[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

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
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
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,550)

Search Parameters:
Keywords = Sentinel-1A

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
21 pages, 33427 KiB  
Article
MSSFNet: A Multiscale Spatial–Spectral Fusion Network for Extracting Offshore Floating Raft Aquaculture Areas in Multispectral Remote Sensing Images
by Haomiao Yu, Yingzi Hou, Fangxiong Wang, Junfu Wang, Jianfeng Zhu and Jianke Guo
Sensors 2024, 24(16), 5220; https://doi.org/10.3390/s24165220 (registering DOI) - 12 Aug 2024
Abstract
Accurately extracting large-scale offshore floating raft aquaculture (FRA) areas is crucial for supporting scientific planning and precise aquaculture management. While remote sensing technology offers advantages such as wide coverage, rapid imaging, and multispectral capabilities for FRA monitoring, the current methods face challenges in [...] Read more.
Accurately extracting large-scale offshore floating raft aquaculture (FRA) areas is crucial for supporting scientific planning and precise aquaculture management. While remote sensing technology offers advantages such as wide coverage, rapid imaging, and multispectral capabilities for FRA monitoring, the current methods face challenges in terms of establishing spatial–spectral correlations and extracting multiscale features, thereby limiting their accuracy. To address these issues, we propose an innovative multiscale spatial–spectral fusion network (MSSFNet) designed specifically for extracting offshore FRA areas from multispectral remote sensing imagery. MSSFNet effectively integrates spectral and spatial information through a spatial–spectral feature extraction block (SSFEB), significantly enhancing the accuracy of FRA area identification. Additionally, a multiscale spatial attention block (MSAB) captures contextual information across different scales, improving the ability to detect FRA areas of varying sizes and shapes while minimizing edge artifacts. We created the CHN-YE7-FRA dataset using Sentinel-2 multispectral remote sensing imagery and conducted extensive evaluations. The results showed that MSSFNet achieved impressive metrics: an F1 score of 90.76%, an intersection over union (IoU) of 83.08%, and a kappa coefficient of 89.75%, surpassing those of state-of-the-art methods. The ablation results confirmed that the SSFEB and MSAB modules effectively enhanced the FRA extraction accuracy. Furthermore, the successful practical applications of MSSFNet validated its generalizability and robustness across diverse marine environments. These findings highlight the performance of MSSFNet in both experimental and real-world scenarios, providing reliable, precise FRA area monitoring. This capability provides crucial data for scientific planning and environmental protection purposes in coastal aquaculture zones. Full article
(This article belongs to the Special Issue Application of Satellite Remote Sensing in Geospatial Monitoring)
21 pages, 42176 KiB  
Article
Application of Getis-Ord Correlation Index (Gi) for Burned Area Detection Improvement in Mediterranean Ecosystems (Southern Italy and Sardinia) Using Sentinel-2 Data
by Antonio Lanorte, Gabriele Nolè and Giuseppe Cillis
Remote Sens. 2024, 16(16), 2943; https://doi.org/10.3390/rs16162943 (registering DOI) - 12 Aug 2024
Viewed by 162
Abstract
This study collects the results obtained using the Getis-Ord local spatial autocorrelation index (Gi) with the aim of improving the classification of burned area detection maps generated from spectral indices (i.e., dNBR index) derived from Sentinel-2 satellite data. Therefore, the work proposes an [...] Read more.
This study collects the results obtained using the Getis-Ord local spatial autocorrelation index (Gi) with the aim of improving the classification of burned area detection maps generated from spectral indices (i.e., dNBR index) derived from Sentinel-2 satellite data. Therefore, the work proposes an adaptive thresholding approach that also includes the application of a similarity index (Sorensen–Dice Similarity Index) with the aim of adaptively correcting classification errors (false-positive burned pixels) related to the spectral response of burned/unburned areas. In this way, two new indices derived from the application of the Getis-Ord local autocorrelation analysis were created to test their effectiveness. Three wildfire events were considered, two of which occurred in Southern Italy in the summer of 2017 and one in Sardinia in the summer of 2019. The accuracy assessment analysis was carried out using the CEMS (Copernicus Emergency Management Service) on-demand maps. The results show the remarkable performance of the two new indices in terms of their ability to reduce the false positives generated by dNBR. In the three sites considered, the false-positive reduction percentage was around 95–96%. The proposed approach seems to be adaptable to different vegetation contexts, and above all, it could be a useful tool for mapping burned areas to support post-fire management activities. Full article
Show Figures

Figure 1

Figure 1
<p>Location and perimeter of the burned areas analysed as provided by CEMS.</p>
Full article ">Figure 2
<p>Workflow of the proposed approach.</p>
Full article ">Figure 3
<p>NBR pre-fire, NBR post-fire, and dNBR maps for Brienza fire.</p>
Full article ">Figure 4
<p>NBR pre-fire, NBR post-fire, and dNBR maps for San Fili-Rende fire.</p>
Full article ">Figure 5
<p>NBR pre-fire, NBR post-fire. and dNBR maps for Tanca-Altara fire.</p>
Full article ">Figure 6
<p>Area of Interest as reported by CEMS (inside the white line) and Region of interest used in the present study (red areas with black squares). On the right are the same views but at a higher zoom level.</p>
Full article ">Figure 7
<p>Comparison of burned areas (red) from CEMS (on the left) with dNBRGi, dGiNBR, and dNBR for the three study cases.</p>
Full article ">Figure 8
<p>Comparison of the dNBR indices dNBRGi and dGiNBR with the reference burned area as reported by CEMS. Highlighting of false positives and negatives.</p>
Full article ">Figure 9
<p>The red colour highlights the burned areas as reported by CEMS (San Fili) and as calculated in the study. The figure shows the improvement obtained by applying one of the indices compared to dNBR.</p>
Full article ">Figure 10
<p>An example of a persistent false-positive has been highlighted within the white circle corresponding to an area of dry vegetation (San Fili). This area was not classified as burned by CEMS (<b>left</b>) but was present in the dNBR (<b>centre</b>) and indices (<b>right</b>) developed in this study.</p>
Full article ">
23 pages, 2702 KiB  
Article
Modelling Soil Moisture Content With Hydrus 2D in a Conti-Nental Climate for Effective Maize Irrigation Planning
by Nxumalo Gift Siphiwe, Tamás Magyar, János Tamás and Attila Nagy
Agriculture 2024, 14(8), 1340; https://doi.org/10.3390/agriculture14081340 - 10 Aug 2024
Viewed by 381
Abstract
In light of climate change and limited water resources, optimizing water usage in agriculture is crucial. This study models water productivity to help regional planners address these challenges. We integrate CROPWAT-based reference evapotranspiration (ETo) with Sentinel 2 data to calculate daily [...] Read more.
In light of climate change and limited water resources, optimizing water usage in agriculture is crucial. This study models water productivity to help regional planners address these challenges. We integrate CROPWAT-based reference evapotranspiration (ETo) with Sentinel 2 data to calculate daily evapotranspiration and water needs for maize using soil and climate data from 2021 to 2023. The HYDRUS model predicted volumetric soil moisture content, validated against observed data. A 2D hydrodynamic model within HYDRUS simulated temporal and spatial variations in soil water distribution for maize at a non-irrigated site in Hungary. The model used soil physical properties and crop evapotranspiration rates as inputs, covering crop development stages from planting to harvest. The model showed good performance, with R² values of 0.65 (10 cm) and 0.81 (60 cm) in 2021, 0.51 (10 cm) and 0.50 (60 cm) in 2022, and 0.38 (10 cm) and 0.72 (60 cm) in 2023. RMSE and NRMSE values indicated reliability. The model revealed water deficits and proposed optimal irrigation schedules to maintain soil moisture between 32.2 and 17.51 V/V%. This integrated approach offers a reliable tool for monitoring soil moisture and developing efficient irrigation systems, aiding maize production’s adaptation to climate change. Full article
(This article belongs to the Section Agricultural Water Management)
26 pages, 14290 KiB  
Article
Exploratory Analysis Using Deep Learning for Water-Body Segmentation of Peru’s High-Mountain Remote Sensing Images
by William Isaac Perez-Torres, Diego Armando Uman-Flores, Andres Benjamin Quispe-Quispe, Facundo Palomino-Quispe, Emili Bezerra, Quefren Leher, Thuanne Paixão and Ana Beatriz Alvarez
Sensors 2024, 24(16), 5177; https://doi.org/10.3390/s24165177 (registering DOI) - 10 Aug 2024
Viewed by 388
Abstract
High-mountain water bodies represent critical components of their ecosystems, serving as vital freshwater reservoirs, environmental regulators, and sentinels of climate change. To understand the environmental dynamics of these regions, comprehensive analyses of lakes across spatial and temporal scales are necessary. While remote sensing [...] Read more.
High-mountain water bodies represent critical components of their ecosystems, serving as vital freshwater reservoirs, environmental regulators, and sentinels of climate change. To understand the environmental dynamics of these regions, comprehensive analyses of lakes across spatial and temporal scales are necessary. While remote sensing offers a powerful tool for lake monitoring, applications in high-mountain terrain present unique challenges. The Ancash and Cuzco regions of the Peruvian Andes exemplify these challenges. These regions harbor numerous high-mountain lakes, which are crucial for fresh water supply and environmental regulation. This paper presents an exploratory examination of remote sensing techniques for lake monitoring in the Ancash and Cuzco regions of the Peruvian Andes. The study compares three deep learning models for lake segmentation: the well-established DeepWaterMapV2 and WatNet models and the adapted WaterSegDiff model, which is based on a combination of diffusion and transformation mechanisms specifically conditioned for lake segmentation. In addition, the Normalized Difference Water Index (NDWI) with Otsu thresholding is used for comparison purposes. To capture lakes across these regions, a new dataset was created with Landsat-8 multispectral imagery (bands 2–7) from 2013 to 2023. Quantitative and qualitative analyses were performed using metrics such as Mean Intersection over Union (MIoU), Pixel Accuracy (PA), and F1 Score. The results achieved indicate equivalent performance of DeepWaterMapV2 and WatNet encoder–decoder architectures, achieving adequate lake segmentation despite the challenging geographical and atmospheric conditions inherent in high-mountain environments. In the qualitative analysis, the behavior of the WaterSegDiff model was considered promising for the proposed application. Considering that WatNet is less computationally complex, with 3.4 million parameters, this architecture becomes the most pertinent to implement. Additionally, a detailed temporal analysis of Lake Singrenacocha in the Vilcanota Mountains was conducted, pointing out the more significant behavior of the WatNet model. Full article
(This article belongs to the Special Issue Application of Satellite Remote Sensing in Geospatial Monitoring)
Show Figures

