[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,238)

Search Parameters:
Keywords = vegetation series

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
24 pages, 5994 KiB  
Article
Mapping Natural Populus euphratica Forests in the Mainstream of the Tarim River Using Spaceborne Imagery and Google Earth Engine
by Jiawei Zou, Hao Li, Chao Ding, Suhong Liu and Qingdong Shi
Remote Sens. 2024, 16(18), 3429; https://doi.org/10.3390/rs16183429 - 15 Sep 2024
Viewed by 244
Abstract
Populus euphratica is a unique constructive tree species within riparian desert areas that is essential for maintaining oasis ecosystem stability. The Tarim River Basin contains the most densely distributed population of P. euphratica forests in the world, and obtaining accurate distribution data in [...] Read more.
Populus euphratica is a unique constructive tree species within riparian desert areas that is essential for maintaining oasis ecosystem stability. The Tarim River Basin contains the most densely distributed population of P. euphratica forests in the world, and obtaining accurate distribution data in the mainstream of the Tarim River would provide important support for its protection and restoration. We propose a new method for automatically extracting P. euphratica using Sentinel-1 and 2 and Landsat-8 images based on the Google Earth Engine cloud platform and the random forest algorithm. A mask of the potential distribution area of P. euphratica was created based on prior knowledge to save computational resources. The NDVI (Normalized Difference Vegetation Index) time series was then reconstructed using the preferred filtering method to obtain phenological parameter features, and the random forest model was input by combining the phenological parameter, spectral index, textural, and backscattering features. An active learning method was employed to optimize the model and obtain the best model for extracting P. euphratica. Finally, the map of natural P. euphratica forests with a resolution of 10 m in the mainstream of the Tarim River was obtained. The overall accuracy, producer’s accuracy, user’s accuracy, kappa coefficient, and F1-score of the map were 0.96, 0.98, 0.95, 0.93, and 0.96, respectively. The comparison experiments showed that simultaneously adding backscattering and textural features improved the P. euphratica extraction accuracy, while textural features alone resulted in a poor extraction effect. The method developed in this study fully considered the prior and posteriori information and determined the feature set suitable for the P. euphratica identification task, which can be used to quickly obtain accurate large-area distribution data of P. euphratica. The method can also provide a reference for identifying other typical desert vegetation. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area and the distribution of sample points. (<b>a</b>): location of the study area in Xinjiang province in China; (<b>b</b>): training dataset distribution; (<b>c</b>): detailed sample area showing <span class="html-italic">P. euphratica</span> and non–<span class="html-italic">P. euphratica</span> in a Sentinel-2 false-color image.</p>
Full article ">Figure 2
<p>Distribution of validation dataset. The black solid line represents the range of the study area; the red and yellow points represent <span class="html-italic">P. euphratica</span> and non–<span class="html-italic">P. euphratica</span>, respectively.</p>
Full article ">Figure 3
<p>Workflow of the research.</p>
Full article ">Figure 4
<p>Threshold segmentation effect of MNDWI and NDVI. (<b>a</b>): false color image of Jieran Lik Reservoir in Xinjiang Province; (<b>b</b>): statistical result of the corresponding frequency distribution of MNDWI values of water and other ground objects in area (<b>a</b>); (<b>c</b>): false color image of Pazili Tamu in Xinjiang; (<b>d</b>): statistical result for the corresponding frequency distribution of NDVI values of desert bare land and other ground objects in region (<b>c</b>).</p>
Full article ">Figure 5
<p>Comparison of NDVI data before and after spatiotemporal fusion: (<b>a</b>) NDVI data derived from Sentinel-2 before fusion, (<b>b</b>) NDVI data after fusion.</p>
Full article ">Figure 6
<p>Comparison of the effects of different filter functions for: (<b>a</b>) <span class="html-italic">P. euphratica</span>; (<b>b</b>) <span class="html-italic">Tamarix</span>; (<b>c</b>) allee tree; (<b>d</b>) farmland; (<b>e</b>) wetland; (<b>f</b>) urban tree.</p>
Full article ">Figure 7
<p>Comparison between phenological curves of six typical vegetation species. Phenology parameters of (<b>a</b>) <span class="html-italic">P. euphratica</span>, (<b>b</b>) <span class="html-italic">Tamarix</span>, (<b>c</b>) allee tree, (<b>d</b>) farmland, (<b>e</b>) wetland, and (<b>f</b>) urban tree.</p>
Full article ">Figure 8
<p>Importance of different features in the RF classification.</p>
Full article ">Figure 9
<p>Natural <span class="html-italic">P. euphratica</span> forest maps extracted using four feature combinations: (<b>a</b>) PS, (<b>b</b>) PSB, (<b>c</b>) PST, and (<b>d</b>) PSBT.</p>
Full article ">Figure 10
<p>Comparison of <span class="html-italic">P. euphratica</span> extraction results using different feature combinations on Sentinel-2 standard false color images. Rows 1 to 4 show the identification of <span class="html-italic">P. euphratica</span> in desert areas, <span class="html-italic">P. euphratica</span>-dense areas, agricultural areas, and large river areas, respectively. The green area represents the classification result of <span class="html-italic">P. euphratica</span>. The yellow circle corresponding to each row is the area where the extraction results of different feature combinations are quite different.</p>
Full article ">Figure 11
<p>(<b>a</b>) Distribution of natural <span class="html-italic">P. euphratica</span> forest in the mainstream of the Tarim River. (<b>b</b>): UAV image of healthy <span class="html-italic">P. euphratica</span>, (<b>c</b>): classification result of healthy <span class="html-italic">P. euphratica</span>, (<b>d</b>): UAV image of unhealthy <span class="html-italic">P. euphratica</span>, (<b>e</b>): classification result of unhealthy <span class="html-italic">P. euphratica</span>, (<b>f</b>): UAV image of dense <span class="html-italic">P. euphratica</span>, (<b>g</b>): classification result of dense <span class="html-italic">P. euphratica</span>, (<b>h</b>): UAV image of sparse <span class="html-italic">P. euphratica</span>, (<b>i</b>): classification result of sparse <span class="html-italic">P. euphratica</span>. The green area represents the classification results of <span class="html-italic">P. euphratica</span>.</p>
Full article ">Figure 12
<p>Mixed pixel problems associated with <span class="html-italic">P. euphratica</span>: (<b>a</b>) <span class="html-italic">P. euphratica</span> occupying less than one pixel; (<b>b</b>) sandy soil interfering with the reflected signal of <span class="html-italic">P. euphratica</span>. The red box represents a pixel on the images for clearer observation. Basemaps of row 1-2 are UAV images while row 3 are Sentinel-2 standard false color images.</p>
Full article ">
18 pages, 4238 KiB  
Article
Combining Vegetation Indices to Identify the Maize Phenological Information Based on the Shape Model
by Huizhu Wu, Bing Liu, Bingxue Zhu, Zhijun Zhen, Kaishan Song and Jingquan Ren
Agriculture 2024, 14(9), 1608; https://doi.org/10.3390/agriculture14091608 - 14 Sep 2024
Viewed by 210
Abstract
Maize is the world’s largest food crop and plays a critical role in global food security. Accurate phenology information is essential for improving yield estimation and enabling timely field management. Yet, much of the research has concentrated on general crop growth periods rather [...] Read more.
Maize is the world’s largest food crop and plays a critical role in global food security. Accurate phenology information is essential for improving yield estimation and enabling timely field management. Yet, much of the research has concentrated on general crop growth periods rather than on pinpointing key phenological stages. This gap in understanding presents a challenge in determining how different vegetation indices (VIs) might accurately extract phenological information across these stages. To address this, we employed the shape model fitting (SMF) method to assess whether a multi-index approach could enhance the precision of identifying key phenological stages. By analyzing time-series data from various VIs, we identified five phenological stages (emergence, seven-leaf, jointing, flowering, and maturity stages) in maize cultivated in Jilin Province. The findings revealed that each VI had distinct advantages depending on the phenological stage, with the land surface water index (LSWI) being particularly effective for jointing and flowering stages due to its correlation with vegetation water content, achieving a root mean square error (RMSE) of three to four days. In contrast, the normalized difference vegetation index (NDVI) was more effective for identifying the emergence and seven-leaf stages, with an RMSE of four days. Overall, combining multiple VIs significantly improved the accuracy of phenological stage identification. This approach offers a novel perspective for utilizing diverse VIs in crop phenology, thereby enhancing the precision of agricultural monitoring and management practices. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the study area and sites.</p>
Full article ">Figure 2
<p>Division of vegetative and reproductive growth stages in maize phenology.</p>
Full article ">Figure 3
<p>Flowchart of phenological period identification.</p>
Full article ">Figure 4
<p>Relationship between field-observed and estimated phenological stages obtained from different indicators and the RMSE. (<b>a</b>,<b>c</b>,<b>e</b>) The data from 2003 to 2014 and (<b>b</b>,<b>d</b>,<b>f</b>) the data from 2015 to 2019. Both the x and y axes represent the DOY.</p>
Full article ">Figure 5
<p>Box plots of the errors in identifying phenological phases in different regions: (<b>a</b>) emergence, (<b>b</b>) seven-leaf stage, (<b>c</b>) jointing, (<b>d</b>) flowering, and (<b>e</b>) maturity. The errors are represented by residuals, with positive values indicating delayed predictions and negative values indicating earlier predictions.</p>
Full article ">Figure 6
<p>The target curve fitting reference curve in Baicheng City, Jilin Province, in 2019, where the orange curve is the reference curve, the blue curve is the target curve, and the green curve is the reference curve after deformation by the fitting function. In the subplots, (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>,<b>m</b>) represent the fitting cases where NDVI is used as the reference curve for the shape model, (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>,<b>n</b>) represent the cases with NDPI as the reference curve, and (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>,<b>o</b>) represent the cases with LSWI as the reference curve, with each subplot labeled accordingly.</p>
Full article ">Figure 7
<p>Spatial distribution of the five phenological stages of maize obtained from the LSWI as a reference curve.</p>
Full article ">Figure 8
<p>Reference curves of different VIs in Jilin City.</p>
Full article ">Figure 9
<p>Reflectance values and reference curves at different phenological stages of different stations. (<b>a</b>) represents the case where the phenological stages of the stations correspond to NDVI values falling on the reference curve, (<b>b</b>) represents the case for NDPI values, and (<b>c</b>) represents the case for LSWI values.</p>
Full article ">
15 pages, 4826 KiB  
Article
Assessing Evapotranspiration Changes in Response to Cropland Expansion in Tropical Climates
by Leonardo Laipelt, Julia Brusso Rossi, Bruno Comini de Andrade, Morris Scherer-Warren and Anderson Ruhoff
Remote Sens. 2024, 16(18), 3404; https://doi.org/10.3390/rs16183404 - 13 Sep 2024
Viewed by 240
Abstract
The expansion of cropland in tropical regions has significantly accelerated in recent decades, triggering an escalation in water demand and changing the total water loss to the atmosphere (evapotranspiration). Additionally, the increase in areas dedicated to agriculture in tropical climates coincides with an [...] Read more.
