[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (2,090)

Search Parameters:
Keywords = cropland

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
16 pages, 8588 KiB  
Article
A Novel Approach for Farmland Size Estimation in Small-Scale Agriculture Using Edge Counting and Remote Sensing
by Jingnan Du, Sucheng Xu, Jinshan Li, Jiakun Duan and Wu Xiao
Remote Sens. 2024, 16(16), 2981; https://doi.org/10.3390/rs16162981 (registering DOI) - 14 Aug 2024
Abstract
Accurate and timely information on farmland size is crucial for agricultural development, resource management, and other related fields. However, there is currently no mature method for estimating farmland size in smallholder farming areas. This is due to the small size of farmland plots [...] Read more.
Accurate and timely information on farmland size is crucial for agricultural development, resource management, and other related fields. However, there is currently no mature method for estimating farmland size in smallholder farming areas. This is due to the small size of farmland plots in these areas, which have unclear boundaries in medium and high-resolution satellite imagery, and irregular shapes that make it difficult to extract complete boundaries using morphological rules. Automatic farmland mapping algorithms using remote sensing data also perform poorly in small-scale farming areas. To address this issue, this study proposes a farmland size evaluation index based on edge frequency (ECR). The algorithm utilizes the high temporal resolution of Sentinel-2 satellite imagery to compensate for its spatial resolution limitations. First, all Sentinel-2 images from one year are used to calculate edge frequencies, which can divide farmland areas into low-value farmland interior regions, medium-value non-permanent edges, and high-value permanent edges (PE). Next, the Otsu’s thresholding algorithm is iteratively applied twice to the edge frequencies to first extract edges and then permanent edges. The ratio of PE to cropland (ECR) is then calculated. Using the North China Plain and Northeast China Plain as study areas, and comparing with existing farmland size datasets, the appropriate estimation radius for ECR was determined to be 1600 m. The study found that the peak ECR value for the Northeast China Plain was 0.085, and the peak value for the North China Plain was 0.105. The overall distribution was consistent with the reference dataset. Full article
(This article belongs to the Special Issue Advances in Remote Sensing for Crop Monitoring and Food Security)
Show Figures

Figure 1

Figure 1
<p>The red borders outline the six major grain-producing provinces in China, which serve as our study area. From top to bottom, left to right, they are Inner Mongolia and Liaoning Province in the Northeast China Plain, and Shandong Province, Henan Province, Hubei Province, and Anhui Province in the North China Plain. There are a total of 1792 sample points from the Geo-Wiki plot size dataset that fall within the study area. Both color and size are used to display the points based on plot size for better visualization.</p>
Full article ">Figure 2
<p>Technical flowchart of this study. Abbreviations: NDVI stands for normalized difference vegetation index; Otsu, Otsu binary segmentation algorithm; ECR, edge cropland ratio [<a href="#B6-remotesensing-16-02981" class="html-bibr">6</a>].</p>
Full article ">Figure 3
<p>The figure illustrates the Edge count generation process within a 3200 m radius of a sample point (Sample ID: 962700, Latitude: 45.417702, Longitude: 121.545998) located in the Inner Mongolia Autonomous Region. Edge count represents the number of times each pixel is marked as an edge. In the grayscale image, brighter areas indicate higher counts, while pure black areas represent non-farmland regions.</p>
Full article ">Figure 4
<p>(<b>a</b>) shows the buffer zones of different radii for the sample point, ranging from 200 m to 3200 m. (<b>b</b>) displays the edge count images obtained at different radii. The edge count result for the 3200 m radius is shown in (<b>c</b>), with non-farmland areas set as transparent. (<b>d</b>–<b>f</b>), respectively, represent the extracted edges, permanent edges, and edge frequency distribution of (<b>c</b>).</p>
Full article ">Figure 5
<p>The boxplot illustrates the distribution of estimated ECR values at different radii. The numerical statistics are shown in the right figure.</p>
Full article ">Figure 6
<p>The distribution of ECR for each parcel size group is shown in separate figures with different radii.</p>
Full article ">Figure 7
<p>The line chart on the left shows the probability density distribution of ECRs at different radii, while the bar chart on the right shows the number of Field size labels at each level, categorized by province.</p>
Full article ">Figure 8
<p>The figure presents three samples of different farmland sizes in separate columns, from left to right: XL, M, and XS. The three rows from top to bottom are: Google Satellite basemap and sample’s metadata, edge count image and grayscale histogram (upper right corner), and identified permanent edges (in red).</p>
Full article ">Figure 9
<p>The probability density curves of the 1600 m ECRs are displayed, grouped by parcel size. The three vertical dashed lines in the left figure represent the optimal thresholds for classifying farmland size based on ECRs, determined using Spearman’s rank correlation coefficient. The confusion matrix on the right shows the comparison between our ECR-predicted farmland sizes and the manually interpreted labeling results.</p>
Full article ">
19 pages, 5213 KiB  
Article
Revealing the Nonlinear Impact of Human Activities and Climate Change on Ecosystem Services in the Karst Region of Southeastern Yunnan Using the XGBoost–SHAP Model
by Bao Zhou, Guoping Chen, Haoran Yu, Junsan Zhao and Ying Yin
Forests 2024, 15(8), 1420; https://doi.org/10.3390/f15081420 - 13 Aug 2024
Viewed by 206
Abstract
The Karst region is a critical ecological barrier and functional zone in China. Understanding the spatiotemporal evolution of its ecosystem services and its relationship with human activities and climate change is of importance for achieving regional ecological protection and high-quality development. In this [...] Read more.
The Karst region is a critical ecological barrier and functional zone in China. Understanding the spatiotemporal evolution of its ecosystem services and its relationship with human activities and climate change is of importance for achieving regional ecological protection and high-quality development. In this study, we used the InVEST model and CASA model to evaluate the spatiotemporal evolution pattern of ecosystem services in the study area from 2000 to 2020. The XGBoost–SHAP model was used to reveal the key indicators and thresholds of changes in major ecosystem services in the study area due to climate change and human activities. The results showed significant land use changes in the study area from 2000 to 2020, particularly the conversion of cropland to construction land, which was more intense in economically developed areas. The areas of forest and grassland increased initially but later decreased due to the impact of human activities and natural factors. Habitat quality (HQ) showed an overall declining trend, while soil retention (SR) and water yield (WY) services exhibited significant interannual variations due to climate change. The changes in rainfall had a particularly notable impact on these services; in years with excessive rainfall, soil erosion intensified, leading to a decline in SR services, whereas in years with moderate rainfall, SR and WY services improved. Carbon fixation (CF) services were enhanced with the expansion of forest areas. The XGBoost–SHAP model further revealed that the effects of rainfall and sunshine duration on ecosystem services were nonlinear, while population density and the proportion of construction land had a significant negative impact on habitat quality and soil retention. The expansion of construction land had the most significant negative impact on habitat quality, whereas the increase in forest land significantly improved carbon fixation and the soil retention capacity. By revealing the mechanisms of the impact of climate change and human activities on ecosystem services, we aimed to provide support for the promotion of ecological conservation and sustainable development strategies in the study area, as well as to provide an important reference for areas with geographic similarities to the study area. Full article
(This article belongs to the Special Issue Application of Remote Sensing in Vegetation Dynamic and Ecology)
Show Figures