Figure 1

Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>Landsat-8 scenes selected for study.</p>
Full article ">Figure 3
<p>Combining process from B2 to B7 into a single 6-channel image.</p>
Full article ">Figure 4
<p>From left to right: <math display="inline"><semantics> <mrow> <mo>(</mo> <mi>θ</mi> <mo>,</mo> <mi>ρ</mi> <mo>)</mo> </mrow> </semantics></math> parameter space, deskwed image, cropped image, and division of the image into 256 × 256 pixel patches.</p>
Full article ">Figure 5
<p>Mask creation process.</p>
Full article ">Figure 6
<p>WatNet model architecture.</p>
Full article ">Figure 7
<p>DeepWaterMapV2 model architecture based on 3 primary blocks.</p>
Full article ">Figure 8
<p>General architecture of WaterSegDiff based on a conditioning model and a diffusion model that integrate their information through two conditioning mechanisms, <math display="inline"><semantics> <mi mathvariant="script">U</mi> </semantics></math>-SA and SS-Former.</p>
Full article ">Figure 9
<p>SS-Former internal architecture consisting of two symmetrical cross-attention modules.</p>
Full article ">Figure 10
<p>Qualitative analysis of 5 selected samples that represent large lakes with compact structures. Showing the RGB image, ground truth, NDWI, WatNet, DeepWaterMapV2, and WaterSegDiff results. (<b>a</b>) Large and irregular lake, (<b>b</b>) two lakes with compact structure, (<b>c</b>) scene with river crossing, (<b>d</b>) large lake in mountainous region, (<b>e</b>) lake surrounded by dense vegetation.</p>
Full article ">Figure 11
<p>Qualitative analysis of 5 selected samples that represent small and dispersed lakes. Showing the RGB image, ground truth, NDWI, WatNet, DeepWaterMapV2, and WaterSegDiff results. (<b>a</b>,<b>b</b>) Snowy scene with shadows with presence of clear and turbid lakes, (<b>c</b>) completely snowy scene, (<b>d</b>,<b>e</b>) partially snowy area with scattered lakes.</p>
Full article ">Figure 12
<p>The edges extracted from Lake Singrenacocha based on NDWI, WatNet, DeepWaterMapV2, and WaterSegDiff. Highlights in yellow, green, blue, and red for the years 2014, 2016, 2018, and 2020, respectively.</p>
Full article ">Figure 13
<p>Graphical representation of the segmentation performance of Lake Singrenacocha during the years 2014, 2016, 2018, and 2020.</p>
Full article ">
18 pages, 4684 KiB  
Article
Monitoring Water Quality Parameters Using Sentinel-2 Data: A Case Study in the Weihe River Basin (China)
by Tieming Liu, Zhao Guo, Xiaoping Li, Teng Xiao, Jiaxin Liu and Yuanzhi Zhang
Sustainability 2024, 16(16), 6881; https://doi.org/10.3390/su16166881 (registering DOI) - 10 Aug 2024
Viewed by 386
Abstract
Based on Sentinel-2 multispectral image data and existing research results, the comprehensive water quality index (CWQI), NH4+-N, and total phosphorus (TP) in the Weihe River and its tributaries were estimated. Furthermore, a verified model was obtained by fitting the regression [...] Read more.
Based on Sentinel-2 multispectral image data and existing research results, the comprehensive water quality index (CWQI), NH4+-N, and total phosphorus (TP) in the Weihe River and its tributaries were estimated. Furthermore, a verified model was obtained by fitting the regression using the measured and inverted data. The verified model results show that the average relative error of the CWQI is only 9.80%, the goodness of fit of NH4+-N and TP concentrations is 0.62 and 0.61, respectively, and the average relative errors are 19.40% and 24.70%, respectively. The accuracy of the verified model is relatively high, and it can approximately invert the distribution of the three parameters of the Weihe River and its tributaries. In December 2023, except for the Bahe River between Puhua Town and Sanli Town in Lantian County, most of the water bodies in the Weihe River and its tributaries had good water quality. The study can provide an example of how to monitor water quality information using Sentinel-2 data in similar river basins. Full article
(This article belongs to the Section Sustainable Water Management)
Show Figures