The expansion of cropland in tropical regions has significantly accelerated in recent decades, triggering an escalation in water demand and changing the total water loss to the atmosphere (evapotranspiration). Additionally, the increase in areas dedicated to agriculture in tropical climates coincides with an increased frequency of drought events, leading to a series of conflicts among water users. However, detailed studies on the impacts of changes in water use due to agriculture expansion, including irrigation, are still lacking. Furthermore, the higher presence of clouds in tropical environments poses challenges for the availability of high-resolution data for vegetation monitoring via satellite images. This study aims to analyze 37 years of agricultural expansion using the Landsat collection and a satellite-based model (geeSEBAL) to assess changes in evapotranspiration resulting from cropland expansion in tropical climates, focusing on the São Marcos River Basin in Brazil. It also used a methodology for estimating daily evapotranspiration on days without satellite images. The results showed a 34% increase in evapotranspiration from rainfed areas, mainly driven by soybean cultivation. In addition, irrigated areas increased their water use, despite not significantly changing water use at the basin scale. Conversely, natural vegetation areas decreased their evapotranspiration rates by 22%, suggesting possible further implications with advancing changes in land use and land cover. Thus, this study underscores the importance of using satellite-based evapotranspiration estimates to enhance our understanding of water use across different land use types and scales, thereby improving water management strategies on a large scale. Full article
Show Figures

Figure 1

Figure 1
<p>São Marcos River Basin: location in Brazil (<b>a</b>), climate zones according to Köppen–Geiger classification and irrigation pivots (<b>b</b>), and Normalized Difference Vegetation Index (NDVI) values computed using average composition of Landsat 8 for 2021 (<b>c</b>).</p>
Full article ">Figure 2
<p>Daily average of <span class="html-italic">ET</span> illustrated as boxplot for each land cover and use (<b>a</b>) for Landsat scenes between 1986 and 2022. We also illustrated the seasonal monthly average of <span class="html-italic">ET</span> (<b>b</b>), and trends of annual average <span class="html-italic">ET</span> for different land types, with natural vegetation (forest and savanna) demonstrating positive trends over the years, as well as irrigated areas (<b>c</b>).</p>
Full article ">Figure 3
<p>Changes in the <span class="html-italic">ET</span> spatial patterns for the São Marcos River Basin from 1986 to 2022 (<b>a</b>). The contribution of the water usage for each land cover and use between 1986 and 2022 is shown in (<b>b</b>), whereas (<b>c</b>) illustrates changes in land cover and use.</p>
Full article ">Figure 4
<p>Annual composition ET (mm day<sup>−1</sup>) in the São Marcus River Basin between 1986 (<b>a</b>) and 2021 (<b>b</b>). Highlighted plots showed the expressive number of pivot irrigation systems over the basin for specific locations.</p>
Full article ">Figure 5
<p>Monthly <span class="html-italic">ET</span> in the São Marcus River basin was analyzed for each month of one water year (2019 and 2020). During the dry season (May to September), precipitation is limited and radiation availability is high, being a water-limited environment. Consequently, lower <span class="html-italic">ET</span> values are observed during the dry season, while the wet season increases <span class="html-italic">ET</span> rates due to higher precipitation availability.</p>
Full article ">Figure 6
<p>Seasonal differences in daily <span class="html-italic">ET</span> for irrigated and rainfed croplands in the São Marcus River Basin (<b>a</b>), and the difference between both estimations (<b>b</b>). We used a simplified method to fill the gap between Landsat scenes by interpolating <span class="html-italic">EF</span> over time and multiplying with the respective reference <span class="html-italic">ET</span>.</p>
Full article ">
13 pages, 4217 KiB  
Article
Effect of Fatty Acids on Vegetable-Oil-Derived Sustainable Polyurethane Coatings for Controlled-Release Fertilizer
by Minhui Pang, Zirui Liu, Hongyan Li, Lina Liang and Lixia Li
Coatings 2024, 14(9), 1183; https://doi.org/10.3390/coatings14091183 - 12 Sep 2024
Viewed by 410
Abstract
Vegetable-oil-based polyurethane has become a promising sustainable candidate for controlled-release fertilizer based on green chemistry. The purpose of this study was to prepare a series of coatings from selective feedstocks including five vegetable oils with a high saturation degree, mono-unsaturation degree, or poly-unsaturation [...] Read more.
Vegetable-oil-based polyurethane has become a promising sustainable candidate for controlled-release fertilizer based on green chemistry. The purpose of this study was to prepare a series of coatings from selective feedstocks including five vegetable oils with a high saturation degree, mono-unsaturation degree, or poly-unsaturation degree, considering that vegetable oil fatty acids played a key role in the synthesis of polyol and polyurethane. The effect of the type and proportion of fatty acids on the physicochemical properties, microstructure, and macro-properties of vegetable-oil-derived polyols and their resulting coatings was characterized and discussed. The position and number of the hydroxy groups were determined by the type and proportion of fatty acid, and polyol from linseed oil with a high poly-unsaturation degree and three carbon–carbon double bonds had a high hydroxyl value and functionality, whereas polyol from palm oil with a high saturation degree possessed the lowest hydroxyl value and functionality. The resultant coating from linseed-oil-based polyol had a good cross-linking density, and the nitrogen release longevity of coated urea was 56 days at a coating percentage of 3%, and its nitrogen use efficiency was increased by 27.15% compared with conventional urea. Although the palm-oil-based coating had good hydrophobicity, its coated urea was not ideal. Overall, this study has enriched theories of bio-based polyurethane coatings for controlled-release fertilizers; using vegetable oil with a poly-unsaturation degree, it is easy to obtain an excellent coating for controlled-release fertilizer, and this will help provide economic and environmental benefits. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Reaction diagram of the VOPs and their coatings.</p>
Full article ">Figure 2
<p>ATR-FTIR (<b>a</b>,<b>b</b>) and <sup>1</sup>H NMR (<b>c</b>,<b>d</b>) spectra of VOs and their polyols, respectively.</p>
Full article ">Figure 3
<p>ATR-FTIR (<b>a</b>) and XPS (<b>b</b>) spectra of coatings; XPS C1s (<b>c</b>–<b>e</b>) and O1s (<b>f</b>–<b>h</b>) fitting curves of PPU, RPU, and LPU, respectively.</p>
Full article ">Figure 4
<p>SEM and AFM images of PPU, RPU, and LPU coatings and films. SEM surface (<b>A1</b>–<b>A3</b>) and cross sections (<b>B1</b>–<b>B3</b>); AFM (<b>C1</b>,<b>D1</b>), (<b>C2</b>,<b>D2</b>), and (<b>C3</b>,<b>D3</b>), respectively.</p>
Full article ">Figure 5
<p>TG (<b>a</b>)/DTG (<b>b</b>) and WCAs (<b>c</b>) of coatings, and nitrogen release behaviors (<b>d</b>) of their coated urea.</p>
Full article ">Figure 6
<p>The results of the pot experiment. Picture of Chinese cabbage after growth for 45 days (<b>a</b>); effect of different treatments on fresh weight and dry weight (<b>b</b>), plant height (<b>c</b>), and nitrogen use efficiency (<b>d</b>) of Chinese cabbage. Various lowercase letters (a and b) demonstrate notable differences between the treatments (<span class="html-italic">p</span> &lt; 0.05).</p>
Full article ">
17 pages, 2704 KiB  
Article
Changes in Ground Cover Layers, Biomass and Diversity of Vascular Plants/Mosses in the Clear-Cuts Followed by Reforested Scots Pine until Maturity Age
by Dovilė Gustienė and Iveta Varnagirytė-Kabašinskienė
Land 2024, 13(9), 1477; https://doi.org/10.3390/land13091477 - 12 Sep 2024
Viewed by 246
Abstract
The distribution of Scots pine (Pinus sylvestris L.) forests, particularly the Vaccinio myrtillo-Pinetum type, is determined by edaphic conditions, and although clear-cutting is used to promote regeneration, it remains controversial. This study evaluated the changes in non-living (forest floor and dead wood) [...] Read more.