Figure 1

Figure 1
<p>Topography and land use of the study area (2000 and 2020).</p>
Full article ">Figure 2
<p>XGBoost model prediction accuracy and error.</p>
Full article ">Figure 3
<p>Sankey diagram of land use transfer in the study area.</p>
Full article ">Figure 4
<p>Spatial distribution map of ecosystem services in the study area from 2000 to 2020.</p>
Full article ">Figure 5
<p>Trade-off synergies and trends in ecosystem services in the study area between 2000 and 2020.</p>
Full article ">Figure 6
<p>Bivariate spatial autocorrelation diagram of ecosystem services in the study area from 2000 to 2020.</p>
Full article ">Figure 7
<p>Average SHAP values and factor importance rankings in the study area shape the habitat pattern. The main influencing factors of NPP were the NDVI, sunshine duration, rainfall, proportion of forested land, and slope. The SHAP value of the NDVI was the highest (0.05), which indicated that the vegetation cover and health were the key factors determining the NPP, and emphasizing the direct influence of vegetation on the productivity of ecosystems. Sunshine duration (SHAP value of 0.02) and rainfall (SHAP value of 0.01) indicated the direct influence of climatic factors on vegetation growth and ecological productivity. The main factors affecting soil retention were slope, rainfall, the NDVI, percentage of forested land, and temperature (Tem). Slope, with a SHAP value of 0.03, was the most important topographic factor affecting soil retention, pointing to the potential risk of steep topography to soil erosion. Rainfall had a SHAP value of 0.02, emphasizing the dual role of precipitation in soil erosion and conservation. Water yield was influenced by rainfall, humidity (Hum), agricultural land share, sunshine hours, and elevation. Rainfall had the highest SHAP value (0.09), showing the dominance of rainfall in determining regional water availability. The SHAP values of 0.01 for both humidity and elevation further illustrated the influence of climatic and topographic conditions on the water resource distribution.</p>
Full article ">Figure 7 Cont.
<p>Average SHAP values and factor importance rankings in the study area shape the habitat pattern. The main influencing factors of NPP were the NDVI, sunshine duration, rainfall, proportion of forested land, and slope. The SHAP value of the NDVI was the highest (0.05), which indicated that the vegetation cover and health were the key factors determining the NPP, and emphasizing the direct influence of vegetation on the productivity of ecosystems. Sunshine duration (SHAP value of 0.02) and rainfall (SHAP value of 0.01) indicated the direct influence of climatic factors on vegetation growth and ecological productivity. The main factors affecting soil retention were slope, rainfall, the NDVI, percentage of forested land, and temperature (Tem). Slope, with a SHAP value of 0.03, was the most important topographic factor affecting soil retention, pointing to the potential risk of steep topography to soil erosion. Rainfall had a SHAP value of 0.02, emphasizing the dual role of precipitation in soil erosion and conservation. Water yield was influenced by rainfall, humidity (Hum), agricultural land share, sunshine hours, and elevation. Rainfall had the highest SHAP value (0.09), showing the dominance of rainfall in determining regional water availability. The SHAP values of 0.01 for both humidity and elevation further illustrated the influence of climatic and topographic conditions on the water resource distribution.</p>
Full article ">Figure 8
<p>Interactions between key variables and ecosystem services.</p>
Full article ">
28 pages, 9121 KiB  
Article
Flood Hazard and Risk Assessment of Flash Floods for Petra Catchment Area Using Hydrological and Analytical Hierarchy (AHP) Modeling
by Mustafa Al Kuisi, Naheel Al Azzam, Tasneem Hyarat and Ibrahim Farhan
Water 2024, 16(16), 2283; https://doi.org/10.3390/w16162283 - 13 Aug 2024
Viewed by 363
Abstract
Floods are a widespread natural disaster that occur in most areas of the world, except for the polar regions. To minimize the damage caused by floods, effective management strategies and policies must be implemented. Petra and Wadi Musa areas are prone to floods, [...] Read more.
Floods are a widespread natural disaster that occur in most areas of the world, except for the polar regions. To minimize the damage caused by floods, effective management strategies and policies must be implemented. Petra and Wadi Musa areas are prone to floods, which happen every 2–3 years and result in significant harm to both lives and properties. To address this issue, a composite hazard and vulnerability index is commonly utilized to evaluate flood risk and guide policy formation for flood risk reduction. These tools are efficient and cost-effective in generating accurate results. Accordingly, the present study aims to determine the morphological and hydrometeorological parameters that affect flash floods in Petra catchment area and to identify high-risk zones using GIS, hydrological, and analytical hierarchy (AHP) modeling. Nine factors, including Elevation (E), Landuse/Landcover LULC, Slope (S), Drainage density (DD), Flood Control Points (FCP) and Rainfall intensity (RI), which make up the six risk indices, and Population Density (PD), Cropland (C), and Transportation (Tr), which make up the three vulnerability indices, were evaluated both individually and in combination using AHP in ArcGIS 10.8.2 software. These parameters were classified as hazard and vulnerability indicators, and a final flood map was generated. The map indicated that approximately 37% of the total area in Petra catchment is at high or very high risk of flooding, necessitating significant attention from governmental agencies and decision-makers for flood risk mitigation. The AHP method proposed in this study is an accurate tool for flood mapping that can be easily applied to other regions in Jordan to manage and prevent flood hazards. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area.</p>
Full article ">Figure 2
<p>(<b>a</b>). Geological map, (<b>b</b>). soil texture, (<b>c</b>). soil hydrological group, (<b>d</b>). DEM (<b>e</b>). slope and (<b>f</b>). land use.</p>
Full article ">Figure 3
<p>Flowchart showing the methodology for this study.</p>
Full article ">Figure 4
<p>Sub-catchments of the study area.</p>
Full article ">Figure 5
<p>Annual rainfall (mm) with the gauge stations.</p>
Full article ">Figure 6
<p>Long-term annual rainfall of (<b>a</b>) Wadi Musa and (<b>b</b>) Petra rainfall gauging stations with a nine-year moving average.</p>
Full article ">Figure 7
<p>IDF curves: (<b>a</b>) Wadi Musa and (<b>b</b>) Petra rain gauging stations.</p>
Full article ">Figure 8
<p>CN distribution value for the Petra catchment.</p>
Full article ">Figure 9
<p>The flood inflow hydrographs of the Petra catchment outlet created by HEC-1 for the different return periods (5, 10, 25, 50, 100 and 1000 years) with rainfall intensities of 10, 30, 60 and 180 min and 24 h.</p>
Full article ">Figure 10
<p>Perpendicular cross sections and water depth along the Wadi course.</p>
Full article ">Figure 11
<p>Flood inundation map with 1 and 24 h’ rainfall intensity for the return periods of 10, 25, 50, 100, and 1000 years.</p>
Full article ">Figure 12
<p>The thematic standardized maps for the hazard and vulnerability indicators, (<b>a</b>). Rainfall Intensities, (<b>b</b>). Elevation, (<b>c</b>). Slope, (<b>d</b>). Flood Control Points, (<b>e</b>). Drainage Density, (<b>f</b>). Land Use/Land Cover, (<b>g</b>). Cropland, (<b>h</b>). Transportation, and (<b>i</b>). Population Density.</p>
Full article ">Figure 13
<p>Flood hazard, vulnerability and risk maps.</p>
Full article ">
20 pages, 8914 KiB  
Article
Driving Forces behind the Reduction in Cropland Area on Hainan Island, China: Implications for Sustainable Agricultural Development
by Jianchao Guo, Shi Qi, Jiadong Chen and Jinlin Lai
Land 2024, 13(8), 1274; https://doi.org/10.3390/land13081274 - 13 Aug 2024
Viewed by 230
Abstract
Food security is a major challenge for China at present and will be in the future. Revealing the spatiotemporal changes in cropland and identifying their driving forces would be helpful for decision-making to maintain grain supply and sustainable development. Hainan Island is endowed [...] Read more.
Food security is a major challenge for China at present and will be in the future. Revealing the spatiotemporal changes in cropland and identifying their driving forces would be helpful for decision-making to maintain grain supply and sustainable development. Hainan Island is endowed with rich agricultural resources due to its unique climatic conditions and is facing tremendous pressure in cropland protection due to the huge variation in natural conditions and human activities over the past few decades. The purpose of this study is to assess the spatiotemporal changes in and driving forces of cropland on Hainan Island in the past and predict future cropland changes under different scenarios. Key findings are as follows: (1) From 2000 to 2020, the cropland area on Hainan Island decreased by 956.22 km2, causing the center of cropland to shift southwestward by 8.20 km. This reduction mainly transformed into construction land and woodland, particularly evident in coastal areas. (2) Among anthropogenic factors, the increase in the human footprint is the primary reason for the decrease in cropland. Land use changes driven by population growth, especially in economically active and densely populated coastal areas, are key factors in this decrease. Natural factors such as topography and climate change also significantly impact cropland changes. (3) Future scenarios show significant differences in cropland area changes. In the natural development scenario, the cropland area is expected to continue decreasing to 597 km2, while in the ecological protection scenario, cropland conversion is restricted to 269.11 km2; however, in the cropland protection scenario, the trend of cropland reduction is reversed, increasing by 448.75 km2. Our findings provide a deep understanding of the driving forces behind cropland changes and, through future scenario analysis, demonstrate the potential changes in cropland area under different policy choices. These insights are crucial for formulating sound land management and agricultural policies to protect cropland resources, maintain food security, and promote ecological balance. Full article
(This article belongs to the Section Land – Observation and Monitoring)
Show Figures