Figure 1

Figure 1
<p>The distribution map of Xi’an City, Xianyang City, and the Weihe River and its tributaries.</p>
Full article ">Figure 2
<p>Water body distribution of the Weihe River and its tributaries in December 2023.</p>
Full article ">Figure 3
<p>Locations of measured sections (S1–S12), towns, and sewage treatment plants (red points indicate measured sections, green points indicate small towns, and black points indicate sewage treatment plants).</p>
Full article ">Figure 4
<p>Line chart of measured and inverted CWQI values.</p>
Full article ">Figure 5
<p>Line chart of measured and inverted NH<sub>4</sub><sup>+</sup>-N concentrations.</p>
Full article ">Figure 6
<p>Line chart of measured and inverted TP concentrations.</p>
Full article ">Figure 7
<p>Line chart of measured and verified CWQI for sections S8–S12.</p>
Full article ">Figure 8
<p>NH<sub>4</sub><sup>+</sup>-N regression results.</p>
Full article ">Figure 9
<p>TP regression results.</p>
Full article ">Figure 10
<p>Distribution of CWQI for the Weihe River and its tributaries in December 2023 estimated by the inversion model.</p>
Full article ">Figure 11
<p>Distribution of CWQI for the Weihe River and its tributaries in December 2023 obtained from the verified model (“a” represents the overall distribution of the river, and “b” represents the river segment in the area of Lantian County).</p>
Full article ">Figure 12
<p>Distribution of NH<sub>4</sub><sup>+</sup>-N for the Weihe River and its tributaries in December 2023 obtained from the inversion model.</p>
Full article ">Figure 13
<p>Distribution of NH<sub>4</sub><sup>+</sup>-N for the Weihe River and its tributaries in December 2023 obtained from the verified model (“c” represents the overall distribution of the river, and “d” represents the river segment in the area of Lantian County).</p>
Full article ">Figure 14
<p>Distribution of TP for the Weihe River and its tributaries in December 2023 obtained from the inversion model.</p>
Full article ">Figure 15
<p>Distribution of TP for the Weihe River and its tributaries in December 2023 obtained from the verified model (“e” represents the overall distribution of the river, and “f” represents the river segment in the area of Lantian County).</p>
Full article ">
20 pages, 2575 KiB  
Article
Recent Developments in Satellite Remote Sensing for Air Pollution Surveillance in Support of Sustainable Development Goals
by Dimitris Stratoulias, Narissara Nuthammachot, Racha Dejchanchaiwong, Perapong Tekasakul and Gregory R. Carmichael
Remote Sens. 2024, 16(16), 2932; https://doi.org/10.3390/rs16162932 (registering DOI) - 9 Aug 2024
Viewed by 209
Abstract
Air pollution is an integral part of climatic, environmental, and socioeconomic current affairs and a cross-cutting component of certain United Nations Sustainable Development Goals (SDGs). Hence, reliable information on air pollution and human exposure is a crucial element in policy recommendations and decisions. [...] Read more.
Air pollution is an integral part of climatic, environmental, and socioeconomic current affairs and a cross-cutting component of certain United Nations Sustainable Development Goals (SDGs). Hence, reliable information on air pollution and human exposure is a crucial element in policy recommendations and decisions. At the same time, Earth Observation is steadily gaining confidence as a data input in the calculation of various SDG indicators. The current paper focuses on the usability of modern satellite remote sensing in the context of SDGs relevant to air quality. We introduce the socioeconomic importance of air quality and discuss the current uptake of geospatial information. The latest developments in Earth Observation provide measurements of finer spatial, temporal, and radiometric resolution products with increased global coverage, long-term continuation, and coherence in measurements. Leveraging on the two latest operational satellite technologies available, namely the Sentinel-5P and the Geostationary Environment Monitoring Spectrometer (GEMS) missions, we demonstrate two potential operational applications for quantifying air pollution at city and regional scales. Based on the two examples and by discussing the near-future anticipated geospatial capabilities, we showcase and advocate that the potential of satellite remote sensing as a, complementary to ground station networks, source of air pollution information is gaining confidence. As such, it can be an invaluable tool for quantifying global air pollution and deriving robust population exposure estimates. Full article
Show Figures

Figure 1

Figure 1
<p>Settlements with available data on PM<sub>2.5</sub> concentrations between 2010 and 2019. Adapted from the World Health Organization [<a href="#B31-remotesensing-16-02932" class="html-bibr">31</a>] with permission.</p>
Full article ">Figure 2
<p>Global gridded map of the adjusted population count. Adapted from the Gridded Population of the World (GPWv4) dataset. Source: Center for International Earth Science Information Network—CIESIN—Columbia University [<a href="#B33-remotesensing-16-02932" class="html-bibr">33</a>].</p>
Full article ">Figure 3
<p>Global annual mean of geophysical PM<sub>2.5</sub> estimates for the year 2015 based on advances in satellite observations. Black dots represent ground stations. Adapted from Hammer et al. [<a href="#B23-remotesensing-16-02932" class="html-bibr">23</a>]. Source: <a href="https://pubs.acs.org/doi/10.1021/acs.est.0c01764" target="_blank">https://pubs.acs.org/doi/10.1021/acs.est.0c01764</a> (accessed on 30 July 2024). Further permissions related to the material excerpted should be directed to the ACS.</p>
Full article ">Figure 4
<p>Annual mean of the tropospheric vertical column of NO<sub>2</sub> for the year 2021 retrieved from Sentinel-5P satellite over Bangkok, Thailand. The blue dots represent the locations of the regulatory-grade ground stations available in this region. The layers are superimposed over a natural-color satellite image of the city of Bangkok.</p>
Full article ">Figure 5
<p>An operational product from the GEMS instrument: estimated surface PM<sub>2.5</sub> concentrations over Asia acquired on 25 February 2022 (retrieved from the NIER) (<b>left</b>) and monthly mean GEMS AOD (550 nm) image for March 2023 (<b>right</b>).</p>
Full article ">Figure 6
<p>Reprocessed monthly mean of NO<sub>2</sub> (<b>left</b>) and monthly maximum for SO<sub>2</sub> (<b>right</b>) from the operational data provided by GEMS for the month of November 2023.</p>
Full article ">
15 pages, 613 KiB  
Article
A Technological Perspective of Bringing Climate Change Adaptation, Disaster Risk Reduction, and Food Security Together in South Africa
by Annegrace Zembe, Livhuwani David Nemakonde, Paul Chipangura, Christo Coetzee and Fortune Mangara
Sustainability 2024, 16(16), 6844; https://doi.org/10.3390/su16166844 - 9 Aug 2024
Viewed by 569
Abstract
As disasters and climate change risks, particularly droughts and floods, continue to affect food security globally, most governments, including South Africa, have resorted to the use of technology to incorporate climate change adaptation and disaster risk reduction to address FS issues. This is [...] Read more.
As disasters and climate change risks, particularly droughts and floods, continue to affect food security globally, most governments, including South Africa, have resorted to the use of technology to incorporate climate change adaptation and disaster risk reduction to address FS issues. This is because most institutions and policies that address climate change adaptation, disaster risk reduction, and food security operate in parallel, which usually leads to the polarisation of interventions and conflicting objectives, thus leaving the issue of FS unresolved. The study aimed to investigate how food security projects are incorporating climate change adaptation and disaster risk reduction using technology. A qualitative research design was applied, whereby in-depth interviews were conducted with ten project participants from two projects, while 24 key informants were purposively selected from government and research institutions. The study’s main findings revealed that both projects incorporate climate change adaptation and disaster risk reduction measures in most of their food value chains. Although the projects are different, they still face similar challenges, such as a lack of expertise, resources, and funding, and an inadequate regulatory environment to improve their farming practices. The study brings in the practical side of addressing the coherence between food security, climate change adaptation, and disaster risk reduction through technology. Full article
(This article belongs to the Section Sustainable Agriculture)
Show Figures

Figure 1

Figure 1
<p>Rooftop farm situated on Chamber of Mines building in Johannesburg [<a href="#B64-sustainability-16-06844" class="html-bibr">64</a>].</p>
Full article ">
18 pages, 3947 KiB  
Article
Potential of the Bi-Static SAR Satellite Companion Mission Harmony for Land-Ice Observations
by Andreas Kääb, Jérémie Mouginot, Pau Prats-Iraola, Eric Rignot, Bernhard Rabus, Andreas Benedikter, Helmut Rott, Thomas Nagler, Björn Rommen and Paco Lopez-Dekker
Remote Sens. 2024, 16(16), 2918; https://doi.org/10.3390/rs16162918 - 9 Aug 2024
Viewed by 238
Abstract
The EarthExplorer 10 mission Harmony by the European Space Agency ESA, scheduled for launch around 2029–2030, consists of two passive C-band synthetic-aperture-radar companion satellites flying in a flexible constellation with one Sentinel-1 radar satellite as an illuminator. Sentinel-1 will serve as transmitter and [...] Read more.
The EarthExplorer 10 mission Harmony by the European Space Agency ESA, scheduled for launch around 2029–2030, consists of two passive C-band synthetic-aperture-radar companion satellites flying in a flexible constellation with one Sentinel-1 radar satellite as an illuminator. Sentinel-1 will serve as transmitter and receiver of radar waves, and the two Harmonys will serve as bistatic receivers without the ability to transmit. During the first and last year of the 5-year mission, the two Harmony satellites will fly in a cross-track interferometric constellation, such as that known from TanDEM-X, about 350 km ahead or behind the assigned Sentinel-1. This constellation will provide 12-day repeat DEMs, among other regions, over most land-ice and permafrost areas. These repeat DEMs will be complemented by synchronous lateral terrain displacements from the well-established offset tracking method. In between the cross-track interferometry phases, one of the Harmony satellites will be moved to the opposite side of the Sentinel-1 to form a symmetric bistatic “stereo” constellation with ±~350 km along-track baseline. In this phase, the mission will provide opportunity for radar interferometry along three lines of sight, or up to six when combining ascending and descending acquisitions, enabling the measurement of three-dimensional surface motion, for instance sub- and emergence components of ice flow, or three-dimensional deformation of permafrost surfaces or slow landslides. Such measurements would, for the first time, be available for large areas and are anticipated to provide a number of novel insights into the dynamics and mass balance of a range of mass movement processes. Full article
(This article belongs to the Special Issue Remote Sensing of the Cryosphere II)
Show Figures