The distribution of Scots pine (Pinus sylvestris L.) forests, particularly the Vaccinio myrtillo-Pinetum type, is determined by edaphic conditions, and although clear-cutting is used to promote regeneration, it remains controversial. This study evaluated the changes in non-living (forest floor and dead wood) and living (mosses, herbs, and dwarf shrubs) ground cover in clear-cut areas and reforested Scots pine stands. Continuous ground cover studies were conducted in clear-cuts, with samples collected over three years after clear-cutting, while data from 8–80-year-old and mature Scots pine stands were collected using the chronological series method with a consistent methodology in temporary plots. The research has shown that, as ecosystem recovery progresses, similarity to the mature forest increases, and a threshold stand age has been identified, beyond which the ecological changes induced by clear-cutting diminish. The study findings demonstrated that clear-cutting in Pinetum vaccinio-myrtillosum-type forest stands lead to a rapid increase in herb and dwarf shrub cover due to reduced competition for light and nutrients. However, clear-cutting caused a significant decline in forest-specific species and a drastic reduction in forest floor and dead wood mass, with a gradual recovery of moss cover over 10–30 years. These findings highlight the importance of managing clear-cutting practices to balance immediate vegetative responses with long-term ecosystem stability and biodiversity conservation. Full article
(This article belongs to the Special Issue Recent Progress in Land Degradation Processes and Control)
Show Figures

Figure 1

Figure 1
<p>Research scheme: three research objects (Trakai, Varėna, and Kazlų Rūda), each included study sites of selected 1–3-year-old clear-cuts and 8–10-, 15–20-, 30–40-, 70–80-, and 110–130-year-old Scots pine stands.</p>
Full article ">Figure 2
<p>The percentage distribution (%) of living ground cover (<b>A</b>) and vascular plant cover (<b>B</b>), each calculated from the total living ground cover (100%) and total vascular plant cover (100%), respectively, in the 1–3-year-old clear-cuts and 8–130-year-old Scots pine stands (aggregated data from three sites).</p>
Full article ">Figure 3
<p>Pair distances between the 1–3-year-old clear-cuts and 8–130-year-old Scots pine stands, illustrated for three sites: Trakai (<b>A</b>), Varėna (<b>B</b>), and Kazlų Rūda (<b>C</b>).</p>
Full article ">Figure 4
<p>The trend of living (<b>A</b>) and non-living (<b>B</b>) ground cover mass in Scots pine stands of <span class="html-italic">Pinetum vaccinio-myrtillosum</span> type throughout the rotation period following clear-cutting (aggregated data from three sites).</p>
Full article ">Figure 5
<p>Relations between the mean aboveground mass of vascular plants and mosses (kg ha<sup>−1</sup>) with the mean stand canopy density in the 1–3-year-old clear-cuts and 8–130-year-old Scots pine stands (aggregated data from three sites).</p>
Full article ">Figure 6
<p>The percentage of vascular plant (key species of herbs and dwarf shrubs) mass (<b>A</b>) and moss species mass (<b>B</b>) of the total vascular plant and moss mass, respectively, and aboveground biomass (kg ha<sup>−1</sup>) of the living ground cover (Total) in 1–3-year-old clear-cuts and 8–130-year-old Scots pine stands (aggregated data from three sites).</p>
Full article ">Figure 7
<p>Relations between the biomass of living soil cover elements (t ha<sup>−1</sup>) and the mean forest floor mass (t ha<sup>−1</sup>) in the 1–3-year-old clear-cuts and 8–130-year-old Scots pine stands (aggregated data from three sites).</p>
Full article ">
22 pages, 8177 KiB  
Article
ANN-Based Filtering of Drone LiDAR in Coastal Salt Marshes Using Spatial–Spectral Features
by Kunbo Liu, Shuai Liu, Kai Tan, Mingbo Yin and Pengjie Tao
Remote Sens. 2024, 16(18), 3373; https://doi.org/10.3390/rs16183373 - 11 Sep 2024
Viewed by 330
Abstract
Salt marshes provide diverse habitats for a wide range of creatures and play a key defensive and buffering role in resisting extreme marine hazards for coastal communities. Accurately obtaining the terrains of salt marshes is crucial for the comprehensive management and conservation of [...] Read more.
Salt marshes provide diverse habitats for a wide range of creatures and play a key defensive and buffering role in resisting extreme marine hazards for coastal communities. Accurately obtaining the terrains of salt marshes is crucial for the comprehensive management and conservation of coastal resources and ecology. However, dense vegetation coverage, periodic tide inundation, and pervasive ditch distribution create challenges for measuring or estimating salt marsh terrains. These environmental factors make most existing techniques and methods ineffective in terms of data acquisition resolution, accuracy, and efficiency. Drone multi-line light detection and ranging (LiDAR) has offered a fire-new perspective in the 3D point cloud data acquisition and potentially exhibited great superiority in accurately deriving salt marsh terrains. The prerequisite for terrain characterization from drone multi-line LiDAR data is point cloud filtering, which means that ground points must be discriminated from the non-ground points. Existing filtering methods typically rely on either LiDAR geometric or intensity features. These methods may not perform well in salt marshes with dense, diverse, and complex vegetation. This study proposes a new filtering method for drone multi-line LiDAR point clouds in salt marshes based on the artificial neural network (ANN) machine learning model. First, a series of spatial–spectral features at the individual (e.g., elevation, distance, and intensity) and neighborhood (e.g., eigenvalues, linearity, and sphericity) scales are derived from the original data. Then, the derived spatial–spectral features are selected to remove the related and redundant ones for optimizing the performance of the ANN model. Finally, the reserved features are integrated as input variables in the ANN model to characterize their nonlinear relationships with the point categories (ground or non-ground) at different perspectives. A case study of two typical salt marshes at the mouth of the Yangtze River, using a drone 6-line LiDAR, demonstrates the effectiveness and generalization of the proposed filtering method. The average G-mean and AUC achieved were 0.9441 and 0.9450, respectively, outperforming traditional geometric information-based methods and other advanced machine learning methods, as well as the deep learning model (RandLA-Net). Additionally, the integration of spatial–spectral features at individual–neighborhood scales results in better filtering outcomes than using either single-type or single-scale features. The proposed method offers an innovative strategy for drone LiDAR point cloud filtering and salt marsh terrain derivation under the novel solution of deeply integrating geometric and radiometric data. Full article
(This article belongs to the Section Ecological Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the two study sites, (<b>b</b>) original point clouds of Site 1 and positions of the training, test, and validation sets, (<b>c</b>) original point clouds of Site 2.</p>
Full article ">Figure 2
<p>Overall flowchart of the proposed method.</p>
Full article ">Figure 3
<p>Counts of ground and non-ground points in different original intensity values for sub-region 5.</p>
Full article ">Figure 4
<p>(<b>a</b>) Visualization of the correlation matrix for the preliminary selected features, (<b>b</b>) normalized feature importance score ranking for the preliminary selected features in LightGBM model.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>f</b>) Elevations of sub-regions 2, 3, 4, 6, 7, and 8, respectively. (<b>g</b>–<b>l</b>) filtering results for sub-regions 2, 3, 4, 6, 7, and 8, respectively.</p>
Full article ">Figure 6
<p>Point cloud filtering results of the ANN model for the validation sets. (<b>a</b>–<b>c</b>) Elevations of sub-regions 1, 5, and 9, respectively. (<b>d</b>–<b>f</b>) filtering results for sub-regions 1, 5, and 9, respectively.</p>
Full article ">Figure 7
<p>Filtering results by different methods of Site 1. (<b>a</b>–<b>h</b>) Filtering results of ANN, RF, XGBoost, LightGBM, SF, PMF, CSF, and RandL-Net, respectively. (<b>i</b>–<b>p</b>) ground points obtained by ANN, RF, XGBoost, LightGBM, SF, PMF, CSF, and RandLA-Net, respectively, where the ground/non-ground points after filtering are 37,234,012/28,364,019, 35,565,000/284,833,031, 32,750,140/287,647,891, 32,701,615/287696416, 129,737,212/190,660,819, 57,745,223/262,652,808, 210,752,452/109,645,579, and 112,776,548/207,621,483.</p>
Full article ">Figure 8
<p>Comparison of filtering results of different filtering methods on the validation set. (<b>a</b>–<b>c</b>) Manual, (<b>d</b>–<b>f</b>) RF, (<b>g</b>–<b>i</b>) XGBoost, (<b>j</b>–<b>l</b>) LightGBM, (<b>m</b>–<b>o</b>) SF, (<b>p</b>–<b>r</b>) PMF, (<b>s</b>–<b>u</b>) CSF, (<b>v</b>–<b>x</b>) RandLA-Net.</p>
Full article ">Figure 8 Cont.
<p>Comparison of filtering results of different filtering methods on the validation set. (<b>a</b>–<b>c</b>) Manual, (<b>d</b>–<b>f</b>) RF, (<b>g</b>–<b>i</b>) XGBoost, (<b>j</b>–<b>l</b>) LightGBM, (<b>m</b>–<b>o</b>) SF, (<b>p</b>–<b>r</b>) PMF, (<b>s</b>–<b>u</b>) CSF, (<b>v</b>–<b>x</b>) RandLA-Net.</p>
Full article ">Figure 9
<p>Filtering results on the validation set using ANN model trained with different features. (<b>a</b>–<b>c</b>) Point-wise features, (<b>d</b>–<b>f</b>) neighborhood-wise features.</p>
Full article ">Figure 10
<p>Filtering results of the entire study site at different scales. (<b>a</b>,<b>b</b>) Filtering results of ANN at individual and neighborhood scales. (<b>c</b>,<b>d</b>) ground points obtained by ANN at individual and neighborhood scales.</p>
Full article ">Figure 10 Cont.