Figure 1

Figure 1
<p>The geographic location of the study area.</p>
Full article ">Figure 2
<p>Spatial distribution of land use types.</p>
Full article ">Figure 3
<p>Cropland transformation from 2000 to 2020.</p>
Full article ">Figure 4
<p>Land transfer Sankey diagram.</p>
Full article ">Figure 5
<p>Spatial distribution of center-of-gravity and standard-deviation ellipses of cropland.</p>
Full article ">Figure 6
<p>Spatial distribution of cropland KD in (<b>a</b>) 2000, (<b>b</b>) 2010, and (<b>c</b>) 2020 and (<b>d</b>) change from 2000 to 2020.</p>
Full article ">Figure 7
<p>Feature variable importance (<b>a</b>) and SHAP summary (<b>b</b>). Note: Ele, Slop, Clay, Sand, Rd, and Wd are static variables, representing elevation, slope, clay content, sand content, road distance, and water distance, respectively. Tem, Pre, Hf, and Pop are dynamic variables, representing changes in temperature, precipitation, human footprint, and population density, respectively.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>j</b>) Correlation between KDC and feature variables.</p>
Full article ">Figure 9
<p>(<b>a</b>–<b>o</b>) Interactions between feature variables.</p>
Full article ">Figure 10
<p>Land use in 2040 on Hainan Island.</p>
Full article ">
22 pages, 11626 KiB  
Article
Assessment of Semi-Automated Techniques for Crop Mapping in Chile Based on Global Land Cover Satellite Data
by Matías Volke, María Pedreros-Guarda, Karen Escalona, Eduardo Acuña and Raúl Orrego
Remote Sens. 2024, 16(16), 2964; https://doi.org/10.3390/rs16162964 - 12 Aug 2024
Viewed by 372
Abstract
In recent years, the Chilean agricultural sector has undergone significant changes, but there is a lack of data that can be used to accurately identify these transformations. A study was conducted to assess the effectiveness of different spatial resolutions used by global land [...] Read more.
In recent years, the Chilean agricultural sector has undergone significant changes, but there is a lack of data that can be used to accurately identify these transformations. A study was conducted to assess the effectiveness of different spatial resolutions used by global land cover products (MODIS, ESA and Dynamic World (DW)), in addition to the demi-automated methods applied to them, for the identification of agricultural areas, using the publicly available agricultural survey for 2021. It was found that lower-spatial-resolution collections consistently underestimated crop areas, while collections with higher spatial resolutions overestimated them. The low-spatial-resolution collection, MODIS, underestimated cropland by 46% in 2021, while moderate-resolution collections, such as ESA and DW, overestimated cropland by 39.1% and 93.8%, respectively. Overall, edge-pixel-filtering and a machine learning semi-automated reclassification methodology improved the accuracy of the original global collections, with differences of only 11% when using the DW collection. While there are limitations in certain regions, the use of global land cover collections and filtering methods as training samples can be valuable in areas where high-resolution data are lacking. Future research should focus on validating and adapting these approaches to ensure their effectiveness in sustainable agriculture and ecosystem conservation on a global scale. Full article
(This article belongs to the Special Issue Application of Satellite and UAV Data in Precision Agriculture)
Show Figures

Figure 1

Figure 1
<p>Study area. On the left, Chile with the Ñuble region is marked in black. On the right, a zoomed image of the Ñuble region and the names of its communes are shown.</p>
Full article ">Figure 2
<p>Flow diagram representing the methodology of this work.</p>
Full article ">Figure 3
<p>Agricultural land area in km<sup>2</sup>, (<b>a</b>) per commune in the Ñuble region and (<b>b</b>) the total in the region. Year: 2021. The calculations were retrieved from the following databases: an agricultural survey (AS), MODIS, ESA, Dynamic World (DW) and improved versions of the latter three (v2).</p>
Full article ">Figure 4
<p>Agricultural area in km<sup>2</sup> per commune in the Ñuble region. (<b>a</b>) Agricultural survey and (<b>b</b>) original ESA dataset, as the most precise original dataset from the three tested. (<b>c</b>) Filtered and reclassified DW, that is, DW version 2 (v2), as the most accurate filtered and reclassified dataset.</p>
Full article ">Figure 4 Cont.
<p>Agricultural area in km<sup>2</sup> per commune in the Ñuble region. (<b>a</b>) Agricultural survey and (<b>b</b>) original ESA dataset, as the most precise original dataset from the three tested. (<b>c</b>) Filtered and reclassified DW, that is, DW version 2 (v2), as the most accurate filtered and reclassified dataset.</p>
Full article ">Figure 5
<p>Zoomed image of the pixel reduction through different stages of quality filtering of <a href="#remotesensing-16-02964-f002" class="html-fig">Figure 2</a>.</p>
Full article ">Figure A1
<p>Maps of agricultural mask retrieved from MODIS dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A2
<p>Maps of agricultural mask retrieved from DW dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A3
<p>Maps of agricultural mask retrieved from ESA dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A3 Cont.
<p>Maps of agricultural mask retrieved from ESA dataset (details in <a href="#remotesensing-16-02964-t002" class="html-table">Table 2</a>) for the year 2021, (<b>a</b>) directly and (<b>b</b>) from the preprocessing described in <a href="#sec3dot2dot2-remotesensing-16-02964" class="html-sec">Section 3.2.2</a>.</p>
Full article ">Figure A4
<p>Maps of agricultural mask retrieved from CONAF dataset for the year 2021.</p>
Full article ">
11 pages, 2193 KiB  
Perspective
A Review of the Application and Impact of Drip Irrigation under Plastic Mulch in Agricultural Ecosystems
by Chunyu Wang, Sien Li, Siyu Huang and Xuemin Feng
Agronomy 2024, 14(8), 1752; https://doi.org/10.3390/agronomy14081752 - 10 Aug 2024
Viewed by 412
Abstract
Food security, a crucial issue for the development of humankind, is often severely constrained by water scarcity. As a globally recognized most advanced agricultural water-saving technology, drip irrigation under plastic mulch (DIPM) has played a significant role in grain production. However, a comprehensive [...] Read more.
Food security, a crucial issue for the development of humankind, is often severely constrained by water scarcity. As a globally recognized most advanced agricultural water-saving technology, drip irrigation under plastic mulch (DIPM) has played a significant role in grain production. However, a comprehensive review of the dual impacts of this practice in farmland remains lacking. This study has conducted an exhaustive review of DIPM research from 1999 to 2023 and employed CiteSpace software to perform a co-occurrence and clustering analysis of keywords in order to reveal research hotspots and trends. The results show that the attention to DIPM technology has increased annually and reached a peak in 2022. China leads in the number of publications in this field, reflecting its emphasis on agricultural water-saving technologies. This study critically discusses the dual impacts of DIPM on farmland. On the positive side, DIPM can improve soil temperature and moisture, enhance nutrient availability, promote water and nutrient absorption by roots, and increase the crop growth rate and yield while reducing evaporation and nitrogen loss, suppressing weed growth, decreasing herbicide usage, and lowering total greenhouse gas emissions. On the negative side, it will cause pollution from plastic mulch residues, damage the soil structure, have impacts on crop growth, and lead to increased clogging of drip irrigation systems, which will increase agricultural costs and energy consumption, hinder crop growth, hamper soil salinization management, and further reduce the groundwater level. The future development of DIPM technology requires optimization and advancement. Such strategies as mechanized residual-mulch recovery, biodegradable mulch substitution, aerated drip irrigation technology, and alternate irrigation are proposed to address existing issues in farmland triggered by DIPM. This review advocates for the active exploration of farming management practices superior to DIPM for future agricultural development. These practices could lead to higher yields, water–nitrogen efficiency, and lower environmental impact in agricultural development. Full article
(This article belongs to the Section Water Use and Irrigation)
Show Figures