Figure 1

Figure 1
<p>Graphical summary of Harmony land-ice measurement objectives.</p>
Full article ">Figure 2
<p>The Harmony mission consists of two satellites with one passive SAR instrument each, Harmony-A and Harmony-B. They will fly in two alternating configurations in convoy with Sentinel-1, which serves as a radar transmitter for the two receive-only Harmony satellites. (<b>a</b>) The stereo configuration is optimized to measure surface motion vectors on land and ocean and is foreseen for years 2–4 of the mission. (<b>b</b>) The cross-track interferometric (XTI) configuration is optimized to measure land surface topography every 12 days during years 1 and 5.</p>
Full article ">Figure 3
<p>Schematics of Harmony time series for the cases of (<b>a</b>) small elevation changes (few meters over several months to years) and of (<b>b</b>) large elevation changes (tens of meters) for a hypothetical glacier point. (<b>b</b>) shows combined elevation changes and horizontal speeds. During the full mission years 1 and 5, Harmony is foreseen to measure dozens of DEMs over glaciers and ice-sheet margins globally (black points with error bars). The blue and brown curves indicate hypothetical idealized glacier elevation variations over time. Note that the vertical axes of panel (<b>a</b>,<b>b</b>) have scales that are different by an order of magnitude. (<b>a</b>) The mission will be able to deliver glacier volume changes Δh from differencing the DEM stacks from years 1 and 5. For small glacier thickness changes, the penetration bias between real surface and a radar-interferometric DEM is substantial, relative to the expected elevation changes, and needs to be dealt with. (<b>b</b>) Harmony’s repeat DEMs, potentially combined with lateral surface displacements from radar offset tracking, can be used to study the interplay between changes in ice dynamics and ice thickness, such as for calving glaciers or glacier surges. Compared to the large elevation changes expected for such cases, radar penetration and DEM errors are relatively small and of less concern than in case (<b>a</b>).</p>
Full article ">Figure 4
<p>Visualization and comparison of some expected Harmony results using Tunabreen (78.46°N, 17.4°E), Svalbard, as an example. (<b>a</b>) Sentinel-1 amplitude image (gamma0, 12 March 2020). (<b>b</b>) Sentinel-1 12-day offsets over 19–31 January 2019. (<b>c</b>) Sentinel-2 image (infrared false color, for orientation only, 2 August 2019). (<b>d</b>) Elevation differences from Arctic-DEM [<a href="#B29-remotesensing-16-02918" class="html-bibr">29</a>] strips of 17 March 2015 and a mosaic of 11 and 15 March 2020 resampled to 100 m resolution with random noise of ±0.5 m/yr added. (<b>e</b>) Same as (<b>d</b>), but resampled to 50 m and with ±0.2 m/yr noise added. (<b>f</b>) Elevation trends from ASTER stacks 2015–2019 [<a href="#B24-remotesensing-16-02918" class="html-bibr">24</a>]. (<b>g</b>) TanDEM-X topographic change product, computed between TanDEM-X elevations compiled over 2010–2014 and elevations from 2017. Note, panels (<b>d</b>,<b>e</b>) are visualizations (not simulations) of Harmony XTI 5-year DEM differences for the threshold (panel (<b>d</b>)) and goal requirements (panel (<b>e</b>)). They do not include potential errors from SAR XTI processing (e.g., steep terrain, penetration), but rather errors from optical stereo processing (e.g., small clouds, lack of visual contrast). Black glacier outlines are from Randolph glacier inventory v7. Note also that the ASTER DEM differences are just shown for visual comparison, and that the potential reasons for the differences between the DEM differences from ASTER and Arctic-DEM are not the subject of the present contribution.</p>
Full article ">Figure 5
<p>Harmony’s interferometric line-of-sight diversity. (<b>a</b>) Sentinel-1 alone is able to provide radar interferometry along one line-of-sight (two if both ascending and descending orbits are combined; grey column to the right). (<b>b</b>) In XTI configuration, Harmony will provide two lines of sight (strictly speaking three, but the lines of sight of both Harmony satellites are very similar), and four if both ascending and descending orbits are combined. (<b>c</b>) In stereo configuration, Harmony will provide interferometry along three, or six, lines of sight, respectively. The three lines of sight per orbit in stereo configuration lie, though, approximately in one oblique plane. (<b>d</b>) Harmony interferograms from one orbit can also be combined with interferograms from an opposite orbit of the other Sentinel-1, providing four lines of sight.</p>
Full article ">Figure 6
<p>Interferometric measurements by the Harmony mission along two or more lines of sight will facilitate the measurement of three-dimensional glacier flow close to the surface. Such measurements will connect between the vertical component of ice flow, thickness changes over time, and local mass balance (i.e., directly measure components of the so-called kinematic boundary condition). Changes in the penetrated snow and firn pack could lead, though, to offsets between the true ice particle displacement and the displacement between the phase centers of the radar waves.</p>
Full article ">Figure 7
<p>2D schemes of different idealized cross-sections and kinematics of slow landslides. (<b>a</b>) surface slope steeper than bedding slope, (<b>b</b>) both similar, (<b>c</b>) bedding slope steeper than surface, (<b>d</b>) rotational landslide. Interferometric measurements from repeat Harmony data would enable estimating two- or three-dimensional surface velocities (red arrows). DEM stacks from Harmony’s XTI phases (or other DEMs) can give elevation changes (blue arrows). Surface at time 1, black line; surface at time 2, black dashed line; bedding plane, grey line.</p>
Full article ">Figure 8
<p>(<b>a</b>) Normalized differences between the two unit-less amplitude images of a TanDEM-X bi-static acquisition of 13 April 2015 over Aletsch Glacier (<b>b</b>) and Bernese Alps, Switzerland, with approx. 500 m along-track baseline and 2 km cross-track baseline, i.e., approx. 2060 m total baseline. The noise-filtered color-coded normalized differences are transparently laid over one of the amplitude images. The more blue or red, resp., the stronger the differences are. Image in raw radar geometry, flying direction from top to bottom (descending), look direction from right to left. Strongest variations in differences between the two amplitude images are found for higher elevations, likely related to variations in snow cover type. (For other examples, see Stefko et al. [<a href="#B81-remotesensing-16-02918" class="html-bibr">81</a>].)</p>
Full article ">
19 pages, 7874 KiB  
Article
Mapping the Fraction of Vegetation Coverage of Potamogeton crispus L. in a Shallow Lake of Northern China Based on UAV and Satellite Data
by Junjie Chen, Quanzhou Yu, Fenghua Zhao, Huaizhen Zhang, Tianquan Liang, Hao Li, Zhentan Yu, Hongli Zhang, Ruyun Liu, Anran Xu and Shaoqiang Wang
Remote Sens. 2024, 16(16), 2917; https://doi.org/10.3390/rs16162917 - 9 Aug 2024
Viewed by 351
Abstract
Under the background of global change, the lake water environment is facing a huge threat from eutrophication. The rapid increase in curly-leaf pondweed (Potamogeton crispus L.) in recent years has seriously threatened the ecological balance and the water diversion safety of the [...] Read more.
Under the background of global change, the lake water environment is facing a huge threat from eutrophication. The rapid increase in curly-leaf pondweed (Potamogeton crispus L.) in recent years has seriously threatened the ecological balance and the water diversion safety of the eastern route of China’s South-to-North Water Diversion Project. The monitoring and control of curly-leaf pondweed is imperative in shallow lakes of northern China. Unmanned Aerial Vehicles (UAVs) have great potential for monitoring aquatic vegetation. However, merely using satellite remote sensing to detect submerged vegetation is not sufficient, and the monitoring of UAVs on aquatic vegetation is rarely systematically evaluated. In this study, taking Nansi Lake as a case, we employed Red–Green–Blue (RGB) UAV and satellite datasets to evaluate the monitoring of RGB Vegetation Indices (VIs) in pondweed and mapped the dynamic patterns of the pondweed Fractional Vegetation Coverage (FVC) in Nansi Lake. The pondweed FVC values were extracted using the RGB VIs and the machine learning method. The extraction of the UAV RGB images was evaluated by correlations, accuracy assessments and separability. The correlation between VIs and FVC was used to invert the pondweed FVC in Nansi Lake. The RGB VIs were also calculated using Gaofen-2 (GF-2) and were compared with UAV and Sentinel-2 data. Our results showed the following: (1) The RGB UAV could effectively monitor the FVC of pondweed, especially when using Support Vector Machine that (SVM) has a high ability to recognize pondweed in UAV RGB images. Two RGB VIs, RCC and RGRI, appeared best suited for monitoring aquatic plants. The correlations between four RGB VIs based on GF-2, i.e., GCC, BRI, VDVI, and RGBVI and FVCSVM calculated by the UAV (p < 0.01) were better than those obtained with other RGB VIs. Thus, the RGB VIs of GF-2 were not as effective as those of the UAV in pondweed monitoring. (2) The binomial estimation model constructed by the Normalized Difference Water Index (NDWI) of Sentinel-2 showed a high accuracy (R2 = 0.7505, RMSE = 0.169) for pondweed FVC and can be used for mapping the FVC of pondweed in Nansi Lake. (3) Combined with the Sentinel-2 time-series data, we mapped the dynamic patterns of pondweed FVC in Nansi Lake. It was determined that the flooding of pondweed in Nansi Lake has been alleviated in recent years, but the rapid increase in pondweed in part of Nansi Lake remains a challenging management issue. This study provides practical tools and methodology for the innovative remote sensing monitoring of submerged vegetation. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area with the distribution of the sample sites. (<b>a</b>) Location of the study area; (<b>b</b>) detailed location of the study area; (<b>c</b>) sample points in the study area; (<b>d</b>) sample points in Dushan Lake; (<b>e</b>) sample points in Weishan Lake.</p>
Full article ">Figure 2
<p>A flowchart of this study.</p>
Full article ">Figure 3
<p>Illustration of RGB VI separability determination.</p>
Full article ">Figure 4
<p>FVC box plots extracted by different methods, including SVM, the Dimidiate Pixel Model and dynamic thresholding.</p>
Full article ">Figure 5
<p>The correlation coefficients between the different FVC values extracted by SVM, the Dimidiate Pixel Model and dynamic thresholding.</p>
Full article ">Figure 6
<p>(<b>a</b>) Correlation between remote sensing VIs, FVC<sub>NDVI</sub>, FVCsvm and the mean RGB VIs, (<b>b</b>) correlation between remote sensing VIs, FVC<sub>NDVI</sub>, FVCsvm and FVC values by UAV. “*” and “**” represent significant differences, with <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">Figure 7
<p>Correlation analysis between RGB VIs by GF-2 and the means of RGB VIs by the UAV (<b>a</b>), FVC by the UAV (<b>b</b>) and remote sensing VIs by Sentinel-2 (<b>c</b>). “*” and “**” represent significant differences, with <span class="html-italic">p</span> &lt; 0.05 and <span class="html-italic">p</span> &lt; 0.01, respectively.</p>
Full article ">Figure 8
<p>Accuracy assessment results for RGB VIs. (<b>a</b>) Overall accuracy, (<b>b</b>) F1 score.</p>
Full article ">Figure 9
<p>Statistical results of separability in the acquired images for RGB VIs.</p>
Full article ">Figure 10
<p>Comparison between estimated and measured pondweed FVC.</p>
Full article ">Figure 11
<p>Mapping pondweed FVC in Nansi Lake based on the NDWI binomial estimation model (14 May 2023).</p>
Full article ">Figure 12
<p>Seasonal change in pondweed FVC in Nansi Lake, 2023.</p>
Full article ">Figure 13
<p>Inter-annual changes in pondweed FVC in Nansi Lake, 2018–2023.</p>
Full article ">Figure 14
<p>Different growth periods of pondweed imaged by the RGB UAV.</p>
Full article ">
23 pages, 11056 KiB  
Article
Co-Kriging-Guided Interpolation for Mapping Forest Aboveground Biomass by Integrating Global Ecosystem Dynamics Investigation and Sentinel-2 Data
by Yingchen Wang, Hongtao Wang, Cheng Wang, Shuting Zhang, Rongxi Wang, Shaohui Wang and Jingjing Duan
Remote Sens. 2024, 16(16), 2913; https://doi.org/10.3390/rs16162913 - 9 Aug 2024
Viewed by 312
Abstract
Mapping wall-to-wall forest aboveground biomass (AGB) at large scales is critical for understanding global climate change and the carbon cycle. In previous studies, a regression-based method was commonly used to map the spatially continuous distribution of forest AGB with the aid of optical [...] Read more.
Mapping wall-to-wall forest aboveground biomass (AGB) at large scales is critical for understanding global climate change and the carbon cycle. In previous studies, a regression-based method was commonly used to map the spatially continuous distribution of forest AGB with the aid of optical images, which may suffer from the saturation effect. The Global Ecosystem Dynamics Investigation (GEDI) can collect forest vertical structure information with high precision on a global scale. In this study, we proposed a collaborative kriging (co-kriging) interpolation-based method for mapping spatially continuous forest AGB by integrating GEDI and Sentinel-2 data. First, by fusing spectral features from Sentinel-2 images with vertical structure features from GEDI, the optimal estimation model for footprint-level AGB was determined by comparing different machine-learning algorithms. Second, footprint-level predicted AGB was used as the main variable, with rh95 and B12 as covariates, to build a co-kriging guided interpolation model. Finally, the interpolation model was employed to map wall-to-wall forest AGB. The results showed the following: (1) For footprint-level AGB, CatBoost achieved the highest accuracy by fusing features from GEDI and Sentinel-2 data (R2 = 0.87, RMSE = 49.56 Mg/ha, rRMSE = 27.06%). (2) The mapping results based on the interpolation method exhibited relatively high accuracy and mitigated the saturation effect in areas with higher forest AGB (R2 = 0.69, RMSE = 81.56 Mg/ha, rRMSE = 40.98%, bias = −3.236 Mg/ha). The mapping result demonstrates that the proposed method based on interpolation combined with multi-source data can be a promising solution for monitoring spatially continuous forest AGB. Full article
(This article belongs to the Special Issue Remote Sensing and Lidar Data for Forest Monitoring)
Show Figures