<p>Filtering results of the entire study site at different scales. (<b>a</b>,<b>b</b>) Filtering results of ANN at individual and neighborhood scales. (<b>c</b>,<b>d</b>) ground points obtained by ANN at individual and neighborhood scales.</p>
Full article ">Figure 11
<p>(<b>a</b>) Filtering results of Site 2 using ANN, (<b>b</b>) ground points of Site 2 obtained by ANN, where the ground/non-ground points after filtering are 65,721,484/348,634,204.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) Elevations of sub-regions A, B, and C in Site 2, respectively. (<b>d</b>–<b>f</b>) filtering results of the ANN model for sub-regions A, B, and C in Site 2, respectively.</p>
Full article ">Figure 13
<p>Filtering results of the ANN model for the validation sets in Site 1 without feature selection. (<b>a</b>) Sub-region 1, (<b>b</b>) sub-region 5, (<b>c</b>) sub-region 9.</p>
Full article ">Figure A1
<p>Some profiles of the filtering result in Site 1 for the ANN model.</p>
Full article ">Figure A2
<p>Some profiles of the filtering result in Site 2 for the ANN model.</p>
Full article ">Figure A3
<p>Loss curves for training the RandLA-Net model.</p>
Full article ">
20 pages, 13462 KiB  
Article
Extraction of Garlic in the North China Plain Using Multi-Feature Combinations from Active and Passive Time Series Data
by Chuang Peng, Binglong Gao, Wei Wang, Wenji Zhu, Yongqi Chen and Chao Dong
Appl. Sci. 2024, 14(18), 8141; https://doi.org/10.3390/app14188141 - 10 Sep 2024
Viewed by 463
Abstract
Garlic constitutes a significant small-scale agricultural commodity in China. A key factor influencing garlic prices is the planted area, which can be accurately and efficiently determined using remote sensing technology. However, the spectral characteristics of garlic and winter wheat are easily confused, and [...] Read more.
Garlic constitutes a significant small-scale agricultural commodity in China. A key factor influencing garlic prices is the planted area, which can be accurately and efficiently determined using remote sensing technology. However, the spectral characteristics of garlic and winter wheat are easily confused, and the widespread intercropping of these crops in the study area exacerbates this issue, leading to significant challenges in remote sensing image analysis. Additionally, remote sensing data are often affected by weather conditions, spatial resolution, and revisit frequency, which can result in delayed and inaccurate area extraction. In this study, historical data were utilized to restore Sentinel-2 remote sensing images, aimed at mitigating cloud and rain interference. Feature combinations were devised, incorporating two vegetation indices into a comprehensive time series, along with Sentinel-1 synthetic aperture radar (SAR) time series and other temporal datasets. Multiple classification combinations were employed to extract garlic within the study area, and the accuracy of the classification results was systematically analyzed. First, we used passive satellite imagery to extract winter crops (garlic, winter wheat, and others) with high accuracy. Second, we identified garlic by applying various combinations of time series features derived from both active and passive remote sensing data. Third, we evaluated the classification outcomes of various feature combinations to generate an optimal garlic cultivation distribution map for each region. Fourth, we developed a garlic fragmentation index to assess the impact of landscape fragmentation on garlic extraction accuracy. The findings reveal that: (1) Better results in garlic extraction can be achieved using active–passive time series remote sensing. The performance of the classification model can be further enhanced by incorporating short-wave infrared bands or spliced time series data into the classification features. (2) Examination of garlic cultivation fragmentation using the garlic fragmentation index aids in elucidating variations in accuracy across the study area’s six counties. (3) Comparative analysis with validation samples demonstrated superior garlic extraction outcomes from the six primary garlic-producing counties of the North China Plain in 2021, achieving an overall precision exceeding 90%. This study offers a practical exploration of target crop identification using multi-source remote sensing data in mixed cropping areas. The methodology presented here demonstrates the potential for efficient, cost-effective, and accurate garlic classification, which is crucial for improving garlic production management and optimizing agricultural practices. Moreover, this approach holds promise for broader applications, such as nationwide garlic mapping. Full article
(This article belongs to the Special Issue Intelligent Computing and Remote Sensing—2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Diagram of garlic fertility period.</p>
Full article ">Figure 2
<p>Overview of the study area.</p>
Full article ">Figure 3
<p>Workflow of garlic extraction based on active–passive remote sensing time series data.</p>
Full article ">Figure 4
<p>The time series of Sentinel-2 NDVI for garlic and winter wheat. The curves in the figure represent the average NDVI values derived from the samples, while the upper and lower boundaries indicate the standard deviation.</p>
Full article ">Figure 5
<p>Time series of Sentinel-1 curves for garlic and winter wheat. (<b>a</b>) Time series data on the ratio of vertical–vertical (VV) and vertical–horizontal (VH) polarization in garlic and winter wheat. (<b>b</b>) Time series data of VV and VH polarization for garlic and winter wheat. The curves in the figure represent the average values derived from the samples.</p>
Full article ">Figure 6
<p>Winter crop distribution maps. This figure illustrates the winter vegetation classification results for each county within the study area. Green indicates areas of winter vegetation, while gray represents other land cover types.</p>
Full article ">Figure 6 Cont.
<p>Winter crop distribution maps. This figure illustrates the winter vegetation classification results for each county within the study area. Green indicates areas of winter vegetation, while gray represents other land cover types.</p>
Full article ">Figure 7
<p>Garlic distribution maps. This figure illustrates the garlic classification results for each county within the study area. Blue indicates areas of garlic, while gray represents other land cover types.</p>
Full article ">
21 pages, 10428 KiB  
Article
Three Decades of Inundation Dynamics in an Australian Dryland Wetland: An Eco-Hydrological Perspective
by Indishe P. Senanayake, In-Young Yeo and George A. Kuczera
Remote Sens. 2024, 16(17), 3310; https://doi.org/10.3390/rs16173310 - 6 Sep 2024
Viewed by 612
Abstract
Wetland ecosystems are experiencing rapid degradation due to human activities, particularly the diversion of natural flows for various purposes, leading to significant alterations in wetland hydrology and their ecological functions. However, understanding and quantifying these eco-hydrological changes, especially concerning inundation dynamics, presents a [...] Read more.
Wetland ecosystems are experiencing rapid degradation due to human activities, particularly the diversion of natural flows for various purposes, leading to significant alterations in wetland hydrology and their ecological functions. However, understanding and quantifying these eco-hydrological changes, especially concerning inundation dynamics, presents a formidable challenge due to the lack of long-term, observation-based spatiotemporal inundation information. In this study, we classified wetland areas into ten equal-interval classes based on inundation probability derived from a dense, 30-year time series of Landsat-based inundation maps over an Australian dryland riparian wetland, Macquarie Marshes. These maps were then compared with three simplified vegetation patches in the area: river red gum forest, river red gum woodland, and shrubland. Our findings reveal a higher inundation probability over a small area covered by river red gum forest, exhibiting persistent inundation over time. In contrast, river red gum woodland and shrubland areas show fluctuating inundation patterns. When comparing percentage inundation with the Normalized Difference Vegetation Index (NDVI), we observed a notable agreement in peaks, with a lag time in NDVI response. A strong correlation between NDVI and the percentage of inundated area was found in the river red gum woodland patch. During dry, wet, and intermediate years, the shrubland patch consistently demonstrated similar inundation probabilities, while river red gum patches exhibited variable probabilities. During drying events, the shrubland patch dried faster, likely due to higher evaporation rates driven by exposure to solar radiation. However, long-term inundation probability exhibited agreement with the SAGA wetness index, highlighting the influence of topography on inundation probability. These findings provide crucial insights into the complex interactions between hydrological processes and vegetation dynamics in wetland ecosystems, underscoring the need for comprehensive monitoring and management strategies to mitigate degradation and preserve these vital ecosystems. Full article
(This article belongs to the Special Issue Remote Sensing for Land Degradation and Drought Monitoring II)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>–<b>c</b>) location of the study area, Macquarie Marshes. (<b>d</b>) LiDAR-derived 1 m digital elevation model (DEM) of Macquarie Marshes. (<b>e</b>) Temporal average NDVI values over the Marshes as captured by the Landsat 8 collections from 2013 to 2020.</p>
Full article ">Figure 2
<p>Inundated area in Northern Marshes as captured by the inundation maps developed with Landsat 5, 7, and 8 datasets using the RaFMIC approach [<a href="#B31-remotesensing-16-03310" class="html-bibr">31</a>].</p>
Full article ">Figure 3
<p>Annual probability of inundation over the Marshes from 1988 to 2020, as captured by the Landsat 5-, 7-, and 8-based inundation maps [<a href="#B31-remotesensing-16-03310" class="html-bibr">31</a>]. Landsat 5-, 7-, and 8-derived maps are demarcated using red, blue, and green map borders, respectively.</p>
Full article ">Figure 4
<p>(<b>a</b>) Annual stream discharge to the Marshes from Marebone Weir (#421090) and Marebone Break (#421088), and (<b>b</b>) annual rainfall captured by the PERSIANN data and rain gauges, #051042 and #051057.</p>