Figure 1

Figure 1
<p>Distribution of articles pertaining to drip irrigation under plastic mulch (DIPM) by year and by major journals. (<b>a</b>) represents the number of articles published in different years. (<b>b</b>) represents the specific number of articles in journals with more than 25 articles.</p>
Full article ">Figure 2
<p>The distribution of articles related to DIPM across various countries around the world.</p>
Full article ">Figure 3
<p>Keyword co-occurrence (red) and clustering (blue) map displaying the most frequently used words associated with DIPM research.</p>
Full article ">
13 pages, 2403 KiB  
Article
The Characteristics and Influential Factors of Earthworm and Vermicompost under Different Land Use in a Temperate Area, China
by Li Ma, Ming’an Shao, Yunqiang Wang, Tongchuan Li, Xuanxuan Jing, Kunyu Jia and Yangyang Zhang
Forests 2024, 15(8), 1389; https://doi.org/10.3390/f15081389 - 8 Aug 2024
Viewed by 419
Abstract
Earthworm communities influence soil carbon and nitrogen circulation by altering the diversity and composition of microbial communities, which improves soil fertility. Studying the soil nutrient composition and bacterial communities change in response to earthworm community natural invasion may be key to exploring earthworm [...] Read more.
Earthworm communities influence soil carbon and nitrogen circulation by altering the diversity and composition of microbial communities, which improves soil fertility. Studying the soil nutrient composition and bacterial communities change in response to earthworm community natural invasion may be key to exploring earthworm ecological functions and accurately assessing C and N mineralization in artificial forests and croplands. In this study, we examined the communities of five earthworm species in ecosystems characterized by six different land-use types, such as buttonwood forest, walnut forest, apple orchard, kiwi orchard, ryegrass land, and corn field. The Metaphire baojiensis (d) and Amynthas carnosus planus were dominant earthworm species. Among different land-use types, earthworm densities ranged from 2 to 27 ind·m−2 in summer and 15 to 40 ind·m−2 in spring. However, surface vermicompost weight in summer (296.7 to 766.0 g·m−2) was greater than in spring. There was a positive correlation between the weight of the vermicompost and earthworm numbers in the same season. Soil carbon (C) and total nitrogen (N) of vermicompost ranged from 5.12 to 20.93 g·kg−1 and from 0.52 to 1.35 g·kg−1, respectively. Compared with soil, the contents of vermicompost C and N increased 2.0 to 4.3 times and 1.6 to 7.7 times, respectively. The average C/N of vermicompost (9.5~23.5) was higher than in the soil (7.3~19.8). Due to the higher abundances of C and N in the soil of corn fields and kiwi orchards, which cultivate higher abundances of earthworms and more vermicompost, the C and N and C/N of vermicompost is higher than in the soil. C and N were accumulated by earthworms’ excreting and feeding activity instead of vegetation in vermicompost. Earthworm community structure plays key roles in decreasing bacterial diversity and adding Proteobacteria, Actinobacteria, Acidobacteria, Bacteroidetes, and Chloroflex in vermicompost, resulting in enriching soil C and N content and increasing C/N in vermicompost. Therefore, the evaluation of different vegetation ecosystems in soil C and N pool accumulation and mineralization should be given more attention regarding the function of earthworm communities in the future. Full article
(This article belongs to the Section Forest Soil)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) The soil organic (<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>), (<b>b</b>) total nitrogen (<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>) content, and (<b>c</b>) C/N (<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>) in vermicompost and soil under different land-use types. (The soil in different land-use types: buttonwood forest (BS), apple orchard (AS), walnut forest (WS), ryegrass land (RS), kiwi orchard (KS), and corn field (CS); the vermicompost in different land-use types: buttonwood forest (BV), apple orchard (AV), walnut forest (WV), ryegrass land (RV), kiwi orchard (KV), and corn field (CV)). (Different lowercase letters indicate significant differences in treatment at the 0.05 level). Blue: soil organic of vermicompost; green: total nitrogen of vermicompost; orange: C/N of vermicompost.</p>
Full article ">Figure 1 Cont.
<p>(<b>a</b>) The soil organic (<b>a<sub>1</sub></b>,<b>a<sub>2</sub></b>), (<b>b</b>) total nitrogen (<b>b<sub>1</sub></b>,<b>b<sub>2</sub></b>) content, and (<b>c</b>) C/N (<b>c<sub>1</sub></b>,<b>c<sub>2</sub></b>) in vermicompost and soil under different land-use types. (The soil in different land-use types: buttonwood forest (BS), apple orchard (AS), walnut forest (WS), ryegrass land (RS), kiwi orchard (KS), and corn field (CS); the vermicompost in different land-use types: buttonwood forest (BV), apple orchard (AV), walnut forest (WV), ryegrass land (RV), kiwi orchard (KV), and corn field (CV)). (Different lowercase letters indicate significant differences in treatment at the 0.05 level). Blue: soil organic of vermicompost; green: total nitrogen of vermicompost; orange: C/N of vermicompost.</p>
Full article ">Figure 2
<p>The characteristics of bacterial classification in soil (<b>a</b>) and vermicompost (<b>b</b>) under different land-use types. (Soil: buttonwood forest (BS1, BS2, BS3), walnut forest (WS1, WS2, WS3), apple orchard (AS1, AS2, AS3), kiwi orchard (KS1, KS2, KS3), ryegrass land (RS1, RS2, RS3), and corn field (CS1, CS2, CS3). Vermicompost: buttonwood forest (BV1, BV2, BV3), walnut forest (WV1, WV2, WV3), apple orchard (AV1, AV2, AV3), kiwi orchard (KV1, KV2, KV3), ryegrass land (RV1, RV2, RV3), and corn field (CV1, CV2, CV3)).</p>
Full article ">Figure 3
<p>RDA (redundancy analysis) map of relationship between earthworm population density, environmental factors, and bacterial community in earthworm compost. ((<b>a</b>): in soil; (<b>b</b>): in vermicompost). (Soil: buttonwood forest (BS), walnut forest (WS), apple orchard (AS), kiwi orchard (KS), ryegrass land (RS), and corn field (CS). Vermicompost: buttonwood forest (BV), walnut forest (WV), apple orchard (AV), kiwi orchard (KV), ryegrass land (RV), and corn field (CV)). The red line were bacterial communities. The blue lines were the environmental factors.</p>
Full article ">
20 pages, 15204 KiB  
Article
Optimizing Land Use to Mitigate Ecosystem Service Trade-Offs Using Multi-Scenario Simulation in the Luo River Basin
by Yulong Dai, Xuning Qiao, Yongju Yang, Liang Liu, Yuru Chen, Jing Zhang and Tongqian Zhao
Land 2024, 13(8), 1243; https://doi.org/10.3390/land13081243 - 8 Aug 2024
Viewed by 206
Abstract
For a long time in the past, China has implemented a large number of “Grain for Green” projects (GFGPs) to improve the ecological environment. However, it is still unclear whether excessive GFGPs will exacerbate the trade-off of ecosystem services (ESs). Additionally, it is [...] Read more.
For a long time in the past, China has implemented a large number of “Grain for Green” projects (GFGPs) to improve the ecological environment. However, it is still unclear whether excessive GFGPs will exacerbate the trade-off of ecosystem services (ESs). Additionally, it is a great challenge to explore the response mechanism of the trade-off relationship to changes in land use and to mitigate the trade-offs by optimizing land use. Taking a typical GFGP basin in the central Yellow River basin as an example, we identified the trade-off areas and measured the nonlinear trade-offs between ESs under different scenarios. This was carried out based on the synergistic potential of the production possibility frontier (PPF) and the first-order derivative. We also identified the optimal scenario for mitigating the trade-offs of ESs. The results showed that excessive GFGPs have intensified the ES trade-offs. The differences in land use types lead to spatial heterogeneity in the relationship of ESs. When carbon storage (CS) is 9.58 t/km2 and habitat quality (HQ) is 0.4, the relationship with water yield (WY) changes from trade-off to synergy, respectively, and the trade-off area is mainly distributed in cropland and construction land. Compared with 2020, the EP scenario has the highest synergy potential and the lowest trade-off intensity, and can alleviate the ES trade-off to the greatest extent. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>Research framework.</p>
Full article ">Figure 3
<p>Illustration of scenario settings.</p>
Full article ">Figure 4
<p>Spatial and temporal changes in the level of supply of the four ESs.</p>
Full article ">Figure 5
<p>Correlation analysis of ESs.</p>
Full article ">Figure 6
<p>Identify ES trade-off area based on production possibility frontier first-order derivative.</p>
Full article ">Figure 7
<p>Production possibility frontier and first-order derivative.</p>
Full article ">Figure 8
<p>Land flow maps for the four scenarios from 2020 to 2030, and the level of ES supply in 2030.</p>
Full article ">Figure 9
<p>Production-possibility frontiers and their first-order derivatives for WY and CS as well as WY and HQ. Scenarios I, II, III, and IV represent the ND, CP, EP, and CD scenarios.</p>
Full article ">
21 pages, 6277 KiB  
Article
Impacts of Drought and Heatwave on the Vegetation and Ecosystem in the Yangtze River Basin in 2022
by Siyuan Chen, Ruonan Qiu, Yumin Chen, Wei Gong and Ge Han
Remote Sens. 2024, 16(16), 2889; https://doi.org/10.3390/rs16162889 - 7 Aug 2024
Viewed by 481
Abstract
In 2022, a severe drought and heatwave occurred in the middle and lower reaches of the Yangtze River Basin. Previous studies have highlighted the severity of this event, yet the relevance of soil moisture (SM), as well as vapor pressure deficit (VPD) and [...] Read more.
In 2022, a severe drought and heatwave occurred in the middle and lower reaches of the Yangtze River Basin. Previous studies have highlighted the severity of this event, yet the relevance of soil moisture (SM), as well as vapor pressure deficit (VPD) and vegetation damage, remained unclear. Here, we utilized solar-induced chlorophyll fluorescence (SIF) and various flux data to monitor the impact of drought on vegetation and analyze the influence of different environmental factors. The results indicated a severe situation of drought and heatwave in the Yangtze River Basin in 2022 that significantly affected vegetation growth and the ecosystem carbon balance. SIF and NDVI have respective advantages in reflecting damage to vegetation under drought and heatwave conditions; SIF is more capable of capturing the weakening of vegetation photosynthesis, while NDVI can more rapidly indicate vegetation damage. Additionally, the correlation of SM and SIF are comparable to that of VPD and SIF. By contrast, the differentiation in the severity of vegetation damage among different types of vegetation is evident; cropland is more vulnerable compared to forest ecosystems and is more severely affected by drought and heatwave. These findings provided important insights for assessing the impact of compound drought and heatwave events on vegetation growth. Full article
(This article belongs to the Special Issue Remote Sensing Application in the Carbon Flux Modelling)
Show Figures

