust the intercept and slope for the models, while the dashed lines are intercepted and idealized as 1 and 0, respectively. R<sup>2</sup> adj: R<sup>2</sup> adjusted value; aEC: apparent electrical conductivity in mS/m (millisiemens per meter).</p>
Full article ">Figure 9
<p>Empirical (circles) and adjusted (lines) semivariograms of apparent electrical conductivity (aEC) in mS/m. (<b>A</b>) Using ordinary kriging with 26 lines (reference); (<b>B</b>) using ordinary kriging with four rows (sparse); (<b>C</b>) using kriging with external drift of aEC data with sparse data as a function of remote sensor data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>D</b>) using geographically weighted regression with sparse data as a function of remote sensor data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>E</b>) semivariogram of the R<sup>2</sup> indices obtained by calculating the GWR for spatialization.</p>
Full article ">Figure 10
<p>Maps of apparent electrical conductivity (aEC) in mS/m. (<b>A</b>) Using ordinary kriging with 26 lines (reference); (<b>B</b>) using ordinary kriging with four rows (sparse); (<b>C</b>) using kriging with external drift of the sparse data as a function of the remote sensing data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>D</b>) using geographically weighted regression with sparse data as a function of remote sensor data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>E</b>) map of the adjusted R<sup>2</sup> obtained by calculating the GWR for the aEC sparse dataset.</p>
Full article ">Figure 10 Cont.
<p>Maps of apparent electrical conductivity (aEC) in mS/m. (<b>A</b>) Using ordinary kriging with 26 lines (reference); (<b>B</b>) using ordinary kriging with four rows (sparse); (<b>C</b>) using kriging with external drift of the sparse data as a function of the remote sensing data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>D</b>) using geographically weighted regression with sparse data as a function of remote sensor data defined in <a href="#sec3dot5-sensors-24-05742" class="html-sec">Section 3.5</a>; (<b>E</b>) map of the adjusted R<sup>2</sup> obtained by calculating the GWR for the aEC sparse dataset.</p>
Full article ">Figure 11
<p>Maps of management zones for soil types. (<b>A</b>) Using ordinary kriging with 26 lines (reference); (<b>B</b>) using ordinary kriging with four rows (sparse); (<b>C</b>) using kriging with external drift of the sparse aEC data as a function of the remote sensing data defined in <a href="#sec2dot6dot2-sensors-24-05742" class="html-sec">Section 2.6.2</a>; (<b>D</b>) using geographically weighted regression with sparse data as a function of remote sensor data defined in <a href="#sec2dot6dot3-sensors-24-05742" class="html-sec">Section 2.6.3</a>.</p>
Full article ">
29 pages, 38452 KiB  
Article
Integration of Multi-Source Datasets for Assessing Ground Swelling/Shrinking Risk in Cyprus: The Case Studies of Pyrgos–Parekklisia and Moni
by Athanasios V. Argyriou, Maria Prodromou, Christos Theocharidis, Kyriaki Fotiou, Stavroula Alatza, Constantinos Loupasakis, Zampela Pittaki-Chrysodonta, Charalampos Kontoes, Diofantos G. Hadjimitsis and Marios Tzouvaras
Remote Sens. 2024, 16(17), 3185; https://doi.org/10.3390/rs16173185 - 28 Aug 2024
Viewed by 629
Abstract
The determination of swelling/shrinking phenomena, from natural and anthropogenic activity, is examined in this study through the synergy of various remote sensing methodologies. For the period of 2016–2022, a time-series InSAR analysis of Sentinel-1 satellite images, with a Coherent Change Detection procedure, was [...] Read more.
The determination of swelling/shrinking phenomena, from natural and anthropogenic activity, is examined in this study through the synergy of various remote sensing methodologies. For the period of 2016–2022, a time-series InSAR analysis of Sentinel-1 satellite images, with a Coherent Change Detection procedure, was conducted to calculate the Normalized Coherence Difference. These were combined with Sentinel-2 multispectral data by exploiting the Normalized Difference Vegetation Index to create multi-temporal image composites. In addition, ALOS-Palsar DEM derivatives highlighted the geomorphological characteristics, which, in conjunction with the satellite imagery outcomes and other auxiliary spatial datasets, were embedded within a Multi-Criteria Decision Analysis (MCDA) model. The synergy of the remote sensing and GIS techniques’ applicability within the MCDA model highlighted the zones undergoing seasonal swelling/shrinking processes in Pyrgos–Parekklisia and Moni regions in Cyprus. The accuracy assessment of the produced final MCDA outcome provided an overall accuracy of 72.4%, with the Kappa statistic being 0.66, indicating substantial agreement of the MCDA outcome with the results from a Persistent Scatterer Interferometry analysis and ground-truth observations. Thus, this study offers decision-makers a powerful procedure to monitor longer- and shorter-term swelling/shrinking phenomena. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Pyrgos–Parekklisia, Moni, and Monagroulli deforming sites in Limassol, Cyprus.</p>
Full article ">Figure 2
<p>The Pyrgos Lemesou–Parekklisia and Moni–Monagroulli geology.</p>
Full article ">Figure 3
<p>Sentinel-1 satellite passes in ascending and descending tracks and satellite image details.</p>
Full article ">Figure 4
<p>The Coherent Change Detection workflow methodology. The step that provides the coherence values is marked in red.</p>
Full article ">Figure 5
<p>Pyrgos–Parekklisia area. (<b>a</b>) Coherence difference and (<b>b</b>) Normalized Coherence difference from descending Sentinel-1 satellite images during 12 February 2021–8 March 2021. (<b>c</b>) Coherence difference and (<b>d</b>) Normalized Coherence difference from ascending Sentinel-1 satellite images during 23 February 2021–7 March 2021.</p>
Full article ">Figure 6
<p>Moni–Monagroulli area. (<b>a</b>) Coherence difference and (<b>b</b>) Normalized Coherence difference from descending Sentinel-1 satellite images during 12 February 2021–8 March 2021. (<b>c</b>) Coherence difference and (<b>d</b>) Normalized Coherence difference from ascending Sentinel-1 satellite images during 23 February 2021–7 March 2021.</p>
Full article ">Figure 7
<p>Annual NDVI variations and corresponding masked areas excluded from further analysis.</p>
Full article ">Figure 8
<p>CCD of the Area of Interest showing the changes that occurred between 2016 and 2022.</p>
Full article ">Figure 9
<p>The TWI with dark bluish hues highlighting the high moisture accumulation.</p>
Full article ">Figure 10
<p>Landform type classification, showing valleys, semi-mountainous, and mountainous zones.</p>
Full article ">Figure 11
<p>The determined precipitation derived from the weather stations.</p>
Full article ">Figure 12
<p>The soil texture map of the AoI.</p>
Full article ">Figure 13
<p>The reclassified soil texture map, highlighting the degree of the swelling/shrinking effect.</p>
Full article ">Figure 14
<p>The reclassified hydrogeological map highlights the swelling degree.</p>
Full article ">Figure 15
<p>The GIS-based MCDA swelling and shrinking effect outcome based on the acknowledged variables of CCD, soil texture, hydrogeology, TWI, landforms, and rainfall. High-risk zones are presented in orange and very high-risk zones in red.</p>
Full article ">Figure 16
<p>(<b>a</b>) Sentinel-1 LoS displacements in Pyrgos–Parekklisia for descending satellite pass and (<b>b</b>) interpolated Sentinel-1 LOS displacements in Pyrgos–Parekklisia for descending satellite pass.</p>
Full article ">Figure 17
<p>The MCDA swelling and shrinking effect outcome with the overlaid ground-truth locations with verified deformed structures, indicated with red arrows, from ground-truth surveys.</p>
Full article ">Figure A1
<p>The distribution of the accuracy assessment points across the final MCDA swelling/shrinking effect outcome.</p>
Full article ">
20 pages, 9326 KiB  
Article
Retrospect on the Ground Deformation Process and Potential Triggering Mechanism of the Traditional Steel Production Base in Laiwu with ALOS PALSAR and Sentinel-1 SAR Sensors
by Chao Ding, Guangcai Feng, Lu Zhang and Wenxin Wang
Sensors 2024, 24(15), 4872; https://doi.org/10.3390/s24154872 - 26 Jul 2024
Viewed by 510
Abstract
The realization of a harmonious relationship between the natural environment and economic development has always been the unremitting pursuit of traditional mineral resource-based cities. With rich reserves of iron and coal ore resources, Laiwu has become an important steel production base in Shandong [...] Read more.