Full article ">Figure 5
<p>(<b>a</b>–<b>c</b>) Inundation probability maps derived from Landsat 5-, 7-, and 8-based inundation maps classified into ten inundation probability classes. (<b>d</b>) Classified inundation probability map based on all the Landsat-derived inundation maps (i.e., Landsat 5, 7, and 8), collectively. <span class="html-italic">n</span> is the number of inundation maps used to develop each probability of inundation map. (<b>e</b>) Classification of vegetation over the Northern Marshes in 2013 based on Bowen et al. [<a href="#B75-remotesensing-16-03310" class="html-bibr">75</a>].</p>
Full article ">Figure 6
<p>Time record of inundated area in each probability of inundation class over the Northern Marshes as captured collectively by the inundation maps derived from Landsat 5, 7, and 8 image collections. Orange dots indicate each data point.</p>
Full article ">Figure 7
<p>(<b>a</b>) The three simplified vegetation patches dominated by river red gum (RRG) forest, RRG woodland, and shrubland (which encompasses common reed, mixed marsh/water couch, and terrestrial vegetation) over the Northern Marshes. (<b>b</b>–<b>d</b>) Time series of percentage areal inundation over the three vegetation patches as collectively captured by the Landsat 5-, 7-, and 8-based inundation maps.</p>
Full article ">Figure 8
<p>Time series between the percentage areal inundation and areal average NDVI values over the three vegetation patches: (<b>a</b>) river red gum forest, (<b>b</b>) river red gum woodland, and (<b>c</b>) shrubland in the Northern Marshes as captured by the Landsat 7-based inundation and NDVI products. (<b>d</b>–<b>f</b>) Linear regressions between Percentage areal inundation and NDVI over the three vegetation patches in the Northern Marshes as captured collectively by Landsat 5,- 7-, and 8-based products.</p>
Full article ">Figure 9
<p>Inundation probability of generally (<b>a</b>) dry, (<b>b</b>) normal, and (<b>c</b>) wet years as captured by Landsat 8-based inundation maps over the Northern Marshes with the three vegetation patches. (<b>d</b>) Average rainfall over the area from gauges #051042 and #051057 with the total discharge of Marebone Weir and Marebone Break from 2013 to 2019.</p>
Full article ">Figure 10
<p>(<b>a</b>) Two drying events as captured by the Landsat 5-based inundation maps from 8 September 1990 to 30 January 1991 (six inundation maps) and from 29 August 1998 to 5 February 1999 (five inundation maps). (<b>b</b>,<b>c</b>) Probability of inundation captured by the Landsat 5-based inundation maps during these two drying events. (<b>d</b>) SAGA Wetness Index (SWI) over the area derived from the 1 m LiDAR-derived DEM.</p>
Full article ">
16 pages, 7535 KiB  
Article
Satellite Observations Reveal Northward Vegetation Greenness Shifts in the Greater Mekong Subregion over the Past 23 Years
by Bowen Deng, Chenli Liu, Enwei Zhang, Mengjiao He, Yawen Li and Xingwu Duan
Remote Sens. 2024, 16(17), 3302; https://doi.org/10.3390/rs16173302 - 5 Sep 2024
Viewed by 385
Abstract
The Greater Mekong Subregion (GMS) economic cooperation program is an effective and fruitful regional cooperation initiative for socioeconomic development in Asia; however, the vegetation change trends and directions in the GMS caused by rapid development remain unknown. In particular, there is a current [...] Read more.
The Greater Mekong Subregion (GMS) economic cooperation program is an effective and fruitful regional cooperation initiative for socioeconomic development in Asia; however, the vegetation change trends and directions in the GMS caused by rapid development remain unknown. In particular, there is a current lack of comparative studies on vegetation changes in various countries in the GMS. Based on the MODIS normalized difference vegetation index (NDVI) time series data, this study analyzed the spatiotemporal patterns of vegetation coverage and their trends in the GMS from 2000 to 2022 using the Theil–Sen slope estimation, the Mann–Kendall mutation test, and the gravity center migration model. The key findings were as follows: (1) the NDVI in the GMS showed an overall upward fluctuating trend over the past 23 years, with an annual growth rate of 0.11%. The NDVI changes varied slightly between seasons, with the greatest increases recorded in summer and winter. (2) The spatial distribution of NDVI in the GMS varied greatly, with higher NDVI values in the north–central region and lower NDVI values in the south. (3) A total of 66.03% of the GMS area showed increments in vegetation during the studied period, mainly in south–central Myanmar, northeastern Thailand, Vietnam, and China. (4) From 2000 to 2022, the gravity center of vegetation greenness shifted northward in the GMS, especially from 2000 to 2005, indicating that the growth rates of vegetation in the north–central part of the GMS were higher than those in the south. Furthermore, the vegetation coverage in all countries, except Cambodia, increased, with the most pronounced growth recorded in China. Overall, these findings can provide scientific evidence for the GMS to enhance ecological protection and sustainable development. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area. (<b>a</b>) Elevation and location of the Great Mekong Subregion (GMS); and (<b>b</b>) land use types in 2020.</p>
Full article ">Figure 2
<p>Spatial distribution of normalized difference vegetation index (NDVI) in the GMS across the studied period (2000–2022). (<b>a</b>) Map showing average annual NDVI, (<b>b</b>) plot showing the total area covered by different NDVI classes, and (<b>c</b>) plot showing the area percentages of each country represented by different NDVI classes.</p>
Full article ">Figure 3
<p>Plots showing temporal variations in NDVI from 2000 to 2022: (<b>a</b>) interannual, (<b>b</b>) spring, (<b>c</b>) summer, (<b>d</b>) autumn, and (<b>e</b>) winter. Plots showing Mann–Kendall mutation test results from 2000 to 2022: (<b>f</b>) interannual, (<b>g</b>) spring, (<b>h</b>) summer, (<b>i</b>) autumn, and (<b>j</b>) winter.</p>
Full article ">Figure 4
<p>Maps showing the spatial distribution of vegetation change trends in the GMS: (<b>a</b>) Sen’s slope estimation, (<b>b</b>) Mann–Kendall significance test, and (<b>c</b>) Sen-MK trend-coupling type.</p>
Full article ">Figure 5
<p>Plots showing temporal variations in annual average NDVI values in the GMS countries from 2000 to 2022: (<b>a</b>) China, (<b>b</b>) Myanmar, (<b>c</b>) Thailand, (<b>d</b>) Laos, (<b>e</b>) Vietnam, and (<b>f</b>) Cambodia.</p>
Full article ">Figure 6
<p>Spatial variation of NDVI in different countries in the GMS. (<b>a</b>) Map showing NDVI spatial variation across the GMS, (<b>b</b>) plot showing trends in NDVI changes across different countries, and (<b>c</b>) plot showing percentage change in NDVI across different countries.</p>
Full article ">Figure 7
<p>Trends in NDVI center of gravity migration in the GMS for the period 2000–2022. (<b>a</b>) Standard deviation ellipse, (<b>b</b>) center of gravity migration trajectory, and (<b>c</b>) interannual migration distance.</p>
Full article ">
22 pages, 4249 KiB  
Article
Estimating Methane Emissions in Rice Paddies at the Parcel Level Using Drone-Based Time Series Vegetation Indices
by Yongho Song, Cholho Song, Sol-E Choi, Joon Kim, Moonil Kim, Wonjae Hwang, Minwoo Roh, Sujong Lee and Woo-Kyun Lee
Drones 2024, 8(9), 459; https://doi.org/10.3390/drones8090459 - 4 Sep 2024
Viewed by 526
Abstract
This study investigated a method for directly estimating methane emissions from rice paddy fields at the field level using drone-based time-series vegetation indices at a town scale. Drone optical and spectral images were captured approximately 15 times from April to November to acquire [...] Read more.
This study investigated a method for directly estimating methane emissions from rice paddy fields at the field level using drone-based time-series vegetation indices at a town scale. Drone optical and spectral images were captured approximately 15 times from April to November to acquire time-series vegetation indices and optical orthoimages. An empirical regression model validated in previous international studies was applied to calculate cumulative methane emissions throughout the rice cultivation process. Methane emissions were estimated using the vegetation index and yield data were used as input variables for each growth phase. Methane emissions from rice paddies showed maximum values of 309 kg CH4 ha−1, within a 7% range compared to similar studies, and minimum values of 138 kg CH4 ha−1, with differences ranging from 29% to 58%. The average emissions were calculated at 247 kg CH4/ha, revealing slightly lower average values but individual field values within a similar range. The results suggest that drone-based remote sensing technology is an efficient and cost-effective alternative to traditional field measurements for greenhouse gas emission assessments. However, adjustments and validations according to rice varieties and local cultivation environments are necessary. Overcoming these limitations can help establish sustainable agricultural management practices and achieve local greenhouse gas reduction targets. Full article
(This article belongs to the Section Drones in Agriculture and Forestry)
Show Figures

Figure 1

Figure 1
<p>Map of the Siu-ri in Gyeonggi-do, South Korea, showing the location of this study area. Red areas indicate the rice paddies in the study field.</p>
Full article ">Figure 2
<p>Flowchart of overall methodology.</p>
Full article ">Figure 3
<p>DEM equidistant flight method used by FMS.</p>
Full article ">Figure 4
<p>Temporal dynamics of parcel level average vegetation index. (<b>a</b>) NDVI, (<b>b</b>) GNDVI, (<b>c</b>) NDRE, (<b>d</b>) OSAVI. # is the drone flight number in <a href="#drones-08-00459-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 5
<p>Utilizing temporal orthoimage for determining harvest day. The red areas indicate the paddy field boundaries, and the numbers represent the indices of the paddy fields. (<b>a</b>) presents images taken shortly before harvest, showing that most areas are about to be harvested, except for fields 33 and 34, which have already been harvested. (<b>b</b>) displays images taken after the harvest, indicating that harvesting has been completed in most areas, except for a few.</p>
Full article ">Figure 5 Cont.