Figure 1

Figure 1
<p>The Yangtze River Basin. The dark gray lines represent the primary and secondary water systems of the Yangtze River. The base map depicts the land cover type based on data processed from 2019 MCD12Q1 v006. Forest is a collective term encompassing evergreen coniferous forests, evergreen broad-leaved forests, deciduous broad-leaved forests, and mixed forests.</p>
Full article ">Figure 2
<p>Spatial distributions of temperature, precipitation, and VPD anomalies in the Yangtze River Basin during July ((<b>a</b>,<b>e</b>,<b>i</b>) in 1st row), August ((<b>b</b>,<b>f</b>,<b>j</b>) in 2nd row), September ((<b>c</b>,<b>g</b>,<b>k</b>) in 3rd row), and October ((<b>d</b>,<b>h</b>,<b>l</b>) in 4th row) in 2022.</p>
Full article ">Figure 3
<p>Spatial distributions of SM1, SM2, SM3, and SM4 anomalies in the Yangtze River Basin during July ((<b>a</b>,<b>e</b>,<b>i</b>,<b>m</b>) in 1st row), August ((<b>b</b>,<b>f</b>,<b>j</b>,<b>n</b>) in 2nd row), September ((<b>c</b>,<b>g</b>,<b>k</b>,<b>o</b>) in 3rd row), and October ((<b>d</b>,<b>h</b>,<b>l</b>,<b>p</b>) in 4th row) in 2022.</p>
Full article ">Figure 4
<p>The seasonal cycles of different environmental metrics: (<b>a</b>) temperature (κ), (<b>c</b>) precipitation (mm), and (<b>e</b>) VPD (hPa). The corresponding change percentages for the year 2022 compared to the average values from 2018 to 2021 are shown in (<b>b</b>) temperature, (<b>d</b>) precipitation, and (<b>f</b>) VPD. In (<b>a</b>,<b>c</b>,<b>e</b>), the black line represents the monthly average for 2022, and the red line represents the monthly average value from 2018 to 2021. In (<b>b</b>,<b>d</b>,<b>f</b>) the blue part represents the proportion of value in 2022 that exceeds the average from 2018 to 2021, while the red part represents the proportion of value in 2022 that falls below the average from 2018 to 2021.</p>
Full article ">Figure 5
<p>The seasonal cycles of different soil moisture metrics: (<b>a</b>) SM1 (unitless), (<b>c</b>) SM2 (unitless), (<b>e</b>) SM3 (unitless), and (<b>g</b>) SM4 (unitless). The corresponding change percentages for the year 2022 compared to the average values from 2018 to 2021 are shown in (<b>b</b>) SM1 (unitless), (<b>d</b>) SM2 (unitless), (<b>f</b>) SM3 (unitless), and (<b>h</b>) SM4 (unitless). In (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>), the black line represents the monthly average for 2022, and the red line represents the monthly average value from 2018 to 2021. In (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>), the blue part represents the proportion of value in 2022 that exceeds the average from 2018 to 2021, while the red part represents the proportion of value in 2022 that falls below the average from 2018 to 2021.</p>
Full article ">Figure 6
<p>The seasonal cycles of different soil moisture metrics: (<b>a</b>) NDVI (unitless), (<b>c</b>) SIF (unitless), (<b>e</b>) GPP (gC m<sup>−2</sup> d<sup>−1</sup>), and (<b>g</b>) NEE (gC m<sup>−2</sup> d<sup>−1</sup>). The corresponding change percentages for the year 2022 compared to the average values from 2018 to 2021 are shown in (<b>b</b>) NDVI (unitless), (<b>d</b>) SIF (unitless), (<b>f</b>) GPP (gC m<sup>−2</sup> d<sup>−1</sup>), and (<b>h</b>) NEE (gC m<sup>−2</sup> d<sup>−1</sup>). In (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>), the black line represents the monthly average for 2022, and the red line represents the monthly average value from 2018 to 2021. In (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>), the blue part represents the proportion of value in 2022 that exceeds the average from 2018 to 2021, while the red part represents the proportion of value in 2022 that falls below the average from 2018 to 2021.</p>
Full article ">Figure 7
<p>Spatial distributions of NDVI, SIF, and GPP anomalies in the Yangtze River Basin during July ((<b>a</b>,<b>e</b>,<b>i</b>,<b>m</b>) in 1st row), August ((<b>b</b>,<b>f</b>,<b>j</b>,<b>n</b>) in 2nd row), September ((<b>c</b>,<b>g</b>,<b>k</b>,<b>o</b>) in 3rd row), and October ((<b>d</b>,<b>h</b>,<b>l</b>,<b>p</b>) in 4th row) in 2022.</p>
Full article ">Figure 8
<p>Spatial distribution of partial correlations between July and October 2022: (<b>a</b>) correlations between SM1 anomalies and SIF anomalies, (<b>b</b>) correlations between VPD anomalies and SIF anomalies. The blue part represents a positive correlation, while the red part represents a negative correlation.</p>
Full article ">Figure 9
<p>The differential values of the seasonal cycles for the year 2022 compared to the average values from 2018 to 2021: (<b>a</b>) temperature(κ), (<b>b</b>) precipitation(mm), and (<b>c</b>–<b>f</b>) SM1–SM4 (unitless). The green line represents forest while the purple line represents cropland.</p>
Full article ">Figure 10
<p>The differential values of the seasonal cycles for the year 2022 compared to the average values from 2018 to 2021: (<b>a</b>) NDVI (unitless), (<b>b</b>) SIF (unitless), (<b>c</b>) GPP (gC m<sup>−2</sup> d<sup>−1</sup>), and (<b>d</b>) NEE (gC m<sup>−2</sup> d<sup>−1</sup>). The green line represents forest, while the purple line represents cropland.</p>
Full article ">
19 pages, 7143 KiB  
Article
Potential Reduction of Spatiotemporal Patterns of Water and Wind Erosion with Conservation Tillage in Northeast China
by Fahui Jiang, Xinhua Peng, Qinglin Li, Yongqi Qian and Zhongbin Zhang
Land 2024, 13(8), 1219; https://doi.org/10.3390/land13081219 - 6 Aug 2024
Viewed by 405
Abstract
Conservational tillage (NT) is widely recognized globally for its efficacy in mitigating soil loss due to wind and water erosion. However, a systematic large-scale estimate of NT’s impact on soil loss reduction in Northeast, China’s primary granary, remains absent. This study aimed to [...] Read more.
Conservational tillage (NT) is widely recognized globally for its efficacy in mitigating soil loss due to wind and water erosion. However, a systematic large-scale estimate of NT’s impact on soil loss reduction in Northeast, China’s primary granary, remains absent. This study aimed to investigate the spatial and temporal variability of soil erosion under NT compared to conventional tillage (CT) in the black soil region and to analyze the underlying mechanisms driving these erosions. The Revised Universal Soil Loss Equation (RUSLE) and the Revised Wind Erosion Equation (RWEQ) models were employed, incorporating previously published plot/watershed data to estimate the potential reduction of water and wind erosion by NT in this region. Results indicated that under CT practices, water- and wind-induced soil losses were widely distributed in the arable land of Northeast China, with intensities of 2603 t km−2 a−1 and 34 t km−2 a−1, respectively. Furthermore, the erosive processes of water and wind erosion were significantly reduced by 56.4% and 91.8%, respectively, under NT practices compared to CT. The highest efficiency in soil conservation using NT was observed in the mountainous regions such as the Changbai Mountains and Greater Khingan Mountains, where water erosion was primarily driven by cropland slopes and wind erosion was driven by the wind speed. Conversely, the largest areas of severe erosion were observed in the Songnen Plain, primarily due to the significant proportion of arable land in this region. In the plain regions, water-induced soil loss was primarily influenced by precipitation, with light and higher levels of erosion occurring more frequently on long gentle slopes (0–3°) than on higher slope areas (3–5°). In the temporal dimension, soil loss induced by water and wind erosion ceased during the winter under both tillage systems due to snow cover and water freezing in the soil combined with the extremely cold climate. Substantial reductions were observed under NT from spring to autumn compared to CT. Ultimately, the temporal and spatial variations of soil loss under CT and NT practices were established from 2010 to 2018 and then projected onto a cropland map of Northeast China. Based on this analysis, NT is recommended as most suitable practice in the southern regions of Northeast China for maintaining soil health and crop yield production, while its suitability decreases in the northern and eastern regions. Full article
(This article belongs to the Topic Slope Erosion Monitoring and Anti-erosion)
Show Figures