The realization of a harmonious relationship between the natural environment and economic development has always been the unremitting pursuit of traditional mineral resource-based cities. With rich reserves of iron and coal ore resources, Laiwu has become an important steel production base in Shandong Province in China, after several decades of industrial development. However, some serious environmental problems have occurred with the quick development of local steel industries, with ground subsidence and consequent secondary disasters as the most representative ones. To better evaluate possible ground collapse risk, comprehensive approaches incorporating the common deformation monitoring with small-baseline subset (SBAS)-synthetic aperture radar interferometry (InSAR) technique, environmental factors analysis, and risk evaluation are designed here with ALOS PALSAR and Sentinel-1 SAR observations. A retrospect on the ground deformation process indicates that ground deformation has largely decreased by around 51.57% in area but increased on average by around −5.4 mm/year in magnitude over the observation period of Sentinel-1 (30 July 2015 to 22 August 2022), compared to that of ALOS PALSAR (17 January 2007 to 28 October 2010). To better reveal the potential triggering mechanism, environmental factors are also utilized and conjointly analyzed with the ground deformation time series. These analysis results indicate that the ground deformation signals are highly correlated with human industrial activities, such underground mining, and the operation of manual infrastructures (landfill, tailing pond, and so on). In addition, the evaluation demonstrates that the area with potential collapse risk (levels of medium, high, and extremely high) occupies around 8.19 km2, approximately 0.86% of the whole study region. This study sheds a bright light on the safety guarantee for the industrial operation and the ecologically friendly urban development of traditional steel production industrial cities in China. Full article
Show Figures

Figure 1

Figure 1
<p>The geographical background of the study area in Laiwu, in which the image coverages for ALOS PALSAR and Sentinel-1 are contoured with blue and cyan polylines, respectively. The background image of this study region is the topographic map. Notably, indicated by the magenta dots, 15 test points coded from A to O are located at Yujiaquan tailing pond (A, B, C, D), the iron ore mining region (E, F, G, H, I), the banksides of the Dawen River (J, K, L), and the coal mining region (M, N, O).</p>
Full article ">Figure 2
<p>Diagram of the methodology utilized for retrieving the ground deformation process, analyzing the potential triggering mechanisms, and evaluating the possible subsidence risks.</p>
Full article ">Figure 3
<p>(<b>a</b>) ALOS PALSAR (17 January 2007~28 October 2010) and (<b>b</b>) Sentinel-1 (30 July 2015~22 August 2022) derived light-of-sight (LOS) deformation velocities (mm/year).</p>
Full article ">Figure 4
<p>The identified LOS deformation region covered by green points for (<b>a</b>) ALOS PALSAR observations from 17 January 2007 to 28 October 2010, and (<b>b</b>) Sentinel-1 observations from 30 July 2015 to 22 August 2022.</p>
Full article ">Figure 5
<p>The LOS deformation velocities of the region near the banksides of the Dawen River derived from (<b>a</b>) ALOS PALSAR observations (17 January 2007~28 October 2010) and (<b>b</b>) Sentinel-1 observations (30 July 2015~22 August 2022). Notably, indicated by the magenta dots, 4 test points coded from J, K, L and M are located at the banksides of the Dawen River (J, K, L) and the coal mining region (M).</p>
Full article ">Figure 6
<p>The LOS deformation velocities of the Yujiaquan tailing pond derived from (<b>a</b>) ALOS PALSAR observations (17 January 2007~28 October 2010) and (<b>b</b>) Sentinel-1 observations (30 July 2015~22 August 2022). Notably, indicated by the magenta dots, 4 test points coded from A, B, C and D are located at Yujiaquan tailing pond.</p>
Full article ">Figure 7
<p>The LOS deformation velocities of the Yaojialing iron ore-mining region derived from (<b>a</b>) ALOS PALSAR observations (17 January 2007~28 October 2010) and (<b>b</b>) Sentinel-1 observations (30 July 2015~22 August 2022). Notably, the magenta point of I indicates the location of Yaojialing iron ore-mining region.</p>
Full article ">Figure 8
<p>The LOS deformation velocities of the newly constructed landfill for industrial wastes derived from (<b>a</b>) ALOS PALSAR observations (17 January 2007~28 October 2010) and (<b>b</b>) Sentinel-1 observations (30 July 2015~22 August 2022). Notably, with the location indicated by the magenta point of H, this landfill was constructed from July 2019 to December 2019.</p>
Full article ">Figure 9
<p>The time-series LOS deformation for test points of (<b>a</b>–<b>i</b>), in which A, B, C, and D are located on the banksides of the Yujiaquan tailing pond; E, F, and G are located in the traditional Luzhong iron ore-mining region; H is located in a newly constructed landfill for industrial wastes; I is located in the Yaojialing iron ore-mining region. In addition, the LOS deformation time series of J, K, L, M, N, and O can be found in <a href="#app1-sensors-24-04872" class="html-app">Figures S1–S6</a>. The environmental factors incorporating the daily precipitation, the cumulative precipitation, and the earthquake events, are cross-compared to the deformation time series derived from ALOS PALSAR and Sentinel-1 SAR observations.</p>
Full article ">Figure 9 Cont.
<p>The time-series LOS deformation for test points of (<b>a</b>–<b>i</b>), in which A, B, C, and D are located on the banksides of the Yujiaquan tailing pond; E, F, and G are located in the traditional Luzhong iron ore-mining region; H is located in a newly constructed landfill for industrial wastes; I is located in the Yaojialing iron ore-mining region. In addition, the LOS deformation time series of J, K, L, M, N, and O can be found in <a href="#app1-sensors-24-04872" class="html-app">Figures S1–S6</a>. The environmental factors incorporating the daily precipitation, the cumulative precipitation, and the earthquake events, are cross-compared to the deformation time series derived from ALOS PALSAR and Sentinel-1 SAR observations.</p>
Full article ">Figure 9 Cont.
<p>The time-series LOS deformation for test points of (<b>a</b>–<b>i</b>), in which A, B, C, and D are located on the banksides of the Yujiaquan tailing pond; E, F, and G are located in the traditional Luzhong iron ore-mining region; H is located in a newly constructed landfill for industrial wastes; I is located in the Yaojialing iron ore-mining region. In addition, the LOS deformation time series of J, K, L, M, N, and O can be found in <a href="#app1-sensors-24-04872" class="html-app">Figures S1–S6</a>. The environmental factors incorporating the daily precipitation, the cumulative precipitation, and the earthquake events, are cross-compared to the deformation time series derived from ALOS PALSAR and Sentinel-1 SAR observations.</p>
Full article ">Figure 9 Cont.
<p>The time-series LOS deformation for test points of (<b>a</b>–<b>i</b>), in which A, B, C, and D are located on the banksides of the Yujiaquan tailing pond; E, F, and G are located in the traditional Luzhong iron ore-mining region; H is located in a newly constructed landfill for industrial wastes; I is located in the Yaojialing iron ore-mining region. In addition, the LOS deformation time series of J, K, L, M, N, and O can be found in <a href="#app1-sensors-24-04872" class="html-app">Figures S1–S6</a>. The environmental factors incorporating the daily precipitation, the cumulative precipitation, and the earthquake events, are cross-compared to the deformation time series derived from ALOS PALSAR and Sentinel-1 SAR observations.</p>
Full article ">Figure 10
<p>(<b>a</b>) The risk level map and (<b>b</b>) corresponding statistical pie chart of the traditional steel production base in Laiwu.</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>) The risk level map and (<b>b</b>) corresponding statistical pie chart of the traditional steel production base in Laiwu.</p>
Full article ">
27 pages, 6641 KiB  
Article
Biomass Estimation and Saturation Value Determination Based on Multi-Source Remote Sensing Data
by Rula Sa, Yonghui Nie, Sergey Chumachenko and Wenyi Fan
Remote Sens. 2024, 16(12), 2250; https://doi.org/10.3390/rs16122250 - 20 Jun 2024
Viewed by 718
Abstract
Forest biomass estimation is undoubtedly one of the most pressing research subjects at present. Combining multi-source remote sensing information can give full play to the advantages of different remote sensing technologies, providing more comprehensive and rich information for aboveground biomass (AGB) estimation research. [...] Read more.