Figure 1

Figure 1
<p>Location of the study area. (<b>a</b>) The administrative boundary of the State of California and the location of the study area within the state (i.e., where the red box is); (<b>b</b>) ground elevation distribution of the study area; (<b>c</b>) local distribution of GEDI footprints.</p>
Full article ">Figure 2
<p>Technology road map of this study.</p>
Full article ">Figure 3
<p>Filtered distribution of GEDI footprints: (<b>a</b>) urban area; (<b>b</b>) vegetation area. The green area represents vegetation, the purple area represents urban areas, and the blue area represents the ocean.</p>
Full article ">Figure 4
<p>The correlation between the features and forest AGB, with all significance levels of the selected features at 0.01.</p>
Full article ">Figure 5
<p>Feature selection outcomes. (<b>a</b>) Features selected using the stepwise method. (<b>b</b>) Features selected using the random forest method.</p>
Full article ">Figure 6
<p>Scatter plots between reference AGB and predicted AGB derived from the combined optical and GEDI data. The red dotted lines represent the fitted trend-lines.</p>
Full article ">Figure 7
<p>Histograms of AGB: (<b>a</b>) ALS-derived and (<b>b</b>) interpolation-method-derived.</p>
Full article ">Figure 8
<p>Boxplot of AGB distribution. IQR = Q3 − Q1, where Q1 is the first quartile and Q3 is the third quartile.</p>
Full article ">Figure 9
<p>Forest AGB mapping results. (<b>a</b>) Satellite images of the study area. (<b>b</b>) Forest AGB map derived from the co-kriging interpolation model. The red line represents the boundary of the study area (i.e., Sonoma County).</p>
Full article ">Figure 10
<p>Comparison of interpolation accuracy for different covariate combinations.</p>
Full article ">Figure 11
<p>Accuracy assessment of the wall-to-wall forest AGB products. (<b>a</b>) The predicted AGB derived from co-kriging interpolation. (<b>b</b>) The predicted AGB derived from the regression method. The red dotted lines represent the fitted trend-lines.</p>
Full article ">
23 pages, 19658 KiB  
Article
Mapping Irrigated Rice in Brazil Using Sentinel-2 Spectral–Temporal Metrics and Random Forest Algorithm
by Alexandre S. Fernandes Filho, Leila M. G. Fonseca and Hugo do N. Bendini
Remote Sens. 2024, 16(16), 2900; https://doi.org/10.3390/rs16162900 - 8 Aug 2024
Viewed by 303
Abstract
Brazil, a leading rice producer globally, faces challenges in systematically mapping its diverse rice fields due to varying cropping systems, climates, and planting calendars. Existing rice mapping methods often rely on complex techniques like deep learning or microwave imagery, posing limitations for large-scale [...] Read more.
Brazil, a leading rice producer globally, faces challenges in systematically mapping its diverse rice fields due to varying cropping systems, climates, and planting calendars. Existing rice mapping methods often rely on complex techniques like deep learning or microwave imagery, posing limitations for large-scale mapping. This study proposes a novel approach utilizing Sentinel-2 spectral–temporal metrics (STMs) in conjunction with a random forest classifier for rice paddy mapping. By extracting diverse STMs and training both regional and global classifiers, we validated the method across independent areas. While regional models tended to overestimate rice areas, the global model effectively reduced discrepancies between our data and the reference maps, achieving an overall classifier accuracy exceeding 80%. Despite the need for further refinement to address confusion with other crops, STM exhibits promise for national-scale rice paddy mapping in Brazil. Full article
(This article belongs to the Special Issue Irrigation Mapping Using Satellite Remote Sensing II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study region and validation sampling distribution. Two tiles (training and validation) were selected for each National Pole of Irrigated Agriculture. Black points represent non-rice samples, while green points represent rice samples.</p>
Full article ">Figure 2
<p>Rice spectral–temporal profile for each study region in Tocantins (<b>a</b>), Santa Catarina (<b>b</b>), and Rio Grande do Sul (<b>c</b>). The spectral indices are NDVI (green), NDMI (orange), and NDWI (blue).</p>
Full article ">Figure 3
<p>An example of spectral–temporal metrics (STM) for Rio Grande do Sul (RS). The red line is the contour of the rice mapping by ANA-CONAB (2019/2020). The polar metrics Q1 (<b>a</b>), Q2 (<b>b</b>), Q3 (<b>c</b>), Q4 (<b>d</b>), and Eccentricity (<b>e</b>) provide less contrast. On the other hand, Gyration Radius (<b>f</b>) and the basic metrics, Max (<b>g</b>), Min (<b>h</b>), Mean (<b>i</b>), AMD (<b>j</b>), Standard Deviation (<b>k</b>), First Quartile (<b>l</b>), Second Quartile (<b>m</b>), Third Quartile (<b>n</b>), and Interquartile Range (<b>o</b>) provide more contrast.</p>
Full article ">Figure 4
<p>Example of areas of training classified by the STM-based approach for the study regions compared to the official mapping for irrigated rice in Brazil. In black, the areas that are not rice are highlighted; in blue, areas of the ANA-CONAB mapping that were not classified by our approach are highlighted; in orange, areas classified as irrigated rice and not included in the official mapping are highlighted; in dark green, irrigated rice areas mapped by ANA-CONAB and our approach are highlighted. For Tocantins, the STM basic (<b>a</b>), STM Polar (<b>b</b>), STM Basic+Polar (<b>c</b>) and STM Global (<b>d</b>) classifications are shown. Similarly, Santa Catarina has STM basic (<b>e</b>), STM Polar (<b>f</b>), STM Basic+Polar (<b>g</b>) and STM Global (<b>h</b>) classifications. Finally, Rio Grande do Sul has classifications by STM basic (<b>i</b>), STM Polar (<b>j</b>), STM Basic+Polar (<b>k</b>) and STM Global (<b>l</b>) classifications.</p>
Full article ">Figure 5
<p>Estimated rice-planted area for training region in Santa Catarina (<b>a</b>), Rio Grande do Sul (<b>b</b>) and Tocantins (<b>c</b>), and comparison between regional and global classifications and ANA-CONAB mapping.</p>
Full article ">Figure 6
<p>Average confusion matrices for STM basic+polar classification in (<b>a</b>) Tocantins, (<b>b</b>) Santa Catarina, (<b>c</b>) Rio Grande do Sul, and (<b>d</b>) for global classification.</p>
Full article ">Figure 7
<p>The most frequent important variables over the 100 iterations of the Monte Carlo simulation for the classification of basic+polar STMs classified by MeanDecreaseGini (<b>a</b>) and MeanDecreaseAccuracy (<b>b</b>).</p>
Full article ">Figure 8
<p>Example of areas of validation classified by the STM-based approach for the study regions compared to the official mapping for irrigated rice in Brazil. In black, areas that are not rice are represented; in blue, areas of the ANA-CONAB mapping that were not classified by our approach are represented; in orange, areas classified as irrigated rice and not included in the official mapping are represented; in dark green, irrigated rice areas mapped by ANA-CONAB and our approach are represented. For Tocantins, the STM basic (<b>a</b>), STM Polar (<b>b</b>), STM Basic+Polar (<b>c</b>) and STM Global (<b>d</b>) classifications are shown. Similarly, Santa Catarina has STM basic (<b>e</b>), STM Polar (<b>f</b>), STM Basic+Polar (<b>g</b>) and STM Global (<b>h</b>) classifications. Finally, Rio Grande do Sul has classifications by STM basic (<b>i</b>), STM Polar (<b>j</b>), STM Basic+Polar (<b>k</b>) and STM Global (<b>l</b>) classifications.</p>
Full article ">Figure 9
<p>Estimated rice-planted area for validation region in Santa Catarina (<b>a</b>), Rio Grande do Sul (<b>b</b>) and Tocantins (<b>c</b>), and comparison between regional and global classifications and ANA-CONAB mapping.</p>
Full article ">Figure A1
<p>Validation pixel protocol for rice classes in Tocantins (validation samples #56 and #108). NDVI is light green, NDMI is dark green and NDWI is dark blue.</p>
Full article ">Figure A2
<p>Validation pixel protocol for rice classes in Santa Catarina (validation samples #86 and #129). NDVI is light green, NDMI is dark green and NDWI is dark blue.</p>
Full article ">Figure A3
<p>Validation pixel protocol for rice classes in Rio Grande do Sul (validation samples #56 and #120). NDVI is light green, NDMI is dark green and NDWI is dark blue.</p>
Full article ">
25 pages, 7111 KiB  
Article
Spatial–Temporal Changes in the Distribution of Populus euphratica Oliv. Forests in the Tarim Basin and Analysis of Influencing Factors from 1990 to 2020
by Xuefei Guo, Lijun Zhu, Zhikun Yang, Chaobin Yang and Zhijun Li
Forests 2024, 15(8), 1384; https://doi.org/10.3390/f15081384 - 7 Aug 2024