Figure 1

Figure 1
<p>The distribution of dryland and the five ecological regions (I–V) in Northeast China.</p>
Full article ">Figure 2
<p>Spatial distribution of factors in the RUSLE model under different tillage practices in 2018, including rainfall erosivity, R (<b>A</b>); soil erodibility, K (<b>B</b>); slope length and steepness, LS (<b>C</b>); crop covering, C (<b>D</b>); soil frozen factor, F (<b>E</b>); and protected effect of conservation tillage, P (<b>F</b>).</p>
Full article ">Figure 3
<p>Spatial distribution of factors in the RWEQ model under different tillage practices in 2018, including wind erosivity, <span class="html-italic">W<sub>f</sub></span> (<b>A</b>); soil wetness, SW (<b>B</b>); snow cover depth, SD (<b>C</b>); soil erodibility factor, K′ (<b>D</b>); soil curst factor, SCF (<b>E</b>); soil roughness, RN (<b>F</b>); soil frozen factor, F′ (<b>G</b>); and straw protection of conservation tillage, P′ (<b>H</b>).</p>
Full article ">Figure 4
<p>Spatial distribution of the water (<b>A</b>–<b>C</b>) and wind (<b>D</b>–<b>F</b>) erosion under different tillage practices in Northeast China’s dryland in 2018. The five ecological regions are the Greater Khingan Mountains (I), the Songnen Plain (II), the Liao River Plain (III), the Changbai Mountains (IV), and the Sanjiang Plain (V).</p>
Full article ">Figure 5
<p>Areas of water (<b>A</b>) and wind (<b>B</b>) erosion under different tillage practices exhibited several grades in Northeast China in 2018. The six erosion grades include tolerable (0–200 t km<sup>−2</sup> a<sup>−1</sup>), slight (200–2500 t km<sup>−2</sup> a<sup>−1</sup>), moderate (2500–5000 t km<sup>−2</sup> a<sup>−1</sup>), severe (5000–8000 t km<sup>−2</sup> a<sup>−1</sup>), very severe (8000–15, 000 t km<sup>−2</sup> a<sup>−1</sup>), and destructive erosion (&gt;15, 000 t km<sup>−2</sup> a<sup>−1</sup>). The five ecological regions are the Greater Khingan Mountains (I), the Songnen Plain (II), the Liao River Plain (III), the Changbai Mountains (IV), and the Sanjiang Plain (V).</p>
Full article ">Figure 6
<p>Annually change in soil loss via water (<b>A</b>) and wind (<b>B</b>) erosion under different tillage practices in Northeast China’s dryland.</p>
Full article ">Figure 7
<p>Monthly change in soil loss via water (<b>A</b>) and wind (<b>B</b>) erosion under different tillage practices in Northeast China’s dryland.</p>
Full article ">Figure 8
<p>The effect of various factors on soil water (<b>A</b>) and wind (<b>B</b>) erosion under different tillage practices in Northeast China’s dryland in 2018.</p>
Full article ">Figure 9
<p>Relationship between soil water erosion and slope under different tillage practices in Northeast China’s dryland in 2018. The slope of the arable land ranged 0–18° in the areas, with 0–5° representing plain areas and 5–18° representing mountainous regions.</p>
Full article ">
33 pages, 17476 KiB  
Article
Driving Analysis and Multi Scenario Simulation of Ecosystem Carbon Storage Changes Based on the InVEST-PLUS Coupling Model: A Case Study of Jianli City in the Jianghan Plain of China
by Jun Shao, Yuxian Wang, Mingdong Tang and Xinran Hu
Sustainability 2024, 16(16), 6736; https://doi.org/10.3390/su16166736 - 6 Aug 2024
Viewed by 651
Abstract
The carbon storage capacity of terrestrial ecosystems serves as a crucial metric for assessing ecosystem health and their resilience to climate change. By evaluating the effects of land use alterations on this storage, carbon management strategies can be improved, thereby promoting carbon reduction [...] Read more.
The carbon storage capacity of terrestrial ecosystems serves as a crucial metric for assessing ecosystem health and their resilience to climate change. By evaluating the effects of land use alterations on this storage, carbon management strategies can be improved, thereby promoting carbon reduction and sequestration. While county-level cities are pivotal to ecological conservation and high-quality development, they often face developmental challenges. Striking a balance between economic growth and meeting peak carbon emissions and carbon neutrality objectives is particularly challenging. Consequently, there is an urgent need to bolster research into carbon storage management. The study focuses on Jianli City, employing the InVEST model and land use data to examine the response patterns of land use changes and terrestrial system carbon storage from 2000 to 2020. Using the PLUS model, the study simulated the land use and carbon storage in Jianli City for the year 2035 under three scenarios: Natural Development scenario, Urban Expansion scenario, and Ecology and food security scenario. Our findings indicate the following: (1) Between 2000 and 2020, significant shifts in land use were observed in Jianli City. These changes predominantly manifested as the interchange between Cropland and Water areas and the enlargement of impervious surfaces, leading to a decrease of 691,790.27 Mg in carbon storage. (2) Under the proposed scenarios—Natural Development scenario, Urban Expansion scenario, and Ecology and food security scenario—the estimated carbon storage capacities in Jianli City were 39.95 Tg, 39.90 Tg, and 40.14 Tg, respectively. When compared with the 2020 data, all these estimates showed an increase. In essence, our study offers insights into optimizing land use structures from a carbon storage standpoint to ensure stability in Jianli’s carbon storage levels while mitigating the risks associated with carbon fixation. This has profound implications for the harmonious evolution of regional eco-economies. Full article
Show Figures