Forest biomass estimation is undoubtedly one of the most pressing research subjects at present. Combining multi-source remote sensing information can give full play to the advantages of different remote sensing technologies, providing more comprehensive and rich information for aboveground biomass (AGB) estimation research. Based on Landsat 8, Sentinel-2A, and ALOS2 PALSAR data, this paper takes the artificial coniferous forests in the Saihanba Forest of Hebei Province as the object of study, fully explores and establishes remote sensing factors and information related to forest structure, gives full play to the advantages of spectral signals in detecting the horizontal structure and multi-dimensional synthetic aperture radar (SAR) data in detecting the vertical structure, and combines environmental factors to carry out multivariate synergistic methods of estimating the AGB. This paper uses three variable selection methods (Pearson correlation coefficient, random forest significance, and the least absolute shrinkage and selection operator (LASSO)) to establish the variable sets, combining them with three typical non-parametric models to estimate AGB, namely, random forest (RF), support vector regression (SVR), and artificial neural network (ANN), to analyze the effect of forest structure on biomass estimation, explore the suitable AGB of artificial coniferous forests estimation of machine learning models, and develop the method of quantifying saturation value of the combined variables. The results show that the horizontal structure is more capable of explaining the AGB compared to the vertical structure information, and that combining the multi-structure information can improve the model results and the saturation value to a great extent. In this study, different sets of variables can produce relatively superior results in different models. The variable set selected using LASSO gives the best results in the SVR model, with an R2 values of 0.9998 and 0.8792 for the training and the test set, respectively, and the highest saturation value obtained is 185.73 t/ha, which is beyond the range of the measured data. The problem of saturation in biomass estimation in boreal medium- and high-density forests was overcome to a certain extent, and the AGB of the Saihanba area was better estimated. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area: (<b>a</b>) the location map of the study area; (<b>b</b>) the HV polarization data of the study area; (<b>c</b>) the true color image of Sentinel-2, with the actual sample locations indicated by green dots.</p>
Full article ">Figure 2
<p>Relationships between forest structural parameters at the sample site level: (<b>a</b>) mean DBH vs. S; (<b>b</b>) CC vs. BA; (<b>c</b>) mean DBH vs. mean forest height, where the size and color shade of the dots vary with biomass.</p>
Full article ">Figure 3
<p>Flowchart of the methodology.</p>
Full article ">Figure 4
<p>Determination of the number of model leaves and decision tree.</p>
Full article ">Figure 5
<p>Parameter optimization diagram of three models. From left to right are the results of RF, SVR, and ANN models. From top to bottom are the results obtained for the horizontal structure indices (V1), vertical structure indices (V2), horizontal + vertical structure indices (V3), horizontal + vertical structure indices + topographical variables (V4), Pearson selection variable (V5), RF importance selection of the variables (V6), and the variable chosen by the LASSO (V7) in each model.</p>
Full article ">Figure 6
<p>Summary graphs of the results of the training and test sets of the three models for estimating AGB. The first and second rows of each model are the training set results, and test set results, respectively. The left side is <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>R</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>, and the right side is RMSE.</p>
Full article ">Figure 7
<p>Summary plot of model results. From left to right, the horizontal structure indices (V1), vertical structure indices (V2), horizontal + vertical structure indices (V3), horizontal + vertical structure indices + topographical variables (V4), Pearson selection variable (V5), RF importance selection of the variables (V6), and the variable chosen by the LASSO (V7) were introduced into the three models to estimate the results of AGB. From top to bottom are the results of RF, SVR, and ANN models, and the first and second rows of each model are the training set results and test set results, respectively. The horizontal axis of the image is the measured data, the vertical axis is the predicted results, the blue is the 1:1 straight line, and the green is the fitted line.</p>
Full article ">Figure 8
<p>AGB map of the study area estimated by the LASSO-based SVR model: (<b>a</b>) AGB map of the study area; (<b>b</b>) histogram of AGB distribution.</p>
Full article ">Figure 9
<p>The spherical model curves for each structure index and different variable sets under different ML models. The left side shows the horizontal structure indices figures (RVI, CTI, MTI, PTI) in order, and the individual index figures of CC, S, and BA are shown on the right. Right side: vertical structure indices figures and a summary plot of the spherical model curves for the different sets of variables under the RF, SVR, and ANN models.</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
Viewed by 792
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 ">
19 pages, 6754 KiB  
Article
Influence of Image Compositing and Multisource Data Fusion on Multitemporal Land Cover Mapping of Two Philippine Watersheds
by Nico R. Almarines, Shizuka Hashimoto, Juan M. Pulhin, Cristino L. Tiburan, Angelica T. Magpantay and Osamu Saito
Remote Sens. 2024, 16(12), 2167; https://doi.org/10.3390/rs16122167 - 14 Jun 2024
Cited by 1 | Viewed by 675
Abstract
Cloud-based remote sensing has spurred the use of techniques to improve mapping accuracy where individual images may have lower quality, especially in areas with complex terrain or high cloud cover. This study investigates the influence of image compositing and multisource data fusion on [...] Read more.
Cloud-based remote sensing has spurred the use of techniques to improve mapping accuracy where individual images may have lower quality, especially in areas with complex terrain or high cloud cover. This study investigates the influence of image compositing and multisource data fusion on the multitemporal land cover mapping of the Pagsanjan-Lumban and Baroro Watersheds in the Philippines. Ten random forest models for each study site were used, all using a unique combination of more than 100 different input features. These features fall under three general categories. First, optical features were derived from reflectance bands and ten spectral indices, which were further subdivided into annual percentile and seasonal median composites; second, radar features were derived from ALOS PALSAR by computing textural indices and a simple band ratio; and third, topographic features were computed from the ALOS GDSM. Then, accuracy metrics and McNemar’s test were used to assess and compare the significance of about 90 pairwise model outputs. Data fusion significantly improved the accuracy of multitemporal land cover mapping in most cases. However, image composition had varied impacts for both sites. This could imply local characteristics and feature inputs as potential determinants of the ideal composite method. Hence, the iterative screening or optimization of both input features and composites is recommended to improve multitemporal mapping accuracy. Full article
Show Figures

Figure 1

Figure 1
<p>Methodological flowchart of this study.</p>
Full article ">Figure 2
<p>A location map of the study sites.</p>
Full article ">Figure 3
<p>Images of the land cover present in the study sites. (<b>a</b>) Inland water (foreground) and residential build-up (background) in the PLW, (<b>b</b>) lowland annual crops in the PLW, (<b>c</b>) open forest in the northeast of the PLW, (<b>d</b>) lowland annual crops in the BW, (<b>e</b>) grassland in the rolling hills of the PLW, and (<b>f</b>) grassland (foreground) and a mosaic of cropland, brushland, and open forest in the uplands of the BW (background).</p>
Full article ">Figure 4
<p>The model performance of various feature sets in (<b>a</b>) the PLW and (<b>b</b>) the BW order based on accuracy.</p>
Full article ">Figure 5
<p>Yate’s <span class="html-italic">p</span>-values of McNemar’s test of pairwise RF model comparisons for the (<b>a</b>) PLW and (<b>b</b>) BW.</p>
Full article ">Figure 6
<p>Generated land cover maps of (<b>a</b>) the PLW and (<b>b</b>) the BW from 2000 to 2020.</p>
Full article ">Figure 7
<p>Net land cover change from the 2000 baseline in (<b>a</b>) the PLW and (<b>b</b>) the BW.</p>
Full article ">
29 pages, 18921 KiB  
Article
RadWet-L: A Novel Approach for Mapping of Inundation Dynamics of Forested Wetlands Using ALOS-2 PALSAR-2 L-Band Radar Imagery
by Gregory Oakes, Andy Hardy, Pete Bunting and Ake Rosenqvist
Remote Sens. 2024, 16(12), 2078; https://doi.org/10.3390/rs16122078 - 8 Jun 2024
Viewed by 747
Abstract
The ability to accurately map tropical wetland dynamics can significantly contribute to a number of areas, including food and water security, protection and enhancement of ecosystems, flood hazard management, and our understanding of natural greenhouse gas emissions. Yet currently, there is not a [...] Read more.