Viewed by 278
Abstract
Understanding the spatiotemporal evolution patterns of Populus euphratica Oliv. (P. euphratica) forests in the Tarim Basin (TB) and their influencing factors is crucial for regional ecological security and high-quality development. However, there is currently a lack of large-area, long-term systematic monitoring. [...] Read more.
Understanding the spatiotemporal evolution patterns of Populus euphratica Oliv. (P. euphratica) forests in the Tarim Basin (TB) and their influencing factors is crucial for regional ecological security and high-quality development. However, there is currently a lack of large-area, long-term systematic monitoring. This study utilized multi-source medium and high-resolution remote sensing images from the Landsat series and Sentinel-2, applying a Random Forest classification model to obtain distribution data of P. euphratica forests and shrublands in 14 areas of the TB from 1990 to 2020. We analyzed the effects of river distance, water transfer, and farmland on their distribution. Results indicated that both P. euphratica forests and shrublands decreased during the first 20 years and increased during the last 10 years. Within 1.5 km of river water transfer zones, P. euphratica forests more frequently converted to shrublands, while both forests and shrublands showed recovery in low-frequency water transfer areas. Farmland encroachment was most significant beyond 3 km from rivers. To effectively protect P. euphratica forests, we recommend intermittent low-frequency water transfers within 3 km of rivers and stricter management of agricultural expansion beyond 3 km. These measures will help maintain a balanced ecosystem and promote the long-term sustainability of P. euphratica forests. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Spatial distribution map of land use types in the TB (Approved map No. GS (2019)1822).</p>
Full article ">Figure 2
<p>The number of images from different sensors during 1990–2020.</p>
Full article ">Figure 3
<p>Schematic representation of the overall methodological workflow.</p>
Full article ">Figure 4
<p>Extraction and classification of <span class="html-italic">P. euphratica</span> forest boundary (1 is Landsat image (red is vegetation); 2 is Sentinel-2 image (red is vegetation); 3 is DJI P4 multi-spectral UAV aerial image (green is vegetation); 4 is the result of random forest classification; 5 for field photos). (A. Up-Tarim River; B. Mid-Tarim River; C. Down-Tarim River; E. Down-Yarkand River; F. Mid-Hotan River; H. Kongque River; I. Kashgar River; J. Cherchen River).</p>
Full article ">Figure 5
<p>Schematic diagram of <span class="html-italic">P. euphratica</span> forest and shrubland area change in the TB (<b>a</b>). 1990; (<b>b</b>). 1995; (<b>c</b>). 2000; (<b>d</b>). 2005; (<b>e</b>). 2010; (<b>f</b>). 2015, (<b>g</b>) 2020 (<b>h</b>) 1990–2020.</p>
Full article ">Figure 6
<p>Distribution of <span class="html-italic">P. euphratica</span> forest (<b>a</b>) and shrublands (<b>b</b>) in different rivers of the TB at different distances from rivers. (A. Up-Tarim River; B. Mid-Tarim River; C. Down-Tarim River; D. Up-Yarkand River; E. Down-Yarkand River; F. Mid-Hotan River; G. Down-Hotan River; H. Kongque River; I. Kashgar River; J. Cherchen River; K. Rivers in the northern Kunlun Mountains; L. Keriya River; N. Sangzhu River).</p>
Full article ">Figure 7
<p>(<b>a</b>) Analysis of the intensity of conversion from <span class="html-italic">P. euphratica</span> forest; (<b>b</b>) analysis of the intensity of conversion from shrublands to <span class="html-italic">P. euphratica</span> forest; (<b>c</b>) distribution of <span class="html-italic">P. euphratica</span> forest and shrublands conversion at different river distances; (<b>d</b>) distribution of <span class="html-italic">P. euphratica</span> forest and shrublands conversion at different rivers in the TB from 1990 to 2020.</p>
Full article ">Figure 8
<p>(<b>a</b>) Mutual conversion of <span class="html-italic">P. euphratica</span> forest, shrublands, and farmland Sankey map; (<b>b</b>) mutual conversion of <span class="html-italic">P. euphratica</span> forest, shrublands, and farmland spatial distribution map; (<b>c</b>) mutual conversion of <span class="html-italic">P. euphratica</span> forest, shrublands, and farmland in different rivers; (<b>d</b>) farmland-to-shrubland conversion in different rivers; (<b>e</b>) mutual conversion of <span class="html-italic">P. euphratica</span> forest, shrublands, and farmland in different distance from rivers; (<b>f</b>) mutual conversion of <span class="html-italic">P. euphratica</span> forest, shrublands, and farmland in different distance from rivers; (<b>f</b>) conversion of farmland to shrubland at different distances from rivers in the TB from 1990 to 2020.</p>
Full article ">Figure 9
<p>(<b>a</b>) Spatial distribution of <span class="html-italic">P. euphratica</span> forest and shrublands inter-conversion under different water transfer frequencies; (<b>b</b>) different frequencies of water transfers on different rivers; (<b>c</b>) <span class="html-italic">P. euphratica</span> and shrublands inter-conversion under different water delivery frequencies of different rivers; (<b>d</b>) <span class="html-italic">P. euphratica</span> and shrublands inter-conversion under different water delivery frequencies at different distances from the rivers in the TB from 1990 to 2020.</p>
Full article ">Figure 10
<p>(<b>a</b>) Spatial distribution map of farmland change and <span class="html-italic">P. euphratica</span> forest and shrubland transformation in the TB from 1990 to 2020; (<b>b</b>) Pearson correlation heatmap of farmland, <span class="html-italic">P. euphratica</span> forest, and shrubland transformation.”→”Indicates the direction of transformation.</p>
Full article ">
34 pages, 4458 KiB  
Article
Evaluating Tree Species Mapping: Probability Sampling Validation of Pure and Mixed Species Classes Using Convolutional Neural Networks and Sentinel-2 Time Series
by Tobias Schadauer, Susanne Karel, Markus Loew, Ursula Knieling, Kevin Kopecky, Christoph Bauerhansl, Ambros Berger, Stephan Graeber and Lukas Winiwarter
Remote Sens. 2024, 16(16), 2887; https://doi.org/10.3390/rs16162887 - 7 Aug 2024
Viewed by 429
Abstract
The accurate large-scale classification of tree species is crucial for the monitoring, protection, and management of the Earth’s invaluable forest ecosystems. Numerous previous studies have recognized the suitability of satellite imagery, particularly Sentinel-2 imagery, for this task. In this study, we utilized a [...] Read more.
The accurate large-scale classification of tree species is crucial for the monitoring, protection, and management of the Earth’s invaluable forest ecosystems. Numerous previous studies have recognized the suitability of satellite imagery, particularly Sentinel-2 imagery, for this task. In this study, we utilized a dense phenology Sentinel-2 time series, which offered consistent data across multiple granules, to map tree species across the entire forested area in Austria. Aiming for the classification scheme to more accurately represent actual forest conditions, we included mixed tree species and sparsely populated classes (classes with sparse canopy cover) alongside pure tree species classes. To enhance the training data for the mixed and sparse classes, synthetic data creation was employed. Autocorrelation has significant implications for the validation of thematic maps. To investigate the impact of spatial dependency on validation data, two methods were employed at numerous split and buffer distances: spatial split validation and a validation method based on a buffered ground reference probability samples provided by the National Forest inventory (NFI). While a random training data holdout set yielded 99% accuracy, the spatial split validation resulted in 74% accuracy, emphasizing the importance of accounting for spatial autocorrelation when validating with holdout sets derived from polygon-based training data. The validation based on NFI data resulted in 55% overall accuracy, 91% post-hoc pure class accuracy, and 79% accuracy when confusions in phenological proximity were disregarded (e.g., spruce–larch confused with spruce). The significant differences in accuracy observed between spatial split and NFI validation underscore the challenge for polygon-based training data to capture ground reference forest complexity, particularly in areas with diverse forests. This hardship is further accentuated by the pure class accuracy of 91%, revealing the substantial impact of mixed stands on the accuracy of tree species maps. Full article
Show Figures