Figure 1

Figure 1
<p>Study area.</p>
Full article ">Figure 2
<p>Research framework.</p>
Full article ">Figure 3
<p>Land use transfer chord map from 2000 to 2020.</p>
Full article ">Figure 4
<p>Land use transfer sankey map from 2000 to 2020.</p>
Full article ">Figure 5
<p>Land use status from 2000 to 2020.</p>
Full article ">Figure 5 Cont.
<p>Land use status from 2000 to 2020.</p>
Full article ">Figure 6
<p>Expansion probability of each land use type.</p>
Full article ">Figure 7
<p>Land use status under three scenarios.</p>
Full article ">Figure 8
<p>Variations in carbon storage and geo-averaged carbon density of terrestrial systems from 2000 to 2020.</p>
Full article ">Figure 9
<p>Spatial distribution of carbon storage from 2000 to 2020.</p>
Full article ">Figure 10
<p>Carbon stock changes from 2000 to 2020.</p>
Full article ">Figure 10 Cont.
<p>Carbon stock changes from 2000 to 2020.</p>
Full article ">Figure 11
<p>The change in carbon storage and carbon sink economy caused by land use type conversion from 2000 to 2020.</p>
Full article ">Figure 11 Cont.
<p>The change in carbon storage and carbon sink economy caused by land use type conversion from 2000 to 2020.</p>
Full article ">Figure 12
<p>Spatial distribution of carbon storage under three scenarios.</p>
Full article ">Figure 13
<p>The change in carbon storage and carbon sink economy caused by land use type conversion under three scenarios.</p>
Full article ">Figure 13 Cont.
<p>The change in carbon storage and carbon sink economy caused by land use type conversion under three scenarios.</p>
Full article ">Figure 14
<p>Driving factors.</p>
Full article ">Figure 15
<p>Importance of driving factors for each land use type.</p>
Full article ">
15 pages, 2251 KiB  
Review
Soil Health Intensification through Strengthening Soil Structure Improves Soil Carbon Sequestration
by Ryusuke Hatano, Ikabongo Mukumbuta and Mariko Shimizu
Agriculture 2024, 14(8), 1290; https://doi.org/10.3390/agriculture14081290 - 5 Aug 2024
Viewed by 1297
Abstract
Intensifying soil health means managing soils to enable sustainable crop production and improved environmental impact. This paper discusses soil health intensification by reviewing studies on the relationship between soil structure, soil organic matter (SOM), and ecosystem carbon budget. SOM is strongly involved in [...] Read more.
Intensifying soil health means managing soils to enable sustainable crop production and improved environmental impact. This paper discusses soil health intensification by reviewing studies on the relationship between soil structure, soil organic matter (SOM), and ecosystem carbon budget. SOM is strongly involved in the development of soil structure, nutrient and water supply power, and acid buffering power, and is the most fundamental parameter for testing soil health. At the same time, SOM can be both a source and a sink for atmospheric carbon. A comparison of the ratio of soil organic carbon to clay content (SOC/Clay) is used as an indicator of soil structure status for soil health, and it has shown significantly lower values in cropland than in grassland and forest soils. This clearly shows that depletion of SOM leads to degradation of soil structure status. On the other hand, improving soil structure can lead to increasing soil carbon sequestration. Promoting soil carbon sequestration means making the net ecosystem carbon balance (NECB) positive. Furthermore, to mitigate climate change, it is necessary to aim for carbon sequestration that can improve the net greenhouse gas balance (NGB) by serving as a sink for greenhouse gases (GHG). The results of a manure application test in four managed grasslands on Andosols in Japan showed that it was necessary to apply more than 2.5 tC ha−1 y−1 of manure to avoid reduction and loss of SOC in the field. Furthermore, in order to offset the increase in GHG emissions due to N2O emissions from increased manure nitrogen input, it was necessary to apply more than 3.5 tC ha−1y−1 of manure. To intensify soil health, it is increasingly important to consider soil management with organic fertilizers that reduce chemical fertilizers without reducing yields. Full article
(This article belongs to the Special Issue Feature Review in Agricultural Soils—Intensification of Soil Health)
Show Figures