The ability to accurately map tropical wetland dynamics can significantly contribute to a number of areas, including food and water security, protection and enhancement of ecosystems, flood hazard management, and our understanding of natural greenhouse gas emissions. Yet currently, there is not a tractable solution for mapping tropical forested wetlands at high spatial and temporal resolutions at a regional scale. This means that we lack accurate and up-to-date information about some of the world’s most significant wetlands, including the Amazon Basin. RadWet-L is an automated machine-learning classification technique for the mapping of both inundated forests and open water using ALOS ScanSAR data. We applied and validated RadWet-L for the Amazon Basin. The proposed method is computationally light and transferable across the range of landscape types in the Amazon Basin allowing, for the first time, regional inundation maps to be produced every 42 days at 50 m resolution over the period 2019–2023. Time series estimates of inundation extent from RadWet-L were significantly correlated with NASA-GFZ GRACE-FO water thickness (Pearson’s r = 0.96, p < 0.01), USDA G-REALM lake hight (Pearson’s r between 0.63 and 0.91, p < 0.01), and in situ river stage measurements (Pearson’s r between 0.78 and 0.94, p < 0.01). Additionally, we conducted an evaluation of 11,162 points against the input ScanSAR data revealing spatial and temporal consistency in the approach (F1 score = 0.97). Serial classifications of ALOS-2 PALSAR-2 ScanSAR data by RadWet-L can provide unique insights into the spatio-temporal inundation dynamics within the Amazon Basin. Understanding these dynamics can inform policy in the sustainable use of these wetlands, as well as the impacts of inundation dynamics on biodiversity and greenhouse gas budgets. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Map of the Amazon River Basin to which the RadWet-L image classification was applied and evaluated. (<b>b</b>) Area within the Central Amazon Basin (indicated by the red box) where training data was generated, and algorithm testing and development was carried out.</p>
Full article ">Figure 2
<p>(<b>a</b>) ALOS-2 PALSAR-2 ScanSAR false colour composite (HH, HV, NPDI: Normalised Difference Polarisation Index) acquired between 12 and 25 August 2019; (<b>b</b>) reference optical Landsat 8 and Sentinel-2 false colour composite (NIR, Red, Green). Example area defined as (<b>c</b>) inundated forest, and (<b>d</b>) open water through visual interpretation of the ALOS-2 PALSAR-2 ScanSAR data. Images centre on the Coari region of the Amazon Basin.</p>
Full article ">Figure 3
<p>Flow diagram summarising the RadWet-L automatic training data generation process and classifier training.</p>
Full article ">Figure 4
<p>Flow diagram outlining the data preparation and classifier application for each ALOS-2 PALSAR-2 in the Amazon River Basin.</p>
Full article ">Figure 5
<p>Location of G-REALM Satellite altimetry measurements on 6 lakes around the Central Amazon and associated drainage basins defined from HydroSheds Level 6 and 7 basin products [<a href="#B45-remotesensing-16-02078" class="html-bibr">45</a>]. (a) Lake Badajos, (b) Lake Mamia, (c) Lake Guaribas, (d) Lake Santo, (e) Lake Ariau, and (f) Lake Erepecu. Basemap ESRI Physical.</p>
Full article ">Figure 6
<p>Locations of river stage gauge stations and associated upslope contributing areas. (a) Estirão Do Repouso, (b) Seringal Moreira, (c) Estirão Da Santa Cruz, (d) Acanaui, and (e) Serrinha. Data provided from [<a href="#B59-remotesensing-16-02078" class="html-bibr">59</a>,<a href="#B60-remotesensing-16-02078" class="html-bibr">60</a>].</p>
Full article ">Figure 7
<p>Classified output over the Amazon River Basin from ALOS-2 PALSAR-2 Orbit Cycle 165, acquired between 01/11/20 and 15/11/20 in low water conditions.</p>
Full article ">Figure 8
<p>Classified output over the Amazon River Basin from ALOS-2 PALSAR-2 orbit cycle 180, acquired between 31/05/21 and 13/06/21 in high water conditions.</p>
Full article ">Figure 9
<p>Examples of the RadWet-L classified outputs from ALOS-2 PALSAR-2 orbit cycle 180, acquired between 31/05/21 and 13/06/21, demonstrating the ability of RadWet-L to detect fine-scale inundation features.</p>
Full article ">Figure 10
<p>Pearson R correlation between RadWet-L total wetted area classified extent and GRACE-FO Gravity Anomaly [<a href="#B57-remotesensing-16-02078" class="html-bibr">57</a>] where measurement and observation dates are within 10 days.</p>
Full article ">Figure 11
<p>Comparison between NASA-GFZ GRACE-FO Gravity Anomaly (Water Equivalent Thickness) [<a href="#B57-remotesensing-16-02078" class="html-bibr">57</a>] from January 2019 to June 2023 (<b>top</b>) and the classified total wetted extent (Inundated Forest and Open Water) between January 2019 and 1st October 2023 as measured by RadWet-L (<b>bottom</b>).</p>
Full article ">Figure 12
<p>Pearson-R correlation coefficient calculated between GRACE-FO and RadWet-L mapped total wetted area at 3° × 3° native GRACE-FO resolution across the Amazon Basin [<a href="#B57-remotesensing-16-02078" class="html-bibr">57</a>]. RadWet-L 2019–2023 maximum total inundation extent is overlaid to show main areas of inundation within the basin.</p>
Full article ">Figure 13
<p>Comparison between RadWet-L total wetted area classified outputs with G-REALM satellite altimetry measurements [<a href="#B58-remotesensing-16-02078" class="html-bibr">58</a>] over 6 candidate sub-basins within the larger Amazon Basin.</p>
Full article ">Figure 14
<p>Comparisons between RadWet-L total wetted area, and in situ river stage height at [<a href="#B59-remotesensing-16-02078" class="html-bibr">59</a>] 5 locations within the Amazon River Basin.</p>
Full article ">Figure 15
<p>In situ measurements of river stage height at 5 gauging stations within the Amazon River Basin between 2019 and 2023, with RadWet—L total wetted [<a href="#B59-remotesensing-16-02078" class="html-bibr">59</a>] area.</p>
Full article ">Figure 16
<p>Comparison between ESA GlobCover [<a href="#B63-remotesensing-16-02078" class="html-bibr">63</a>] (<b>top</b>), the Global Lakes and Wetlands Database [<a href="#B61-remotesensing-16-02078" class="html-bibr">61</a>,<a href="#B62-remotesensing-16-02078" class="html-bibr">62</a>] (<b>centre</b>), and RadWet-L Inundation Frequency (2019–2023) (<b>bottom</b>).</p>
Full article ">
18 pages, 12154 KiB  
Article
Blind Edge-Retention Indicator for Assessing the Quality of Filtered (Pol)SAR Images Based on a Ratio Gradient Operator and Confidence Interval Estimation
by Xiaoshuang Ma, Le Li and Gang Wang
Remote Sens. 2024, 16(11), 1992; https://doi.org/10.3390/rs16111992 - 31 May 2024
Viewed by 415
Abstract
Speckle reduction is a key preprocessing approach for the applications of Synthetic Aperture Radar (SAR) data. For many interpretation tasks, high-quality SAR images with a rich texture and structure information are useful. Therefore, a satisfactory SAR image filter should retain this information well [...] Read more.
Speckle reduction is a key preprocessing approach for the applications of Synthetic Aperture Radar (SAR) data. For many interpretation tasks, high-quality SAR images with a rich texture and structure information are useful. Therefore, a satisfactory SAR image filter should retain this information well after processing. Some quantitative assessment indicators have been presented to evaluate the edge-preservation capability of single-polarization SAR filters, among which the non-clean-reference-based (i.e., blind) ones are attractive. However, most of these indicators are derived based only on the basic fact that the speckle is a kind of multiplicative noise, and they do not take into account the detailed statistical distribution traits of SAR data, making the assessment not robust enough. Moreover, to our knowledge, there are no specific blind assessment indicators for fully Polarimetric SAR (PolSAR) filters up to now. In this paper, a blind assessment indicator based on an SAR Ratio Gradient Operator (RGO) and Confidence Interval Estimation (CIE) is proposed. The RGO is employed to quantify the edge gradient between two neighboring image patches in both the speckled and filtered data. A decision is then made as to whether the ratio gradient value in the filtered image is close to that in the unobserved clean image by considering the statistical traits of speckle and a CIE method. The proposed indicator is also extended to assess the PolSAR filters by transforming the polarimetric scattering matrix into a scalar which follows a Gamma distribution. Experiments on the simulated SAR dataset and three real-world SAR images acquired by ALOS-PALSAR, AirSAR, and TerraSAR-X validate the robustness and reliability of the proposed indicator. Full article
(This article belongs to the Special Issue Remote Sensing: 15th Anniversary)
Show Figures

Figure 1

Figure 1
<p>The unbounded PDF curves of the ratio gradient in the cases of <span class="html-italic">r</span><sub>0</sub> = 0.5 and <span class="html-italic">r</span><sub>0</sub> = 2.</p>
Full article ">Figure 2
<p>Basic idea of the proposed indicator. Red squares denote image patches and each arrow with a certain color denotes the gradient between two neighboring patches along a certain direction (for simplicity, only the gradients along two directions are shown). The gradient values in the filtered images should be close to that in the clean image if the filter retains edges well.</p>
Full article ">Figure 3
<p>Diagram of the confidence interval estimation approach.</p>
Full article ">Figure 4
<p>Single-look simulated SAR images. (<b>a</b>) The building image. (<b>b</b>) The homogeneous image.</p>
Full article ">Figure 5
<p>Real SAR images. (<b>a</b>) The ALOS-PALSAR image. (<b>b</b>) The TerraSAR-X image. (<b>c</b>) The AirSAR image.</p>
Full article ">Figure 6
<p>Filtering experiment on the single-look simulated SAR image. (<b>a</b>) The speckled building image. (<b>b</b>) The refined Lee filtered image. (<b>c</b>) The PPB filtered image. (<b>d</b>) The SAR-BM3D filtered image.</p>
Full article ">Figure 7
<p>Signal intensity of the clean reference (black line) and the filtered image (red line) along a line. (<b>a</b>) The refined Lee filter. (<b>b</b>) The PPB filter. (<b>c</b>) The SAR-BM3D filter.</p>
Full article ">Figure 8
<p>Filtering experiment on the ALOS-PALSAR image. (<b>a</b>) The original image. (<b>b</b>) The refined Lee filtered image. (<b>c</b>) The SAR-BM3D filtered image. (<b>d</b>–<b>f</b>) The images filtered by the PPB filter with IT varied from 2 to 4. (<b>g</b>–<b>k</b>) The ratio images of (<b>b</b>–<b>f</b>), respectively.</p>
Full article ">Figure 9
<p>Filtering experiment on the TerraSAR-X image. (<b>a</b>) The original image. (<b>b</b>,<b>c</b>) The filtered image and ratio image of the refined Lee filter, respectively. (<b>d</b>,<b>e</b>) The PPB filtered image with BS = 5 × 5 and 7 × 7, respectively. (<b>f</b>,<b>g</b>) The SAR-BM3D filtered image with BS = 5 × 5 and 7 × 7, respectively. (<b>h</b>–<b>k</b>) The ratio images of (<b>d</b>–<b>g</b>), respectively.</p>
Full article ">Figure 10
<p>Filtering experiment on the AirSAR image. (<b>a</b>) The original image. (<b>b</b>–<b>d</b>) The filtered image by nonlocal means, PNGF and NLTV, respectively.</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 1 | Viewed by 901
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 ">
25 pages, 5950 KiB  
Article
Forest Structure Mapping of Boreal Coniferous Forests Using Multi-Source Remote Sensing Data
by Rula Sa and Wenyi Fan
Remote Sens. 2024, 16(11), 1844; https://doi.org/10.3390/rs16111844 - 22 May 2024
Cited by 1 | Viewed by 704
Abstract
Modeling forest structure using multi-source satellite data is beneficial to understanding the relationship between vertical and horizontal structure and image features to provide more comprehensive and abundant information for the study of forest structural complexity. This study investigates and models forest structure as [...] Read more.