<p>Utilizing temporal orthoimage for determining harvest day. The red areas indicate the paddy field boundaries, and the numbers represent the indices of the paddy fields. (<b>a</b>) presents images taken shortly before harvest, showing that most areas are about to be harvested, except for fields 33 and 34, which have already been harvested. (<b>b</b>) displays images taken after the harvest, indicating that harvesting has been completed in most areas, except for a few.</p>
Full article ">Figure 6
<p>Thematic map of methane emissions by rice paddy field (<b>a</b>) EVI2-JS, (<b>b</b>) EVI2-JS-HS, (<b>c</b>) EVI2-HS-GS, (<b>d</b>) EVI2-AS (Unit: g CH<sub>4</sub>/pixel).</p>
Full article ">Figure 7
<p>Identification of abnormal methane emission regions using drone optical imaging-based methane emission model results.</p>
Full article ">Figure 8
<p>Time series EVI2 value comparison by paddy field. (<b>a</b>) Fields showing similar trends were indicated with different colors using points and lines. Fields 33 and 34 showed a rapid decline in EVI2 values after the heading stage compared to other plots. (<b>b</b>) Optical image verification indicated that early harvesting was conducted.</p>
Full article ">Figure 9
<p>Methane emission estimation by rice paddy field area.</p>
Full article ">Figure 10
<p>Box plot for methane emissions per unit area by models.</p>
Full article ">
24 pages, 6269 KiB  
Article
Aboveground Biomass Mapping in SemiArid Forests by Integrating Airborne LiDAR with Sentinel-1 and Sentinel-2 Time-Series Data
by Linjing Zhang, Xinran Yin, Yaru Wang and Jing Chen
Remote Sens. 2024, 16(17), 3241; https://doi.org/10.3390/rs16173241 - 1 Sep 2024
Viewed by 649
Abstract
Aboveground biomass (AGB) is a vital indicator for studying carbon sinks in forest ecosystems. Semiarid forests harbor substantial carbon storage but received little attention due to the high spatial–temporal heterogeneity that complicates the modeling of AGB in this environment. This study assessed the [...] Read more.
Aboveground biomass (AGB) is a vital indicator for studying carbon sinks in forest ecosystems. Semiarid forests harbor substantial carbon storage but received little attention due to the high spatial–temporal heterogeneity that complicates the modeling of AGB in this environment. This study assessed the performance of different data sources (annual monthly time-series radar was Sentinel-1 [S1]; annual monthly time series optical was Sentinel-2 [S2]; and single-temporal airborne light detection and ranging [LiDAR]) and seven prediction approaches to map AGB in the semiarid forests on the border between Gansu and Qinghai Provinces in China. Five experiments were conducted using different data configurations from synthetic aperture radar backscatter, multispectral reflectance, LiDAR point cloud, and their derivatives (polarimetric combination indices, texture information, vegetation indices, biophysical features, and tree height- and canopy-related indices). The results showed that S2 acquired better prediction (coefficient of determination [R2]: 0.62–0.75; root mean square error [RMSE]: 30.08–38.83 Mg/ha) than S1 (R2: 0.24–0.45; RMSE: 47.36–56.51 Mg/ha). However, their integration further improved the results (R2: 0.65–0.78; RMSE: 28.68–35.92 Mg/ha). The addition of single-temporal LiDAR highlighted its structural importance in semiarid forests. The best mapping accuracy was achieved by XGBoost, with the metrics from the S2 and S1 time series and the LiDAR-based canopy height information being combined (R2: 0.87; RMSE: 21.63 Mg/ha; relative RMSE: 14.45%). Images obtained during the dry season were effective for AGB prediction. Tree-based models generally outperformed other models in semiarid forests. Sequential variable importance analysis indicated that the most important S1 metric to estimate AGB was the polarimetric combination indices sum, and the S2 metrics were associated with red-edge spectral regions. Meanwhile, the most important LiDAR metrics were related to height percentiles. Our methodology advocates for an economical, extensive, and precise AGB retrieval tailored for semiarid forests. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Study area: (<b>a</b>) a geographical position map of Qinghai and Gansu Provinces in China; (<b>b</b>) a geographical position map of the study area at the border between Qinghai and Gansu Provinces; and (<b>c</b>) a true color image of the study area formed by clipping an S2 image acquired on 11 August 2019.</p>
Full article ">Figure 2
<p>Data-acquisition timeline: (<b>a</b>) time of acquisition for the S1 and S2 single image; (<b>b</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2019; (<b>c</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2020; and (<b>d</b>) time of acquisition for the synthetic cloud-free S2 images from May to Jul 2021.</p>
Full article ">Figure 3
<p>Flowchart of the study.</p>
Full article ">Figure 4
<p>Accuracy evaluation for different models and data sources. The bar represents R<sup>2</sup>, and the line represents RMSE<sub>r</sub>.</p>
Full article ">Figure 5
<p>Scatter plots of the measured AGB and predicted AGB from the best performance combination (experiment E-S1S2LiDAR) for different models (<b>a</b>) RF; (<b>b</b>) XGBoost; (<b>c</b>) SGB; (<b>d</b>) CNN; (<b>e</b>) GPR; (<b>f</b>) MLP; (<b>g</b>) LASSO.</p>
Full article ">Figure 6
<p>Flow chart of the sequential forward selection.</p>
Full article ">Figure 7
<p>Variable importance plots of the top 15 predictors for the XGBoost model with three combined datasets: (<b>a</b>) experiment C; (<b>b</b>) experiment D; (<b>c</b>) experiment E.</p>
Full article ">Figure 8
<p>AGB map in the study area from the XGBoost model with the optical, SAR, and LiDAR metrics.</p>
Full article ">
21 pages, 15871 KiB  
Article
Tracking Forest Disturbance in Northeast China’s Cold-Temperate Forests Using a Temporal Sequence of Landsat Data
by Yueting Wang, Xiang Jia, Xiaoli Zhang, Lingting Lei, Guoqi Chai, Zongqi Yao, Shike Qiu, Jun Du, Jingxu Wang, Zheng Wang and Ran Wang
Remote Sens. 2024, 16(17), 3238; https://doi.org/10.3390/rs16173238 - 1 Sep 2024
Viewed by 578
Abstract
Cold-temperate forests (CTFs) are not only an important source of wood but also provide significant carbon storage in China. However, under the increasing pressure of human activities and climate change, CTFs are experiencing severe disturbances, such as logging, fires, and pest infestations, leading [...] Read more.
Cold-temperate forests (CTFs) are not only an important source of wood but also provide significant carbon storage in China. However, under the increasing pressure of human activities and climate change, CTFs are experiencing severe disturbances, such as logging, fires, and pest infestations, leading to evident degradation trends. Though these disturbances impact both regional and global carbon budgets and their assessments, the disturbance patterns in CTFs in northern China remain poorly understood. In this paper, the Genhe forest area, which is a typical CTF region located in the Inner Mongolia Autonomous Region, Northeast China (with an area of about 2.001 × 104 km2), was selected as the study area. Based on Landsat historical archived data on the Google Earth Engine (GEE) platform, we used the continuous change detection and classification (CCDC) algorithm and considered seasonal features to detect forest disturbances over nearly 30 years. First, we created six inter-annual time series seasonal vegetation index datasets to map forest coverage using the maximum between-class variance algorithm (OTSU). Second, we used the CCDC algorithm to extract disturbance information. Finally, by using the ECMWF climate reanalysis dataset, MODIS C6, the snow phenology dataset, and forestry department records, we evaluated how disturbances relate to climate and human activities. The results showed that the disturbance map generated using summer (June–August) imagery and the enhanced vegetation index (EVI) had the highest overall accuracy (88%). Forests have been disturbed to the extent of 12.65% (2137.31 km2) over the last 30 years, and the disturbed area generally showed a trend toward reduction, especially after commercial logging activities were banned in 2015. However, there was an unusual increase in the number of disturbed areas in 2002 and 2003 due to large fires. The monitoring of potential widespread forest disturbance due to extreme drought and fire events in the context of climate change should be strengthened in the future, and preventive and salvage measures should be taken in a timely manner. Our results demonstrate that CTF disturbance can be robustly mapped by using the CCDC algorithm based on Landsat time series seasonal imagery in areas with complex meteorological conditions and spatial heterogeneity, which is essential for understanding forest change processes. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area.</p>
Full article ">Figure 2
<p>Forest disturbance analysis workflow.</p>
Full article ">Figure 3
<p>Mapping the forest. (<b>a</b>) OTSU value and forest area every year (Summer &amp; EVI); (<b>b</b>) forest cover synthesis map from 1990 to 2021 (Summer &amp; EVI).</p>
Full article ">Figure 4
<p>Schematic diagram of typical disturbance types and disturbance processes. (<b>a</b>) Logging in 1999; (<b>b</b>) anthropogenic fire in 2003; (<b>c</b>) wildfires in 2003 and 2010, respectively; (<b>d</b>) logging in 1990 and anthropogenic fire in 2003; (<b>e</b>) EVI curves for the sample points in disturbance areas from (<b>a</b>–<b>d</b>), with red boxes indicating disturbance events.</p>
Full article ">Figure 5
<p>Forest disturbance extraction. (<b>a</b>) Forest disturbance zone; (<b>b</b>) disturbance caused by logging after 1990; (<b>c</b>) disturbance caused by man-made fire in 2003; (<b>d</b>) other disturbances caused by multiple factors such as wildfire, etc.; <b>b1</b>–<b>d1</b> show the results of extracting forest disturbance information, <b>b2</b>–<b>d2</b> display the satellite images that correspond to these areas after the disturbance has occurred; (<b>e</b>) forest disturbance zone after fire-induced disturbances have been removed; (<b>f</b>) annual forest disturbance area caused by fires and other factors; the lines in the plot are the univariate linear trendlines.</p>