Figure 1

Figure 1
<p>Tree species mapping workflow.</p>
Full article ">Figure 2
<p>MLP-ResNet hybrid schematic architecture.</p>
Full article ">Figure 3
<p>Analysis of accuracy measures over spatial split distances.</p>
Full article ">Figure 4
<p>National Forest inventory data validation (NFI-VAL) buffer distance analysis.</p>
Full article ">Figure 5
<p>National Forest inventory data validation (NFI-VAL) buffer distance analysis with training data discard kept constant up to 15,000 m.</p>
Full article ">Figure 6
<p>Tree species map over the entire forest in the study area.</p>
Full article ">Figure 7
<p>Tree species map on an intermediate zoom level.</p>
Full article ">Figure 8
<p>Tree species map intermediate–close zoom level: predominantly mountain pine, white pine, spruce, and larch.</p>
Full article ">Figure 9
<p>Tree species map intermediate–close zoom level: predominantly black pine, black pine–other deciduous, and beech.</p>
Full article ">Figure 10
<p>Tree species map intermediate–close zoom level: predominantly spruce and beech.</p>
Full article ">Figure 11
<p>Tree species map intermediate–close zoom level: typical mixed forest.</p>
Full article ">
21 pages, 11248 KiB  
Article
Transferability of Empirical Models Derived from Satellite Imagery for Live Fuel Moisture Content Estimation and Fire Risk Prediction
by Eva Marino, Lucía Yáñez, Mercedes Guijarro, Javier Madrigal, Francisco Senra, Sergio Rodríguez and José Luis Tomé
Fire 2024, 7(8), 276; https://doi.org/10.3390/fire7080276 - 6 Aug 2024
Viewed by 355
Abstract
Estimating live fuel moisture content (LFMC) is critical for assessing vegetation flammability and predicting potential fire behaviour, thus providing relevant information for wildfire prevention and management. Previous research has demonstrated that empirical modelling based on spectral data derived from remote sensing is useful [...] Read more.
Estimating live fuel moisture content (LFMC) is critical for assessing vegetation flammability and predicting potential fire behaviour, thus providing relevant information for wildfire prevention and management. Previous research has demonstrated that empirical modelling based on spectral data derived from remote sensing is useful for retrieving LFMC. However, these types of models are often very site-specific and generally considered difficult to extrapolate. In the present study, we analysed the performance of empirical models based on Sentinel-2 spectral data for estimating LFMC in fire-prone shrubland dominated by Cistus ladanifer. We used LFMC data collected in the field between June 2021 and September 2022 in 27 plots in the region of Andalusia (southern Spain). The specific objectives of the study included (i) to test previous existing models fitted for the same shrubland species in a different study area in the region of Madrid (central Spain); (ii) to calibrate empirical models with the field data from the region of Andalusia, comparing the model performance with that of existing models; and (iii) to test the capacity of the best empirical models to predict decreases in LFMC to critical threshold values in historical wildfire events. The results showed that the empirical models derived from Sentinel-2 data provided accurate LFMC monitoring, with a mean absolute error (MAE) of 15% in the estimation of LFMC variability throughout the year and with the MAE decreasing to 10% for the critical lower LFMC values (<100%). They also showed that previous models could be easily recalibrated for extrapolation to different geographical areas, yielding similar errors to the specific empirical models fitted in the study area in an independent validation. Finally, the results showed that decreases in LFMC in historical wildfire events were accurately predicted by the empirical models, with LFMC <80% in this fire-prone shrubland species. Full article
Show Figures