Modeling forest structure using multi-source satellite data is beneficial to understanding the relationship between vertical and horizontal structure and image features to provide more comprehensive and abundant information for the study of forest structural complexity. This study investigates and models forest structure as a multivariate structure based on sample data and active-passive remote sensing data (Landsat8, Sentinel-2A, and ALOS-2 PALSAR) from the Saihanba Forest in Hebei Province, Northern China, to measure forest structural complexity, relying on a relationship-driven model between field and satellite data. In this study, we considered the effects of the role of satellite variables in different vertical structure types and horizontal structure ranges, used two methods to stepwise select significant variables (stepwise forward selection and Pearson correlation coefficient), and employed a multivariate modeling technique (redundancy analysis) to derive a forest composite structure index (FSI), combining both horizontal and vertical structure attributes. The results show that optical texture can better represent forest structure characteristics, polarization interferometric radar information can represent the vertical structure information of forests, and combining the two can represent 77% of the variance of multiple forest structural attributes. The new FSI can explain 93% of the relationship between stand structure and satellite variables, and the linear fit R2 to the measured data reaches 0.91, which largely shows the situation of the measured data. The generated forest structure map more accurately reflects the complexity of the forest structure in the Saihanba Forest, achieving a supplementary explanation of the measured data. Full article
(This article belongs to the Special Issue Vegetation Structure Monitoring with Multi-Source Remote Sensing Data)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location map of the study area, with the actual sample locations indicated by green dots. The detail on the upper right is the spatial location of ALOS-2 PALSAR data, the image on the left is the HV polarization data of the 11 July 2020 view image of the study area, and the right image is the true color image of Sentinel-2A of the 21 June 2020.</p>
Full article ">Figure 2
<p>Relationships between forest structural parameters across sample plots: (<b>a</b>) DBH vs. S; (<b>b</b>) CC vs. BA; (<b>c</b>) DBH vs. H; (<b>d</b>) H vs. CT.</p>
Full article ">Figure 3
<p>Flow chart of the method used in this study.</p>
Full article ">Figure 4
<p>Map of vertical structure types. The length of the thick line in the figure represents the relative number of trees of this size in the sample plot. Small trees (DBH <math display="inline"><semantics> <mrow> <mo>&lt;</mo> </mrow> </semantics></math> 10, H <math display="inline"><semantics> <mrow> <mo>&lt;</mo> </mrow> </semantics></math> 15) are most abundant in the SC1 structure, medium size trees (10 <math display="inline"><semantics> <mrow> <mo>≤</mo> </mrow> </semantics></math> DBH <math display="inline"><semantics> <mrow> <mo>≤</mo> </mrow> </semantics></math> 25, 15 <math display="inline"><semantics> <mrow> <mo>≤</mo> </mrow> </semantics></math> H <math display="inline"><semantics> <mrow> <mo>≤</mo> </mrow> </semantics></math> 20) are most abundant in the SC2 structure, and large trees (DBH <math display="inline"><semantics> <mrow> <mo>&gt;</mo> </mrow> </semantics></math> 25, H <math display="inline"><semantics> <mrow> <mo>&gt;</mo> </mrow> </semantics></math> 20) are most abundant in the SC3 structure.</p>
Full article ">Figure 5
<p>(<b>a</b>) Gravel plot, where eigenvalue is shown in blue and explained fitted variation is shown in red line. (<b>b</b>) 2D ordination plot for redundancy analysis of relationships between stand structural characteristics and remote sensing variables. RDA biplots with lines for explanatory (red) and response (blue) variables.</p>
Full article ">Figure 6
<p>Gravel diagram, where eigenvalue is shown in blue and cumulative is shown in red line.</p>
Full article ">Figure 7
<p>Scatterplot of linear fit of FSI to measured data. The x-axis is the result obtained by PCA for the sample data, and the y-axis is the FSI created from the satellite data.</p>
Full article ">Figure 8
<p>Forest structure image of the Saihanba research area.</p>
Full article ">Figure 9
<p>Using stepwise forward selection of variables on different types and ranges, variables with more than 1% contribution were selected to enter the model for RDA. From left to right, the different SC1, SC2, SC3, and sparse, medium, and dense structure types, and from top to bottom are SC, S, and CC. RDA biplots with lines for explanatory (red) and response (blue) variables.</p>
Full article ">Figure 10
<p>Selection of variables using the Pearson correlation coefficient on different structure types and ranges, and the selection of variables that were significantly correlated at the 0.05 level to be introduced into the model for RDA. From left to right, the different SC1, SC2, and SC3 and sparse, medium, and dense structure types, and from top to bottom are SC, S, and CC. RDA biplots with lines for explanatory (red) and response (blue) variables.</p>
Full article ">Figure 11
<p>The proportion of the degree of contribution of significant variables in the model. The two figures show the proportion of the degree of contribution to the RDA model of the significant variables obtained by stepwise forward selection and Pearson correlation coefficient selection of variables on the nine structures and ranges. The S, M, and D in parentheses represent sparse, medium, and dense. The selected variables were plotted according to eight parts: band, BP, HSI, BC, PDP, PDV, SI, and H-VSI.</p>
Full article ">Figure 12
<p>(<b>a</b>,<b>b</b>) show the results of RDA for optical and SAR variables, respectively. From top to bottom, they are RDA biplots and plots of the contribution of each type of variable (the left y-axis is the amount explained (%), the right y-axis is the contribution (%), and the x-axis is each type of variable). RDA biplots with lines for explanatory (red) and response (blue) variables.</p>
Full article ">Figure 13
<p>The results of RDA performed by F1, F2, and F3 are shown from left to right. RDA biplots with lines for explanatory (red) and response (blue) variables.</p>
Full article ">Figure 14
<p>From left to right are the significant variables obtained by performing RDA for F1, F2, and F3. Where PTI_CC is V1, PTI_S is V2, PTI_BA is V3, 0808-0919H is V4, and 0905-0919H is V5. The x-axis is the significant variable, the left y-axis is the amount of explanation of the variable, and the right y-axis is the contribution of the variable to the model.</p>
Full article ">
21 pages, 23185 KiB  
Article
InSAR-DEM Block Adjustment Model for Upcoming BIOMASS Mission: Considering Atmospheric Effects
by Kefu Wu, Haiqiang Fu, Jianjun Zhu, Huacan Hu, Yi Li, Zhiwei Liu, Afang Wan and Feng Wang
Remote Sens. 2024, 16(10), 1764; https://doi.org/10.3390/rs16101764 - 16 May 2024
Viewed by 713
Abstract
The unique P-band synthetic aperture radar (SAR) instrument, BIOMASS, is scheduled for launch in 2024. This satellite will enhance the estimation of subcanopy topography, owing to its strong penetration and fully polarimetric observation capability. In order to conduct global-scale mapping of the subcanopy [...] Read more.