Full article ">Figure 6
<p>Distance of forest disturbance patches from roads and rivers. (<b>a</b>,<b>c</b>) represent the distance of disturbance patches from roads; (<b>b</b>,<b>d</b>) represent the distance of disturbance patches from rivers.</p>
Full article ">Figure 7
<p>Number of forest disturbance events. (<b>a</b>) Number of forest disturbance events in Genhe; (<b>b</b>) number of forest disturbance events in each administrative unit.</p>
Full article ">Figure 8
<p>The relationship between forest disturbance and its influencing factors. a<sub>i</sub>, b<sub>i</sub>, c<sub>i</sub>, d<sub>i</sub>, and e<sub>i</sub> are models of the area of disturbance and its influencing factors (annual precipitation, annual average temperature, annual snow cover days, the annual number of fires, and annual commercial logging output, respectively) for every year; the pink circles are for the anomalous years (2002 and 2003); the period of (<b>a<sub>1</sub></b>–<b>d<sub>1</sub></b>) is from 1991 to 2020; the period of (<b>a<sub>2</sub></b>–<b>d<sub>2</sub></b>) is from 2011 to 2020; the period of (<b>a<sub>3</sub></b>–<b>c<sub>3</sub></b>) is from 1991 to 2020, the disturbance area of (<b>a<sub>3</sub></b>–<b>c<sub>3</sub></b>) is the disturbance caused by factors other than fire; (<b>d<sub>3</sub></b>) is the model of the disturbance area and burned area for every year from 1991 to 2020; (<b>e<sub>1</sub></b>) annual commercial logging output; the period of (<b>e<sub>2</sub></b>,<b>e<sub>3</sub></b>) is from 1991 to 2015; the disturbance area of (<b>e<sub>3</sub></b>) is the disturbance caused by factors other than fire; the red line in the figure is the univariate linear trendline.</p>
Full article ">Figure 9
<p>Time and location of forest disturbance and fires. Note: the blue, pink, and purple areas are the areas where disturbances and fires occurred. It should be noted that the purple areas are areas where disturbances and fires occurred in the same year, and other areas with colors (except white and gray) are all where disturbances were detected but the region of fire did not exist in the auxiliary dataset.</p>
Full article ">
37 pages, 76788 KiB  
Article
Machine Learning-Based Remote Sensing Inversion of Non-Photosynthetic/Photosynthetic Vegetation Coverage in Desertified Areas and Its Response to Drought Analysis
by Zichen Guo, Shulin Liu, Kun Feng, Wenping Kang and Xiang Chen
Remote Sens. 2024, 16(17), 3226; https://doi.org/10.3390/rs16173226 - 31 Aug 2024
Viewed by 521
Abstract
Determining the responses of non-photosynthetic vegetation (NPV) and photosynthetic vegetation (PV) communities to climate change is crucial in illustrating the sensitivity and sustainability of these ecosystems. In this study, we evaluated the accuracy of inverting NPV and PV using Landsat imagery with random [...] Read more.
Determining the responses of non-photosynthetic vegetation (NPV) and photosynthetic vegetation (PV) communities to climate change is crucial in illustrating the sensitivity and sustainability of these ecosystems. In this study, we evaluated the accuracy of inverting NPV and PV using Landsat imagery with random forest (RF), backpropagation neural network (BPNN), and fully connected neural network (FCNN) models. Additionally, we inverted MODIS NPV and PV time-series data using spectral unmixing. Based on this, we analyzed the responses of NPV and PV to precipitation and drought across different ecological regions. The main conclusions are as follows: (1) In NPV remote sensing inversion, the softmax activation function demonstrates greater advantages over the ReLU activation function. Specifically, the use of the softmax function results in an approximate increase of 0.35 in the R2 value. (2) Compared with a five-layer FCNN with 128 neurons and a three-layer BPNN with 12 neurons, a random forest model with over 50 trees and 5 leaf nodes provides better inversion results for NPV and PV (R2_RF-NPV = 0.843, R2_RF-PV = 0.861). (3) Long-term drought or heavy rainfall events can affect the utilization of precipitation by NPV and PV. There is a high correlation between extreme precipitation events following prolonged drought and an increase in PV coverage. (4) Under long-term drought conditions, the vegetation in the study area responded to precipitation during the last winter and growing season. This study provides an illustration of the response of semi-arid ecosystems to drought and wetting events, thereby offering a data basis for the effect evaluation of afforestation projects. Full article
(This article belongs to the Special Issue Machine Learning for Spatiotemporal Remote Sensing Data (2nd Edition))
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Desertification types in the study area; (<b>b</b>) annual mean monthly precipitation in the study area; (<b>c</b>) location of the study area in a semi-arid region of China.</p>
Full article ">Figure 2
<p>Technical workflow diagram of this study.</p>
Full article ">Figure 3
<p>Box plot of non-photosynthetic vegetation coverage of different desertification types and degrees.</p>
Full article ">Figure 4
<p>Box plot of photosynthetic vegetation coverage of different desertification types and degrees.</p>
Full article ">Figure 5
<p>(<b>a</b>) Response of non-photosynthetic vegetation to annual precipitation. (<b>b</b>) Response of photosynthetic vegetation to annual precipitation. (<b>c</b>) Response of non-photosynthetic vegetation to annual mean temperature. (<b>d</b>) Response of photosynthetic vegetation to annual mean temperature.</p>
Full article ">Figure 6
<p>(<b>a</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on mobile dune desertification during Dry Years. (<b>b</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on coppice dune desertification during Dry Years. (<b>c</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on Gobi desertification during Dry Years. (<b>d</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on mobile dune desertification during Wet Years. (<b>e</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on coppice dune desertification Wet Years. (<b>f</b>) Time-Lagged Response of NPV and PV to Monthly Precipitation on Gobi desertification Wet Years. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. K is the response degree of NPV and PV to precipitation in the desertification type and degree region.</p>
Full article ">Figure 7
<p>(<b>a</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on mobile dune desertification during Dry Years. (<b>b</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on coppice dune desertification during Dry Years. (<b>c</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on Gobi desertification during Dry Years. (<b>d</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on mobile dune desertification during Wet Years. (<b>e</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on coppice dune desertification Wet Years. (<b>f</b>) Time-delay correlation (R<sup>2</sup>) of NPV and PV to Monthly Precipitation on Gobi desertifi-cation Wet Years. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. R<sup>2</sup> is the correlation.</p>
Full article ">Figure 8
<p>(<b>a</b>) Response of NPV and PV to SPEI on mobile dune desertification in the Year 2000. (<b>b</b>) Response of NPV and PV to SPEI on coppice dune desertification in the Year 2000. (<b>c</b>) Response of NPV and PV to SPEI on Gobi desertification in the Year 2000. (<b>d</b>) Response of NPV and PV to SPEI on mobile dune desertification in the Year 2005. (<b>e</b>) Response of NPV and PV to SPEI on coppice dune desertification in the Year 2005. (<b>f</b>) Response of NPV and PV to SPEI on Gobi desertification in the Year 2005. (<b>g</b>) Response of NPV and PV to SPEI on mobile dune desertification in the Year 2010. (<b>h</b>) Response of NPV and PV to SPEI on coppice dune desertification in the Year 2010. (<b>i</b>) Response of NPV and PV to SPEI on Gobi desertification in the Year 2010. (<b>j</b>) Response of NPV and PV to SPEI on mobile dune desertification in the Year 2015. (<b>k</b>) Response of NPV and PV to SPEI on coppice dune desertification in the Year 2015. (<b>l</b>) Response of NPV and PV to SPEI on Gobi desertification in the Year 2015. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. K is the response degree of NPV and PV to precipitation in the desertification type and degree region.</p>
Full article ">Figure 9
<p>(<b>a</b>) Correlation of NPV and PV to SPEI on mobile dune desertification in the Year 2000. (<b>b</b>) Correlation of NPV and PV to SPEI on coppice dune desertification in the Year 2000. (<b>c</b>) Correlation of NPV and PV to SPEI on Gobi desertification in the Year 2000. (<b>d</b>) Correlation of NPV and PV to SPEI on mobile dune desertification in the Year 2005. (<b>e</b>) Correlation of NPV and PV to SPEI on coppice dune desertification in the Year 2005. (<b>f</b>) Correlation of NPV and PV to SPEI on Gobi desertification in the Year 2005. (<b>g</b>) Correlation of NPV and PV to SPEI on mobile dune desertification in the Year 2010. (<b>h</b>) Correlation of NPV and PV to SPEI on coppice dune desertification in the Year 2010. (<b>i</b>) Correlation of NPV and PV to SPEI on Gobi desertification in the Year 2010. (<b>j</b>) Correlation of NPV and PV to SPEI on mobile dune desertification in the Year 2015. (<b>k</b>) Response of NPV and PV to SPEI on coppice dune desertification in the Year 2015. (<b>l</b>) Response of NPV and PV to SPEI on Gobi desertification in the Year 2015. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. R<sup>2</sup> is the correlation.</p>
Full article ">Figure 10
<p>(<b>a</b>) Local NPV cover of MODIS in 2019; (<b>b</b>) SWIR67 Index in 2019; (<b>c</b>) Local NPV of MODIS in 2018.</p>
Full article ">Figure 11