Figure 1

Figure 1
<p>Effects of soil organic matter improvement managements [<a href="#B8-agriculture-14-01290" class="html-bibr">8</a>] on soil carbon sequestration [<a href="#B4-agriculture-14-01290" class="html-bibr">4</a>], nutrient and water supply improvement [<a href="#B32-agriculture-14-01290" class="html-bibr">32</a>,<a href="#B33-agriculture-14-01290" class="html-bibr">33</a>] and environmental impact reduction [<a href="#B34-agriculture-14-01290" class="html-bibr">34</a>,<a href="#B35-agriculture-14-01290" class="html-bibr">35</a>,<a href="#B36-agriculture-14-01290" class="html-bibr">36</a>].</p>
Full article ">Figure 2
<p>Schematic illustration of net ecosystem carbon balance (NECB) and net greenhouse gas balance (NGB) in agroecosystem.</p>
Full article ">Figure 3
<p>Relationship between manure application rate and net ecosystem carbon balance (NECB) and net greenhouse gas balance (NGB) in Japanese managed grasslands at Nakashibetsu (NKS), Shizunai (SZN), Nas-Shiobara (NSS) and Kobayashi (KBY) (produced from [<a href="#B62-agriculture-14-01290" class="html-bibr">62</a>]).</p>
Full article ">Figure 4
<p>Breakdown of net ecosystem carbon balance (NECB) in four grasslands in Japan: net ecosystem production (NEP), carbon imported through manure application (Manure), carbon exported through harvest (Harvest); Nakashibetsu (NKS), Shizunai (SZN), Nas-Shiobara (NSS) and Kobayashi (KBY) (Produced from [<a href="#B62-agriculture-14-01290" class="html-bibr">62</a>]).</p>
Full article ">
23 pages, 9401 KiB  
Article
Refinement of Cropland Data Layer with Effective Confidence Layer Interval and Image Filtering
by Reza Maleki, Falin Wu, Amel Oubara, Loghman Fathollahi and Gongliu Yang
Agriculture 2024, 14(8), 1285; https://doi.org/10.3390/agriculture14081285 - 4 Aug 2024
Viewed by 383
Abstract
Various systems have been developed to process agricultural land data for better management of crop production. One such system is Cropland Data Layer (CDL), produced by the National Agricultural Statistics Service of the United States Department of Agriculture (USDA). The CDL has been [...] Read more.
Various systems have been developed to process agricultural land data for better management of crop production. One such system is Cropland Data Layer (CDL), produced by the National Agricultural Statistics Service of the United States Department of Agriculture (USDA). The CDL has been widely used for training deep learning (DL) segmentation models. However, it contains various errors, such as salt-and-pepper noise, and must be refined before being used in DL training. In this study, we used two approaches to refine the CDL for DL segmentation of major crops from a time series of Sentinel-2 monthly composite images. Firstly, different confidence intervals of the confidence layer were used to refine the CDL. Secondly, several image filters were employed to improve data quality. The refined CDLs were then used as the ground-truth in DL segmentation training and evaluation. The results demonstrate that the CDL with +45% and +55% confidence intervals produced the best results, improving the accuracy of DL segmentation by approximately 1% compared to non-refined data. Additionally, filtering the CDL using the majority and expand–shrink filters yielded the best performance, enhancing the evaluation metrics by about 1.5%. The findings suggest that pre-filtering the CDL and selecting an effective confidence interval can significantly improve DL segmentation performance, contributing to more accurate and reliable agricultural monitoring. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic overview and data layers used in the study. The top-left map shows the study area within the Mississippi Delta (blue rectangle) and T14TNK test area (red rectangle). The top-right image displays Sentinel-2 composite imagery from May 2021. The bottom-left map illustrates the distribution of crops in the CDL. The bottom-right map depicts the CDL confidence layer, indicating the confidence values associated with the CDL.</p>
Full article ">Figure 2
<p>NDVI profiles for the major crops in the study area throughout the year 2021.</p>
Full article ">Figure 3
<p>The 2021 CDL land cover distribution in the study area.</p>
Full article ">Figure 4
<p>Flowchart summarizing the research approach for major crop mapping from Sentinel-2 imagery using various CDL confidence levels and image filters.</p>
Full article ">Figure 5
<p>Distribution of crop pixels refined by different confidence levels of major crops within the study area for the year 2021.</p>
Full article ">Figure 6
<p>Impact of varying confidence intervals on the refinement of the CDL. The percentages indicate the confidence thresholds used to refine the CDL.</p>
Full article ">Figure 7
<p>DL segmentation results of croplands from Sentinel-2 imagery. The DL models were trained using different CDL confidence layer intervals.</p>
Full article ">Figure 8
<p>Confusion matrices illustrating the performance of DL models trained with different CDL confidence intervals in segmenting major crops. Diagonal numbers represent the percentage of correctly classified instances for each crop, while non-diagonal numbers indicate the percentage of misclassified instances between different crops.</p>
Full article ">Figure 9
<p>Filtered CDL and corresponding DL segmentation results for major crops using various image filters. The “No Filter” and “R-CDL” results are included for comparison.</p>
Full article ">Figure 10
<p>Confusion matrices illustrating the performance of DL models using different image filters on the CDL for major crop segmentation. The values within the matrix represent the percentage of correctly and incorrectly classified instances, ranging from 0 to 100.</p>
Full article ">Figure 11
<p>Comparison of DL accuracy metric results using different CDL refinement methods, including the R-CDL from Lin et al.’s study [<a href="#B9-agriculture-14-01285" class="html-bibr">9</a>].</p>
Full article ">Figure 12
<p>F1-score trends comparing the performance of different refining methods between the Mississippi Delta and T14TNK areas. The chart illustrates the accuracy metrics for various confidence intervals and filtering techniques, highlighting the generalizability and robustness of the methods across different geographical regions.</p>
Full article ">
22 pages, 16958 KiB  
Article
Analysis of the Spatiotemporal Changes in Cropland Occupation and Supplementation Area in the Pearl River Delta and Their Impacts on Carbon Storage
by Shu-Qi Huang, Da-Fang Wu, Jin-Yao Lin, Yue-Ling Pan and Ping Zhou
Land 2024, 13(8), 1195; https://doi.org/10.3390/land13081195 - 3 Aug 2024
Viewed by 309
Abstract
In recent years, the “dual carbon” issue has become a major focus of the international community. Changes in land use driven by anthropogenic activities have a profound impact on ecosystem structure and carbon cycling. This study quantitatively assesses the spatiotemporal changes in cropland [...] Read more.
In recent years, the “dual carbon” issue has become a major focus of the international community. Changes in land use driven by anthropogenic activities have a profound impact on ecosystem structure and carbon cycling. This study quantitatively assesses the spatiotemporal changes in cropland occupation and supplementation in the Pearl River Delta from 2000 to 2020 using the InVEST model, analyzing the spatial clustering of carbon storage changes caused by variations in cropland area. The PLUS model was employed to simulate land-use patterns and the spatial distribution of carbon storage in four future development scenarios. The results indicate the following: (1) From 2000 to 2020, the net change rate of cropland area in the Pearl River Delta was −0.81%, with a decrease of 16.49 km2 in cropland area, primarily converted to built-up land and forest land. (2) Carbon storage in the Pearl River Delta exhibited a pattern of lower values in the center and higher values in the periphery. The terrestrial ecosystem carbon storage in the Pearl River Delta was 534.62 × 106 t in 2000, 518.60 × 106 t in 2010, and 512.57 × 106 t in 2020, showing an overall decreasing trend. The conversion of cropland and forest land was the main reason for the decline in total regional carbon storage. (3) The area of carbon sequestration lost due to cropland occupation was significantly greater than the area of carbon loss compensated by new cropland, indicating an imbalance in the quality of cropland occupation and supplementation as a crucial factor contributing to regional carbon loss. (4) Under the ecological priority scenario, the expansion of built-up land and the reduction in ecological land such as cropland and forest land were effectively controlled, resulting in the minimal loss of carbon storage. The soil organic carbon pool of cropland is the most active carbon pool in terrestrial ecosystems and has a significant impact on carbon storage. Clarifying the relationship between “cropland protection measures–land use changes–ecosystem carbon storage” will improve cropland protection policies, provide references for regional carbon sequestration enhancement, and support sustainable socio-economic development. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the Pearl River Delta.</p>
Full article ">Figure 2
<p>Temporal and spatial changes in cultivated land area in the Pearl River Delta from 2000 to 2020.</p>
Full article ">Figure 3
<p>Distribution of farmland occupation and compensation from 2000 to 2020.</p>
Full article ">Figure 4
<p>Net change rate of farmland area in various cities of the Pearl River Delta from 2000 to 2020.</p>
Full article ">Figure 5
<p>Variations in carbon reserves over time and space in the Pearl River Delta, 2000–2020.</p>
Full article ">Figure 6
<p>Carbon storage of various land use types from 2000 to 2020.</p>
Full article ">Figure 7
<p>Spatial distribution changes in carbon storage induced by cultivated land occupation and compensation in the Pearl River Delta from 2000 to 2020.</p>
Full article ">Figure 8
<p>Hot spot analysis of carbon storage changes resulting from land conversion in the Pearl River Delta.</p>
Full article ">Figure 9
<p>Land use in the Pearl River Delta for 2030: spatial simulation in four distinct scenarios.</p>
Full article ">Figure 10
<p>Spatial distribution simulation of carbon stock in different scenarios in the Pearl River Delta in 2030.</p>
Full article ">
20 pages, 6206 KiB  
Article
From Expansion to Shrinkage: An Assessment of the Carbon Effect from Spatial Reconfiguration of Rural Human Settlements in the Wuhan Metropolitan Area
by Yingxue Rao, Chenxi Wu and Qingsong He
Land 2024, 13(8), 1176; https://doi.org/10.3390/land13081176 - 30 Jul 2024
Viewed by 330
Abstract
Nowadays, the reorganization of rural land-use space exhibits a dynamic process of expansion and shrinkage. Taking the Wuhan Metropolitan Area as an example, this study used the InVEST model to quantitatively assess changes in rural built-up land between 1995 and 2020 and its [...] Read more.
Nowadays, the reorganization of rural land-use space exhibits a dynamic process of expansion and shrinkage. Taking the Wuhan Metropolitan Area as an example, this study used the InVEST model to quantitatively assess changes in rural built-up land between 1995 and 2020 and its impact on regional carbon storage. Combined with the PLUS model, further simulations were carried out to predict the heterogeneous mechanisms of shrinkage and expansion of rural habitable space under three scenarios in 2030. The results indicate that the area of rural built-up land in the Wuhan Metropolitan Area showed an overall increasing trend, with shrinkage mainly concentrated in the Wuhan-Ezhou border, Tianmen, and southern Xiantao, while expansion displayed a decentralized point distribution. The PLUS model predicts that, in the scenario of rural built-up land expansion, a significant amount of cropland is encroached upon. This study provides a new perspective for understanding the impact of rural habitat changes on the carbon cycle. Future land management and planning should pay more attention to maintaining ecosystem services and considering the environmental effects of changes in rural built-up land layout. Full article
(This article belongs to the Section Land Environmental and Policy Impact Assessment)
Show Figures

Figure 1

Figure 1
<p>Location of The Wuhan Metropolitan Area.</p>
Full article ">Figure 2
<p>Research Framework.</p>
Full article ">Figure 3
<p>Rural built-up land area and proportion of overall land area, 1995–2020. The bar chart represents the area of rural built-up land. The line represents the proportion of rural built-up land area to the total land area.</p>
Full article ">Figure 4
<p>Changes in rural built-up land in the Wuhan Metropolitan Area, 1995–2020. (<b>a</b>) represents the spatial distribution of the shrinkage of rural built-up land; (<b>b</b>)represents the spatial distribution of the expansion of rural built-up land. Note: Due to the small size of the parcel changes, this map bolds the outlines of parcels where rural built-up land has changed. Consequently, the map only represents the location of land type changes, and the pixel occupancy does not reflect the actual area converted. For specific conversion areas, please refer to <a href="#land-13-01176-t004" class="html-table">Table 4</a> and <a href="#land-13-01176-t005" class="html-table">Table 5</a>.</p>
Full article ">Figure 5
<p>Carbon storage in seven land use types in Wuhan urban area, 1995–2020. Note: * is urban built-up land, ** is rural built-up land.</p>
Full article ">Figure 6
<p>Comparison between the current land use status in 2020 and PLUS simulation. (<b>a</b>) represents the actual land use; (<b>b</b>) represents the land use simulated by PLUS model.</p>
Full article ">Figure 7
<p>Spatial simulation of land use under three different scenarios in Wuhan Metropolitan Area in 2030. (<b>a</b>) represents a natural development scenario; (<b>b</b>) represents a scenario of rural built-up land shrinkage; (<b>c</b>) represents a scenario of rural built-up land expansion.</p>
Full article ">Figure 8
<p>Changes in the spatial pattern of rural built-up land in Wuhan Metropolitan Area. (<b>a</b>) represents a natural development scenario; (<b>b</b>) represents a scenario of rural built-up land shrinkage; (<b>c</b>) represents a scenario of rural built-up land expansion.</p>
Full article ">Figure 9
<p>Simulation of the spatial distribution of carbon storages in Wuhan Metropolitan Area under three different scenarios in 2030 (t/ha). (<b>a</b>) represents a natural development scenario; (<b>b</b>) represents a scenario of rural built-up land shrinkage; (<b>c</b>) represents a scenario of rural built-up land expansion.</p>
Full article ">
Back to TopTop