The unique P-band synthetic aperture radar (SAR) instrument, BIOMASS, is scheduled for launch in 2024. This satellite will enhance the estimation of subcanopy topography, owing to its strong penetration and fully polarimetric observation capability. In order to conduct global-scale mapping of the subcanopy topography, it is crucial to calibrate systematic errors of different strips through interferometric SAR (InSAR) DEM (digital elevation model) block adjustment. Furthermore, the BIOMASS mission will operate in repeat-pass interferometric mode, facing the atmospheric delay errors introduced by changes in atmospheric conditions. However, the existing block adjustment methods aim to calibrate systematic errors in bistatic mode, which can avoid possible errors from atmospheric effects through interferometry. Therefore, there is still a lack of systematic error calibration methods under the interference of atmospheric effects. To address this issue, we propose a block adjustment model considering atmospheric effects. Our model begins by employing the sub-aperture decomposition technique to form forward-looking and backward-looking interferograms, then multi-resolution weighted correlation analysis based on sub-aperture interferograms (SA-MRWCA) is utilized to detect atmospheric delay errors. Subsequently, the block adjustment model considering atmospheric effects can be established based on the SA-MRWCA. Finally, we use robust Helmert variance component estimation (RHVCE) to build the posterior stochastic model to improve parameter estimation accuracy. Due to the lack of spaceborne P-band data, this paper utilized L-band Advanced Land Observing Satellite (ALOS)-1 PALSAR data, which is also long-wavelength, to emulate systematic error calibration of the BIOMASS mission. We chose climatically diverse inland regions of Asia and the coastal regions of South America to assess the model’s effectiveness. The results show that the proposed block adjustment model considering atmospheric effects improved accuracy by 72.2% in the inland test site, with root mean square error (RMSE) decreasing from 10.85 m to 3.02 m. Moreover, the accuracy in the coastal test site improved by 80.2%, with RMSE decreasing from 16.19 m to 3.22 m. Full article
(This article belongs to the Special Issue Remote Sensing for Geology and Mapping)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flow chart of InSAR-DEM block adjustment considering atmospheric effects.</p>
Full article ">Figure 2
<p>The location and the ALOS-1 footprint of the inland test site. The numbers are the index of the SAR data.</p>
Full article ">Figure 3
<p>The location and the ALOS-1 footprint of the coastal test site. The numbers are the index of the SAR data.</p>
Full article ">Figure 4
<p>(<b>a</b>) The control point database of the inland test site; (<b>b</b>) the distribution of GCPs and TPs in the inland test site.</p>
Full article ">Figure 5
<p>Comparison of different methods for removing orbit error phase in inland test site: (<b>a</b>–<b>d</b>) scene-by-scene polynomial fitting; (<b>e</b>–<b>h</b>) block adjustment estimates. The boxes show the difference.</p>
Full article ">Figure 6
<p>The atmospheric correction results of the inland test site: (<b>a</b>–<b>d</b>) ATP estimated by the SA-MRWCA; (<b>e</b>–<b>h</b>) residual phase after atmospheric correction.</p>
Full article ">Figure 7
<p>DEM calibration results and accuracy verification in the inland test site: (<b>a</b>) corrected DEM; (<b>b</b>) error statistical histogram of the corrected DEMs by block adjustment considering and neglecting atmospheric effects. (<b>c</b>–<b>e</b>) are partially enlarged views of the red frame in (<b>a</b>): (<b>c</b>) partial map of the corrected DEM; (<b>d</b>) difference map between (<b>c</b>) and TanDEM-X DEM; (<b>e</b>) difference map between the partial DEM ignoring atmospheric effects and TanDEM-X DEM.</p>
Full article ">Figure 8
<p>(<b>a</b>) The control point database of the coastal test site; (<b>b</b>) the distribution of GCPs and TPs in the coastal test site.</p>
Full article ">Figure 9
<p>The atmospheric correction results of the coastal test site: (<b>a</b>–<b>d</b>) ATP estimated by the SA-MRWCA; (<b>e</b>–<b>h</b>) residual phase after atmospheric correction.</p>
Full article ">Figure 10
<p>DEM calibration results and accuracy verification in the coastal test site: (<b>a</b>) corrected DEM; (<b>b</b>) error statistical histogram of the corrected DEMs by block adjustment considering and neglecting atmospheric effects. (<b>c</b>–<b>e</b>) are partially enlarged views of the red frame in (<b>a</b>): (<b>c</b>) partial map of the corrected DEM; (<b>d</b>) difference map between (<b>c</b>) and TanDEM-X DEM; (<b>e</b>) difference map between the partial DEM ignoring atmospheric effects and TanDEM-X DEM.</p>
Full article ">Figure 11
<p>Comparison of systematic errors between scene-by-scene fitting and block adjustment fitting in the inland test site: (<b>a</b>–<b>d</b>) scene-by-scene fitting; (<b>e</b>–<b>h</b>) block adjustment fitting; (<b>i</b>–<b>l</b>) difference maps in estimated systematic errors between scene-by-scene fitting and block adjustment fitting.</p>
Full article ">Figure 12
<p>Comparison between the original DEM, corrected DEM, TanDEM-X DEM, SRTM DEM, and ASTER GDEM at varying slopes: (<b>a</b>,<b>c</b>,<b>e</b>) the RMSE, MAE, and STD of different DEMs in inland test site; (<b>b</b>,<b>d</b>,<b>f</b>) the RMSE, MAE, and STD of different DEMs in coastal test site.</p>
Full article ">
22 pages, 14092 KiB  
Article
Spaceborne Radars for Mapping Surface and Subsurface Salt Pan Configuration: A Case Study of the Pozuelos Salt Flat in Northern Argentina
by José Manuel Lattus, Matías Ernesto Barber, Dražen Skoković, Waldo Pérez-Martínez, Verónica Rocío Martínez and Laura Flores
Remote Sens. 2024, 16(8), 1411; https://doi.org/10.3390/rs16081411 - 16 Apr 2024
Viewed by 1485
Abstract
Lithium mining has become a controversial issue in the transition to green technologies due to the intervention in natural basins that impact the native flora and fauna in these environments. Large resources of this element are concentrated in Andean salt flats in South [...] Read more.
Lithium mining has become a controversial issue in the transition to green technologies due to the intervention in natural basins that impact the native flora and fauna in these environments. Large resources of this element are concentrated in Andean salt flats in South America, where extraction is much easier than in other geological configurations. The Pozuelos highland salt flat, located in northern Argentina (Salta’s Province), was chosen for this study due to the presence of different evaporitic crusts and its proven economic potential in lithium-rich brines. A comprehensive analysis of a 5.5-year-long time series of its microwave backscatter with Synthetic Aperture Radar (SAR) images yielded significant insights into the dynamics of their crusts. During a field campaign conducted near the acquisition of three SAR images (Sentinel-1, ALOS-2/PALSAR-2, and SAOCOM-1), field measurements were collected for computational modeling of the SAR response. The temporal backscattering coefficients for the crusts in the salt flat are directly linked to rainfall events, where changes in surface roughness, soil moisture, and water table depth represent the most critical variables. Field parameters were employed to model the backscattering response of the salt flat using the Small Slope Approximation (SSA) model. Salt concentration of the subsurface brine and the water table depth over the slightly to moderately roughed crusts were quantitatively derived from Bayesian inference of the ALOS-2/PALSAR-2 and SAOCOM-1 SAR backscattering coefficient data. The results demonstrated the potential for subsurface estimation with L-band dual-polarization images, constrained to crusts compatible with the feasibility range of the layered model. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the Pozuelos salt flat study area. A zoomed-in area of the Pozuelos salt flat is presented in a true color Sentinel-2 product acquired on 30 May 2023, with the location (1 to 16) of the trenches.</p>
Full article ">Figure 2
<p>(<b>a</b>) Illustration of fieldwork on the trench at sampling site ID-1. (<b>b</b>) Photograph sample for surface roughness calculation at ID-16.</p>
Full article ">Figure 3
<p>Monthly availability of SAR images between January 2018 and May 2023. Numbers on circles indicate the number of images for each sensor.</p>
Full article ">Figure 4
<p>Methodology flowchart for Sentinel-1, ALOS-2/PALSAR-2, and SAOCOM-1 scenes.</p>
Full article ">Figure 5
<p>Crust differentiation in the Pozuelos salt flat. (<b>a</b>) Crust classification over true color image (based on [<a href="#B31-remotesensing-16-01411" class="html-bibr">31</a>]), (<b>b</b>) Sentinel-2 true color image acquired on 30 May 2023, (<b>c</b>) Sentinel-1 SAR image (VV) acquired on 25 May 2023, and (<b>d</b>) the corresponding despeckled image with a 17 × 17 Lee Sigma filter.</p>