<p>Spectra of photosynthetic and non-photosynthetic components of major vegetation types and major soil types in the study area.</p>
Full article ">Figure A1
<p>(<b>a</b>) Process of parameter selection for RF models. (<b>b</b>) Error distribution of random forest models. (<b>c</b>) Process of parameter selection for BPNN. (<b>d</b>) Process of parameter selection for FCNN.</p>
Full article ">Figure A2
<p>(<b>a</b>) The regression relationship between Landsat 8 B<sub>2</sub> and Landsat 5 B<sub>1</sub>. (<b>b</b>) The regression relationship between Landsat 8 B<sub>3</sub> and Landsat 5 B<sub>2</sub>. (<b>c</b>) The regression relationship between Landsat 8 B<sub>4</sub> and Landsat 5 B<sub>3</sub>. (<b>d</b>) The regression relationship between Landsat 8 B<sub>5</sub> and Landsat 5 B<sub>4</sub>. (<b>e</b>) The regression relationship between Landsat 8 B<sub>6</sub> and Landsat 5 B<sub>5</sub>. (<b>f</b>) The regression relationship between Landsat 8 B<sub>7</sub> and Landsat 5 B<sub>6</sub>.</p>
Full article ">Figure A3
<p>(<b>a</b>) Non-photosynthetic vegetation coverage at the end of the growing season in 2000. (<b>b</b>) Photosynthetic vegetation coverage at the end of the growing season in 2000.</p>
Full article ">Figure A4
<p>(<b>a</b>) Non-photosynthetic vegetation coverage at the end of the growing season in 2005. (<b>b</b>) Photosynthetic vegetation coverage at the end of the growing season in 2005.</p>
Full article ">Figure A5
<p>(<b>a</b>) Non-photosynthetic vegetation coverage at the end of the growing season in 2010. (<b>b</b>) Photosynthetic vegetation coverage at the end of the growing season in 2010.</p>
Full article ">Figure A6
<p>(<b>a</b>) Non-photosynthetic vegetation coverage at the end of the growing season in 2015. (<b>b</b>) Photosynthetic vegetation coverage at the end of the growing season in 2015.</p>
Full article ">Figure A7
<p>(<b>a</b>) Response degree of non-photosynthetic vegetation at the end of growing season to total precipitation in 1–9 months. (<b>b</b>) Response degree of non-photosynthetic vegetation at the end of growing season to total precipitation in 10–21 months. (<b>c</b>) Response degree of non-photosynthetic vegetation at the end of growing season to total precipitation in 22–33 months. (<b>d</b>) Response degree of non-photosynthetic vegetation at the end of growing season to total precipitation in 34–45 months. (<b>e</b>) Response degree of non-photosynthetic vegetation at the end of growing season to total precipitation in 46–57 months.</p>
Full article ">Figure A8
<p>(<b>a</b>) Response degree of photosynthetic vegetation at the end of growing season to total precipitation in 0–9 months. (<b>b</b>) Response degree of photosynthetic vegetation at the end of growing season to total precipitation in 9–21 months. (<b>c</b>) Response degree of photosynthetic vegetation at the end of growing season to total precipitation in 21–33 months. (<b>d</b>) Response degree of photosynthetic vegetation at the end of growing season to total precipitation in 33–45 months. (<b>e</b>) Response degree of photosynthetic vegetation at the end of growing season to total precipitation in 45–57 months.</p>
Full article ">Figure A9
<p>(<b>a</b>) Response of non-photosynthetic vegetation to 0–9 month mean temperature at the end of growing season. (<b>b</b>) Response of non-photosynthetic vegetation to 9–21 month mean temperature at the end of growing season. (<b>c</b>) Response of non-photosynthetic vegetation to 21–33 month mean temperature at the end of growing season. (<b>d</b>) Response of non-photosynthetic vegetation to 33–45 month mean temperature at the end of growing season. (<b>e</b>) Response of non-photosynthetic vegetation to 45–57 month mean temperature at the end of growing season.</p>
Full article ">Figure A10
<p>(<b>a</b>) Response of photosynthetic vegetation to 0–9 month mean temperature at the end of growing season. (<b>b</b>) Response of photosynthetic vegetation to 9–21 month mean temperature at the end of growing season. (<b>c</b>) Response of photosynthetic vegetation to 21–33 month mean temperature at the end of growing season. (<b>d</b>) Response of photosynthetic vegetation to 33–45 month mean temperature at the end of growing season. (<b>e</b>) Response of photosynthetic vegetation to 45–57 month mean temperature at the end of growing season.</p>
Full article ">Figure A11
<p>Time-delay responses of non-photosynthetic and photosynthetic vegetation coverage to monthly precipitation in different desertification types and degrees. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. K is the response degree of NPV and PV to precipitation in the desertification type and degree region.</p>
Full article ">Figure A12
<p>Time-delay correlation (R<sup>2</sup>) of non-photosynthetic and photosynthetic vegetation cover with monthly precipitation in different desertification types and degrees. Figure note: MBD represents mobile dune desertification; CD represents coppice dune desertification; GD represents Gobi desertification; MD represents mild desertification; MOD represents moderate desertification; and SD represents severe desertification. R<sup>2</sup> is the correlation.</p>
Full article ">Figure A13
<p>(<b>a</b>) Spatial distribution map of 1-month SPEI series in September 2000. (<b>b</b>) Spatial distribution map of 3-month SPEI series in September 2000. (<b>c</b>) Spatial distribution map of 9-month SPEI series in September 2000. (<b>d</b>) Spatial distribution map of 12-month SPEI series in September 2000.</p>
Full article ">Figure A14
<p>(<b>a</b>) Spatial distribution map of 1-month SPEI series in September 2005. (<b>b</b>) Spatial distribution map of 3-month SPEI series in September 2005. (<b>c</b>) Spatial distribution map of 9-month SPEI series in September 2005. (<b>d</b>) Spatial distribution map of 12-month SPEI series in September 2005.</p>
Full article ">Figure A15
<p>(<b>a</b>) Spatial distribution map of 1-month SPEI series in September 2010. (<b>b</b>) Spatial distribution map of 3-month SPEI series in September 2010. (<b>c</b>) Spatial distribution map of 9-month SPEI series in September 2010. (<b>d</b>) Spatial distribution map of 12-month SPEI series in September 2010.</p>
Full article ">Figure A16
<p>(<b>a</b>) Spatial distribution map of 1-month SPEI series in September 2015. (<b>b</b>) Spatial distribution map of 3-month SPEI series in September 2015. (<b>c</b>) Spatial distribution map of 9-month SPEI series in September 2015. (<b>d</b>) Spatial distribution map of 12-month SPEI series in September 2015.</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 636
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 ">
29 pages, 3577 KiB  
Review
Recent Advances in Light Penetration Depth for Postharvest Quality Evaluation of Fruits and Vegetables
by Yuping Huang, Jie Xiong, Ziang Li, Dong Hu, Ye Sun, Haojun Jin, Huichun Zhang and Huimin Fang
Foods 2024, 13(17), 2688; https://doi.org/10.3390/foods13172688 - 26 Aug 2024
Viewed by 679
Abstract
Light penetration depth, as a characteristic parameter reflecting light attenuation and transmission in biological tissues, has been applied in nondestructive detection of fruits and vegetables. Recently, with emergence of new optical detection technologies, researchers have begun to explore methods evaluating optical properties of [...] Read more.
Light penetration depth, as a characteristic parameter reflecting light attenuation and transmission in biological tissues, has been applied in nondestructive detection of fruits and vegetables. Recently, with emergence of new optical detection technologies, researchers have begun to explore methods evaluating optical properties of double-layer or even multilayer fruit and vegetable tissues due to the differences between peel and pulp in the chemical composition and physical properties, which has gradually promoted studies on light penetration depth. A series of demonstrated research on light penetration depth could ensure the accuracy of the optical information obtained from each layer of tissue, which is beneficial to enhance detection accuracy for quality assessment of fruits and vegetables. Therefore, the aim of this review is to give detailed outlines about the theory and principle of light penetration depth based on several emerging optical detection technologies and to focus primarily on its applications in the field of quality evaluation of fruits and vegetables, its future applicability in fruits and vegetables and the challenges it may face in the future. Full article
(This article belongs to the Section Food Packaging and Preservation)
Show Figures

Figure 1

Figure 1
<p>Schematic of the interaction between light and an object.</p>
Full article ">Figure 2
<p>Energy variations of Rayleigh scattering and Raman scattering (energy level: Em &lt; En).</p>
Full article ">Figure 3
<p>Relationship between light attenuation and light penetration depth in tissues.</p>
Full article ">Figure 4
<p>(<b>a</b>) Schematic representation of extrapolated boundary; (<b>b</b>) MC simulation for diffuse reflectance and absorption of tissues; (<b>c</b>) transmission process of photons in the AD method.</p>
Full article ">Figure 5
<p>(<b>a</b>) Flowchart of the MC simulation of a single photon; (<b>b</b>) flowchart of the IAD method.</p>
Full article ">Figure 6
<p>(<b>a</b>) Short-pulsed illumination at the surface of a semi-infinite turbid medium; (<b>b</b>) schematic of time-resolved system for measuring optical properties, in which PMT is a photomultiplier tube and SYNC is the synchronization signal.</p>
Full article ">Figure 7
<p>(<b>a</b>) Schematic illustrations of configuration of single-fiber and “banana-shape” path of light transfer; (<b>b</b>) multifiber array based on a multiplexer; (<b>c</b>) multifiber array based on a multiplexer; (<b>d</b>) multichannel curved array based on spatially resolved system.</p>
Full article ">Figure 8
<p>(<b>a</b>) Schematic illustrations of noncontact SRS systems; (<b>b</b>) schematic of an SFDI system for spectral image acquisition.</p>
Full article ">
Back to TopTop