Figure 1

Figure 1
<p>Study area (black rectangle) within the region of Andalusia (red) in Spain (green), with details on the location of the 27 sampling plots (INIA + INFOCA) and the historical wildfires (burnt areas for each year in different colours) in the pilot areas (Z1, Z2 and Z3).</p>
Full article ">Figure 2
<p>Observed LFMC values in INIA (green) and INFOCA (orange) field sampling plots during the study period (June 2021 to September 2022).</p>
Full article ">Figure 3
<p>Pearson correlation coefficients for spectral indices and LFMC in the calibration dataset (values highlighted in dark blue denote stronger correlations, r &gt; 0.90).</p>
Full article ">Figure 4
<p>Extrapolation of the best empirical models previously fitted for the region of Madrid to the sampling plots in Andalusia, NLR-exp (<b>left</b>) and NLR-sqr (<b>right</b>), depicting validation results of original models (<b>upper</b>) and after recalibration (<b>lower</b>) with the linear regression (Y = a + bX) between observed and predicted values: NLR-exp (a = 17.33, b = 0.712); NLR-sqr (a = 7.659, b = 0.832). Y = recalibrated LFMC; X = predicted LFMC.</p>
Full article ">Figure 5
<p>Observed vs. predicted LFMC values in the best models for each formulation tested with the calibration dataset (n = 224) for field plots in Andalusia.</p>
Full article ">Figure 5 Cont.
<p>Observed vs. predicted LFMC values in the best models for each formulation tested with the calibration dataset (n = 224) for field plots in Andalusia.</p>
Full article ">Figure 6
<p>Observed vs. predicted LFMC values in the independent validation (n = 111) of the best models fitted for field plots in Andalusia.</p>
Full article ">Figure 6 Cont.
<p>Observed vs. predicted LFMC values in the independent validation (n = 111) of the best models fitted for field plots in Andalusia.</p>
Full article ">Figure 7
<p>Performance of the best empirical model for LFMC estimation before the 7 selected historical wildfire events in the pilot areas (Z1, Z2 and Z3), depicting the mean value of reference plots available for each wildfire (blue) and the LFMC range for each date (vertical bars). Headings indicate the wildfire name, with the ignition date in brackets and depicted as red symbols.</p>
Full article ">Figure A1
<p>Changes in LFMC estimation obtained with the best empirical model derived from Sentinel-2 data in the Z2 pilot area before the Almonaster wildfire (ignited on 27 August 2020), which burned an area of 14,957 ha in 12 days (perimeter in black). This wildfire was the biggest event in the region of Andalusia during the study period (2018–2022). LFMC is only shown for pixels corresponding to shrubland. Reference plots used in <a href="#fire-07-00276-f007" class="html-fig">Figure 7</a> for LFMC value calculations are depicted as green triangles.</p>
Full article ">Figure A1 Cont.
<p>Changes in LFMC estimation obtained with the best empirical model derived from Sentinel-2 data in the Z2 pilot area before the Almonaster wildfire (ignited on 27 August 2020), which burned an area of 14,957 ha in 12 days (perimeter in black). This wildfire was the biggest event in the region of Andalusia during the study period (2018–2022). LFMC is only shown for pixels corresponding to shrubland. Reference plots used in <a href="#fire-07-00276-f007" class="html-fig">Figure 7</a> for LFMC value calculations are depicted as green triangles.</p>
Full article ">
20 pages, 38319 KiB  
Article
Calculating Vegetation Index-Based Crop Coefficients for Alfalfa in the Mesilla Valley, New Mexico Using Harmonized Landsat Sentinel-2 (HLS) Data and Eddy Covariance Flux Tower Data
by Robert Sabie, A. Salim Bawazir, Michaela Buenemann, Caitriana Steele and Alexander Fernald
Remote Sens. 2024, 16(16), 2876; https://doi.org/10.3390/rs16162876 - 6 Aug 2024
Viewed by 742
Abstract
The goal of this study is to investigate the usefulness of the relatively new 30 m spatial and <5.7-day temporal resolution Harmonized Landsat Sentinel-2 (HLS) dataset for calculating vegetation index-based crop coefficients (KcVI) for estimating field scale crop evapotranspiration (ETc [...] Read more.
The goal of this study is to investigate the usefulness of the relatively new 30 m spatial and <5.7-day temporal resolution Harmonized Landsat Sentinel-2 (HLS) dataset for calculating vegetation index-based crop coefficients (KcVI) for estimating field scale crop evapotranspiration (ETc). Increased spatial and temporal resolution ETc estimates are needed for improving irrigation scheduling, monitoring impacts of water conservation programs, and improving crop yield. The crop coefficient (Kc) method is widely used for estimating ETc. Remote sensing vegetation indices (VI) are highly correlated to Kc and allow the creation of a KcVI but the approach is limited by the availability of high temporal and spatial resolutions. We selected and calculated sixteen commonly used VIs using HLS data and regressed them against field-measured ET for alfalfa in the Mesilla Valley, New Mexico to create linear KcVI models. All models showed good agreement with Kc (r2 > 0.67 and RMSE < 0.15). ETc prediction resulted in an MAE ranging between 0.35- and 0.64-mm day−1, an MSE ranging between 0.20- and 0.75-mm day−1 and an MAPD ranging between 10.0 and 16.5%. The largest differences in predicted ETc occurred early in the growing season and during cutting periods when the spectral signal could be influenced by soil background or irrigation events. The results suggest that applying the KcVI approach to the HLS dataset can help fill in the data gap in remote sensing ET tools. Future work should focus on assessing additional crops and integration into other tools such as the emerging OpenET platform. Full article
(This article belongs to the Section Remote Sensing in Agriculture and Vegetation)
Show Figures

Figure 1

Figure 1
<p>The location of the alfalfa study site (purple border) in the Mesilla Valley (red circle), New Mexico, USA.</p>
Full article ">Figure 2
<p>Processing steps used for creating and analyzing vegetation index-based crop coefficients.</p>
Full article ">Figure 3
<p>(<b>a</b>) Distribution of measured Kc values for the 2017 growing season, and (<b>b</b>) satellite acquisitions plotted on the daily measured Kc values with cuts indicated by the vertical dashed lines.</p>
Full article ">Figure 4
<p>Linear regressions between 16 vegetation indices and K<sub>c</sub> from in situ data. The red line represents the best-fit line.</p>
Full article ">Figure 5
<p>Graphs showing higher and lower prediction of ET<sub>c</sub> using the K<sub>cVI</sub> method compared to measured ET<sub>a</sub>. The red line is a visualization of zero deviation.</p>
Full article ">Figure 6
<p>Comparison of predicted crop ET (ET<sub>c</sub>) using the K<sub>CVI</sub> method and measured ET<sub>a</sub> from an eddy covariance tower for alfalfa during the 2017 growing season.</p>
Full article ">Figure 7
<p>Example of K<sub>CVI</sub> output map showing estimated ET (mm day<sup>−1</sup>) using NGRDI for 19 May 2017.</p>
Full article ">
Back to TopTop