Full article ">Figure 6
<p>Four major crust types gathered at fieldwork. IDs refer to the trench locations in <a href="#remotesensing-16-01411-f001" class="html-fig">Figure 1</a>. Insets depict crust roughness by comparison with the gridded board. (<b>a</b>) Type I at ID-1. (<b>b</b>) Type II at ID-9. (<b>c</b>) Type III at ID-7. (<b>d</b>) Type IV at ID-6.</p>
Full article ">Figure 7
<p>Two-layer (surface and subsurface) model for the Pozuelos salt flat.</p>
Full article ">Figure 8
<p>Sentinel-2 color infrared composition showing the water dynamics preceding and following heavy rainfall events. (<b>a</b>–<b>c</b>) correspond to a 31 mm rainfall accumulated between 2 January 2019 and 4 January 2019. (<b>d</b>–<b>f</b>) correspond to a 30 mm rainfall accumulated between 14 January 2020 to 21 January 2020. (<b>g</b>–<b>i</b>) correspond to a 69 mm rainfall accumulated from 19 December 2021 to 27 December 2021.</p>
Full article ">Figure 9
<p>Temporal evolution of co-polarized backscatter responses of selected crusts in the Pozuelos salt flat for Sentinel-1 VV (S1), ALOS-2/PALSAR-2 HH (A2P2), and SAOCOM-1 HH (SC) and daily rainfall [mm].</p>
Full article ">Figure 10
<p>Temporal evolution of cross-polarized backscatter responses of selected crusts in the Pozuelos salt flat for Sentinel-1 VH (S1), ALOS-2/PALSAR-2 HV (A2P2), and SAOCOM-1 HV (SC) and daily rainfall [mm].</p>
Full article ">Figure 11
<p>Backscattering coefficients in test samples for Sentinel-1 (S1) and ALOS-2/PALSAR-2 (A2P2) and water table depths. For ID-6, the excavator could only dig up to 70 cm, due to the hardness of the soil, without reaching the water table.</p>
Full article ">Figure 12
<p>Two-layer SSA model contour levels for VV-polarized Sentinel-1 (blue) and HH-polarized SAOCOM-1 (black) backscattering coefficient. Dotted contours correspond to the measured backscattering coefficients given by <a href="#remotesensing-16-01411-t003" class="html-table">Table 3</a>. Red crosses correspond to the in situ measurements. (<b>a</b>) ID-1, (<b>b</b>) ID-7, (<b>c</b>) ID-16.</p>
Full article ">Figure 13
<p>Posterior distribution sampled using an MCMC algorithm for HH and HV polarization with the SSA model. Red cross refers to the measurements at the corresponding sampling locations and black cross to the Q2 quartile. (<b>a</b>) ID-1 (ALOS-2/PALSAR-2), (<b>b</b>) ID-7 (ALOS-2/PALSAR-2), (<b>c</b>) ID-16 (ALOS-2/PALSAR-2), (<b>d</b>) ID-1 (SAOCOM-1), (<b>e</b>) ID-7 (SAOCOM-1), and (<b>f</b>) ID-16 (SAOCOM-1).</p>
Full article ">
18 pages, 41235 KiB  
Article
Exploring the InSAR Deformation Series Using Unsupervised Learning in a Built Environment
by Mengshi Yang, Menghua Li, Cheng Huang, Ruisi Zhang and Rui Liu
Remote Sens. 2024, 16(8), 1375; https://doi.org/10.3390/rs16081375 - 13 Apr 2024
Cited by 1 | Viewed by 915
Abstract
As a city undergoes large-scale construction and expansion, there is an urgent need to monitor the stability of the ground and infrastructure. The time-series InSAR technique is an effective tool for measuring surface displacements. However, interpreting these displacements in a built environment, where [...] Read more.
As a city undergoes large-scale construction and expansion, there is an urgent need to monitor the stability of the ground and infrastructure. The time-series InSAR technique is an effective tool for measuring surface displacements. However, interpreting these displacements in a built environment, where observed displacements consist of mixed signals, poses a challenge. This study uses principal component analysis (PCA) and the k-means clustering method for exploring deformation series within an unsupervised learning context. The PCA method extracts the dominant components in deformation series, whereas the clustering method identifies similar deformation series. This method was tested on Kunming City (KMC) using C-band Sentinel-1, X-band TerraSAR-X, and L-band ALOS-2 PALSAR-2 data acquired between 2017 to 2022. The experiment demonstrated that the suggested unsupervised learning approach can group PS points with similar kinematic characteristics. Five types of deformation kinematic characteristics were discovered in the three SAR datasets: upward, slight upward, stability, slight downward, and downward. According to the results, less than 20% of points exhibit significant motion trends, whereas 50% show small velocity values but still demonstrate movement trends. The remaining 30% are relatively stable. Similar clustering results were obtained from the three datasets using unsupervised methods, highlighting the effectiveness of identifying spatial–temporal patterns over the study area. Moreover, It was found that clustering based on kinematic characteristics enhances the interpretation of InSAR deformation, particularly for points with small deformation velocities. Finally, the significance of PCA decomposition in interpreting InSAR deformation was discussed, as it can better represent series with noise, enabling their accurate identification. Full article
(This article belongs to the Special Issue Imaging Geodesy and Infrastructure Monitoring II)
Show Figures

Figure 1

Figure 1
<p>The left column of the figure displays the building heights in Kunming City’s central districts using the CNBH-10 product [<a href="#B37-remotesensing-16-01375" class="html-bibr">37</a>]. The top right figure shows Kunming City’s location in southwestern China. The bottom right figure presents the collected SAR coverages, depicted with Kunming’s topography.</p>
Full article ">Figure 2
<p>The data processing workflow in this study includes two parts: (<b>a</b>) time series InSAR analysis, including four blocks: S: PS selection, U: phase unwrapping, A: atmospheric mitigation, and O: output; (<b>b</b>) unsupervised learning InSAR deformation kinematic characteristics, including two parts: D: PCA decomposition and C: clustering.</p>
Full article ">Figure 3
<p>Deformation velocity map of the KMC area, estimated from Sentinel-1, TerraSAR-X, and PALSAR-2 images. The deformation velocities are color-coded from the range of −12 to 12 mm/yr. The white circle indicates the position of the reference point.</p>
Full article ">Figure 4
<p>Land cover fine classification map of KMC area [<a href="#B48-remotesensing-16-01375" class="html-bibr">48</a>]: (<b>a</b>) 2010; (<b>b</b>) 2020; (<b>c</b>) the changes between 2010 and 2020.</p>
Full article ">Figure 5
<p>Panels (<b>a</b>,<b>e</b>) display detailed maps of surface changes in areas A and B and areas C and D, respectively; (<b>b</b>,<b>f</b>) depict Sentinel-1 deformation results for areas A and B and areas C and D; (<b>c</b>,<b>g</b>) show detailed TerraSAR-X deformation results for areas A and B and areas C and D; (<b>d</b>,<b>h</b>) present ALOS-2 PALSAR-2 deformation results for areas A and B and areas C and D.</p>
Full article ">Figure 6
<p>Panel (<b>a</b>) displays the clustering result from the unsupervised learning of the time series InSAR deformation sequence that is estimated from Sentinel-1 data. The color indicates the category, and the average time series deformation result for each category is depicted in (<b>b</b>). Panel (<b>c</b>) represents the violin plot of the velocity distribution for corresponding categories of PS points.</p>
Full article ">Figure 7
<p>Panel (<b>a</b>) displays the clustering result from the unsupervised learning of the time series InSAR deformation sequence that is estimated from TerraSAR-X data. The color indicates the category, and the average time series deformation result for each category is depicted in (<b>b</b>). Panel (<b>c</b>) represents the violin plot of the velocity distribution for corresponding categories of PS points.</p>
Full article ">Figure 8
<p>Panel (<b>a</b>) displays the clustering result from the unsupervised learning of the time series InSAR deformation sequence that is estimated from PALSAR-2 data. The color indicates the category, and the average time series deformation result for each category is depicted in (<b>b</b>). Panel (<b>c</b>) represents the violin plot of the velocity distribution for corresponding categories of PS points.</p>
Full article ">Figure 9
<p>Ten sequences randomly selected from the Sentinel-1 deformation series and the results of their PCA decomposition.</p>
Full article ">
20 pages, 10049 KiB  
Article
Ground Subsidence, Driving Factors, and Risk Assessment of the Photovoltaic Power Generation and Greenhouse Planting (PPG&GP) Projects in Coal-Mining Areas of Xintai City Observed from a Multi-Temporal InSAR Perspective
by Chao Ding, Guangcai Feng, Zhiqiang Xiong and Lu Zhang
Remote Sens. 2024, 16(6), 1109; https://doi.org/10.3390/rs16061109 - 21 Mar 2024
Cited by 1 | Viewed by 969
Abstract
In recent years, photovoltaic power generation and greenhouse planting (PPG&GP) have become effective approaches for reconstructing and restoring the ecological environment of old coal-mining industry bases, such as Xintai City. However, the ecological impacts or improvements of the PPG&GP projects and their daily [...] Read more.
In recent years, photovoltaic power generation and greenhouse planting (PPG&GP) have become effective approaches for reconstructing and restoring the ecological environment of old coal-mining industry bases, such as Xintai City. However, the ecological impacts or improvements of the PPG&GP projects and their daily operations on the local environment are still unclear. To solve these problems, this study retrieved the ground deformation velocities and time series of the study region by performing the Small-Baseline Subset (SBAS)-Interferometric Synthetic Aperture Radar (InSAR) technique on the Advanced Land Observing Satellite (ALOS) PALSAR and Sentinel-1 SAR datasets. With these deformation results, the spatial analysis indicated that the area of the subsidence region within the PPG&GP projects reached 10.70 km2, with a magnitude of approximately −21.61 ± 12.10 mm/yr. Also, even though the ground deformations and their temporal changes were both visible in the construction and operation stages of the PPG&GP projects, the temporal analysis demonstrated that most observation points finally entered into the stationary phases in the late stage of the observation period. This phenomenon validated the effectiveness of the PPG&GP projects in enhancing the ground surface stability in coal-mining areas. Additionally, the precipitation, geological structure, increased coal-mining depths, and emergent agricultural modes were assumed to be the major impact factors controlling the ground deformation within the local PPG&GP projects. Finally, a novel risk assessment method with a designed index of IRA was utilized to classify the ground subsidence risks of the PPG&GP projects into three levels: Low (69.7%), Medium (16.9%), and High (9.4%). This study sheds a bright light on the ecological monitoring and risk management of the burgeoning industrial and agricultural infrastructures, such as the PPG&GP projects, constructed upon the traditional coal-mining areas in China from a multi-temporal InSAR perspective. Full article
Show Figures

Figure 1

Figure 1
<p>The geographical and geological background of the study area in Xintai City, with the image coverages of ALOS PALSAR and Sentinel-1 SAR ascending observations. The coalfield boundary, faults, fault branches, and geological structure were revised from [<a href="#B12-remotesensing-16-01109" class="html-bibr">12</a>].</p>
Full article ">Figure 2
<p>The flow chart of the methodology was used to evaluate the risks of construction and operation of the photovoltaic power generation and greenhouse planting (PPG&amp;GP) projects in the coal-mining areas of Xintai City.</p>
Full article ">Figure 3
<p>The spatial and temporal baseline network of (<b>a</b>) Advanced Land Observing Satellite (ALOS) PALSAR images (spatial baseline threshold: 75~1200 m, temporal baseline threshold: 45–900 days) and (<b>b</b>) Sentinel-1 images (spatial baseline threshold: 0~200 m, temporal baseline threshold: 12~180 days) using SBAS-InSAR processing.</p>
Full article ">Figure 4
<p>The ALOS PALSAR-derived ground line-of-sight (LOS) deformation velocity (mm/yr) over the observation period of 17 January 2007~28 October 2010 in the coal-mining region of Xintai City.</p>
Full article ">Figure 5
<p>The Sentinel-1 derived ground line-of-sight (LOS) deformation velocity (mm/yr) over the observation period of 30 July 2015~22 August 2022 in the region of local photovoltaic power generation and greenhouse planting (PPG&amp;GP) projects in Xintai City.</p>
Full article ">Figure 6
<p>The ground deformation region (green areas) detected by (<b>a</b>) ALOS PALSAR over the observation period of 17 January 2007~28 October 2010 and (<b>b</b>) Sentinel-1 over the observation period of 30 July 2015~22 August 2022.</p>
Full article ">Figure 7
<p>The cumulative ALOS PALSAR (17 January 2007~28 October 2010) and Sentinel-1 (30 July 2015~22 August 2022) deformation time series (<b>a</b>–<b>l</b>) for 12 observation points of A~L located within the photovoltaic power generation and greenhouse planting (PPG&amp;GP) projects of Xintai City. The geological locations of these points were consistent with the ones in <a href="#remotesensing-16-01109-f004" class="html-fig">Figure 4</a> and <a href="#remotesensing-16-01109-f005" class="html-fig">Figure 5</a>. The ground deformation time series of A~L were also cross-compared with the local precipitation and nearby tectonic seismicity within 200 km. In each subgraph, the construction temporal period and the operating temporal period of the PPG&amp;GP projects are marked by the light red block and the light green block, respectively. The specific start and end times for the construction and operation of the PPG&amp;GP projects can be found in <a href="#remotesensing-16-01109-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 8
<p>The ALOS PALSAR (<b>a</b>) and Sentinel-1 (<b>b</b>) derived ground deformation time series along the Profile of PP’, compared with corresponding geological column map (<b>c</b>) revised from [<a href="#B12-remotesensing-16-01109" class="html-bibr">12</a>]. The vertical axis of (<b>a</b>,<b>b</b>) from top to bottom denote the time sequence from 17 January 2007 to 28 October 2010 for ALOS PALSAR (<b>a</b>) and the time sequence from 30 July 2015 to 22 August 2022 for Sentinel-1 (<b>b</b>), respectively.</p>
Full article ">Figure 9
<p>The risk assessment map of the photovoltaic power generation and greenhouse planting (PPG&amp;GP) projects constructed in the coal-mining areas of Xintai City, which mainly evaluated the deformation velocities and deformation time series derived from the recent Sentinel-1 observations from 30 July 2015 to 22 August 2022.</p>
Full article ">
13 pages, 3387 KiB  
Technical Note
Polarimetric Measures in Biomass Change Prediction Using ALOS-2 PALSAR-2 Data
by Henrik J. Persson and Ivan Huuva
Remote Sens. 2024, 16(6), 953; https://doi.org/10.3390/rs16060953 - 8 Mar 2024
Viewed by 1224
Abstract
The use of multiple synthetic aperture radar polarizations can improve biomass estimations compared to using a single polarization. In this study, we compared predictions of aboveground biomass change from ALOS-2 PALSAR-2 backscatter using linear regression based on (1) the cross-polarization channels, (2) the [...] Read more.
The use of multiple synthetic aperture radar polarizations can improve biomass estimations compared to using a single polarization. In this study, we compared predictions of aboveground biomass change from ALOS-2 PALSAR-2 backscatter using linear regression based on (1) the cross-polarization channels, (2) the co- and cross-polarizations from fully polarimetric SAR, (3) Freeman–Durden polarimetric decomposition, and (4) the polarimetric radar vegetation index (RVI). Additionally, the impact of forest structure on the sensitivity of the polarimetric backscatter to AGB and AGB change was assessed. The biomass consisted of mainly coniferous trees at the hemi-boreal test site Remningstorp, located in southern Sweden. We found some improvements in the predictions when quad-polarized data (RMSE = 79.4 tons/ha) were used instead of solely cross-polarized data (RMSE = 84.9 tons/ha). However, when using Freeman–Durden decomposition, the prediction accuracy improved further (RMSE = 69.7 tons/ha), and the highest accuracy was obtained with the radar vegetation index (RMSE = 50.4 tons/ha). The corresponding R2 values ranged from 0.45 to 0.82. The bias was less than 1 t/ha for all models. An analysis of forest variables showed that the sensitivity to AGB was reduced for high values of basal-area-weighted mean height, basal area, and stem density when predicting absolute AGB, but the best change prediction model was sensitive to changes larger than the apparent saturation point for AGB state estimates. We conclude that by using fully polarimetric SAR images, forest biomass changes can be estimated more accurately compared to using single- or dual-polarization images. The results were improved the most (in terms of RMSE and R2) by using the Freeman–Durden decomposition model or the RVI, which captured especially the large changes better. Full article
(This article belongs to the Special Issue SAR for Forest Mapping III)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Extent of satellite scene in orange and location of test site, Remningstorp, in red. (<b>b</b>) Plot locations of the 46 field plots in red.</p>
Full article ">Figure 2
<p>Scatter plots of predicted vs. reference (field) AGB change between 2015 and 2021. (<b>a</b>) Model 1—based on cross-polarization. (<b>b</b>) Model 2—based on cross- and co-polarizations. (<b>c</b>) Model 3—Freeman–Durden-based. (<b>d</b>) Model 4—RVI-based.</p>
Full article ">Figure 3
<p>Scatter plots of predicted vs. reference (field) AGB change for model 4—RVI—with colors representing different ranges of structure variables. (<b>a</b>) Lorey’s height, (<b>b</b>) basal area, (<b>c</b>) stems per hectare. Values of structure variables are from the 2014 survey.</p>
Full article ">Figure 4
<p>Scatter plots of predicted vs. reference (field) AGB 2014. (<b>a</b>) Model 1—based on cross-polarization. (<b>b</b>) Model 2—based on cross- and co-polarizations. (<b>c</b>) Model 3—Freeman–Durden-based. (<b>d</b>) Model 4—based on RVI.</p>
Full article ">Figure 5
<p>Scatter plots of predicted vs. reference (field) AGB 2014 for model 4—RVI—with colors representing different ranges of structure variables. (<b>a</b>) Lorey’s height, (<b>b</b>) basal area, (<b>c</b>) stems per hectare.</p>
Full article ">
Back to TopTop