Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (6,307)

Search Parameters:
Keywords = spatial predictive modeling

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 872 KiB  
Article
Estimating the Relative Risks of Spatial Clusters Using a Predictor–Corrector Method
by Majid Bani-Yaghoub, Kamel Rekab, Julia Pluta and Said Tabharit
Mathematics 2025, 13(2), 180; https://doi.org/10.3390/math13020180 (registering DOI) - 7 Jan 2025
Abstract
Spatial, temporal, and space–time scan statistics can be used for geographical surveillance, identifying temporal and spatial patterns, and detecting outliers. While statistical cluster analysis is a valuable tool for identifying patterns, optimizing resource allocation, and supporting decision-making, accurately predicting future spatial clusters remains [...] Read more.
Spatial, temporal, and space–time scan statistics can be used for geographical surveillance, identifying temporal and spatial patterns, and detecting outliers. While statistical cluster analysis is a valuable tool for identifying patterns, optimizing resource allocation, and supporting decision-making, accurately predicting future spatial clusters remains a significant challenge. Given the known relative risks of spatial clusters over the past k time intervals, the main objective of the present study is to predict the relative risks for the subsequent interval, k+1. Building on our prior research, we propose a predictive Markov chain model with an embedded corrector component. This corrector utilizes either multiple linear regression or an exponential smoothing method, selecting the one that minimizes the relative distance between the observed and predicted values in the k-th interval. To test the proposed method, we first calculated the relative risks of statistically significant spatial clusters of COVID-19 mortality in the U.S. over seven time intervals from May 2020 to March 2023. Then, for each time interval, we selected the top 25 clusters with the highest relative risks and iteratively predicted the relative risks of clusters from intervals three to seven. The predictive accuracies ranged from moderate to high, indicating the potential applicability of this method for predictive disease analytic and future pandemic preparedness. Full article
(This article belongs to the Section Computational and Applied Mathematics)
20 pages, 15608 KiB  
Article
Study of Shale Gas Source Rock S-Wave Structure Characteristics via Dense Array Ambient Noise Tomography in Zhangjiakou, China
by Si Chen, Zhanwu Lu, Haiyan Wang, Qingyu Wu, Wei Cai, Guowei Wu and Guangwen Wang
Remote Sens. 2025, 17(2), 183; https://doi.org/10.3390/rs17020183 - 7 Jan 2025
Abstract
Utilizing short-period dense seismic arrays, ambient noise tomography has proven effective in delineating continuous geological structures, a task critical for characterizing shale gas reservoir configurations. This study deployed 153 short-period seismic stations across the Xiahuayuan District in Zhangjiakou, a region with prospective shale [...] Read more.
Utilizing short-period dense seismic arrays, ambient noise tomography has proven effective in delineating continuous geological structures, a task critical for characterizing shale gas reservoir configurations. This study deployed 153 short-period seismic stations across the Xiahuayuan District in Zhangjiakou, a region with prospective shale gas deposits, to perform an ambient noise tomography survey. Through a meticulous process involving cross-correlation analysis, dispersion curve extraction, and subsequent inversion, a three-dimensional velocity structure model of the area was constructed. The model discerns subtle velocity changes within the 0–3 km depth interval, achieving a horizontal resolution of approximately 1.5 km in the 0–3 km stratum, thereby effectively delineating the shale reservoir structure. Integration of the velocity model with regional geological data facilitated a comprehensive interpretation and structural analysis of the prospective shale gas zone. Low-velocity anomalies observed within the velocity structure correspond to the spatial distribution of the Xiahuayuan Formation, likely attributable to the prevalent stratum of mudstone shale deposits within this formation. Employing a binary stratigraphic model, the study predicted shale content based on the velocity structure, with predictions exhibiting a moderate correlation (correlation coefficient of 0.58) with empirical data. This suggests the presented method as a viable rapid estimation technique for assessing the shale content of target strata. Full article
Show Figures

Figure 1

Figure 1
<p>Location of nodal seismographs station deployment in the research area.</p>
Full article ">Figure 2
<p>Regional geological features and station locations (geological map provided by the Hebei Coalfield Geological Exploration Institute).</p>
Full article ">Figure 3
<p>Stratigraphic column of the Xiahuayuan Formation. the (<b>a</b>,<b>b</b>) image is the mudstone shale and coal of the Xiahuayuan Formation. The rocks mainly exhibit thin layers and are relatively fragmented. (<b>c</b>) mainly shows the sandstone layer in the lower part of the Xiahuayuan Formation with intact structure. The data are provided by the Hebei Coalfield Geological Exploration Institute.</p>
Full article ">Figure 4
<p>Station 6153 cross-correlates results with other stations. The image illustrates the characteristics of surface waves in the array with various bandpass filters, and these characteristics are effectively captured within the filtering parameters of 0.8–6 s.</p>
Full article ">Figure 5
<p>Dispersion curve extraction process schematic diagram.</p>
Full article ">Figure 6
<p>Characteristics of the overall distribution of the dispersion curve. The features of the dispersion curve collection are illustrated via a 2D histogram. Between 0 and 3 s, the dispersion curve features are more concentrated, whereas after 3 s, the dispersion curves become more dispersed. This suggests that the shallow subsurface structures are straightforward, whereas the deep subsurface structures are complex.</p>
Full article ">Figure 7
<p>The number of iterations and the trend of parameter changes.</p>
Full article ">Figure 8
<p>Average sensitivity kernel testing. The figure illustrates that the velocity structure model responds well to depths of 0–5 km.</p>
Full article ">Figure 9
<p>Checkerboard test results at different depths. The dense array data were tested respectively with synthetic 5% Gaussian random noise. The figure reflects the ability of the velocity structure model to resolve anomalies with a scale of 1.5 km. The (<b>a</b>) represents the checkerboard pattern model, while (<b>b</b>–<b>i</b>) correspond to depths of 0.8 km, 1.2 km, 1.6 km, 2.0 km, 2.4 km, 2.8 km, 3.6 km, and 4.8 km, respectively. The dashed box is the area with the better recovery effect. Based on the results, it is evident that the recovery effect is more pronounced within the range of 1–3 km.</p>
Full article ">Figure 10
<p>Depth slice features of shear wave velocity structure. The image demonstrates the characteristics of low-velocity structures. In the strata at a depth of 0–3 km, the occurrence of low-velocity structures closely aligns with the distribution of the Xiahuayuan Formation. Triangles represent stations.</p>
Full article ">Figure 11
<p>Mudstone shale contours of Xiahuayuan formation. This study compiled contour features of the thickness of the Xiahuayuan Formation from all the drilling wells. Note that some wells did not penetrate the Xiahuayuan Formation; thus, the thicknesses represent only those revealed by the drilling.</p>
Full article ">Figure 12
<p>City structure vertical slice. The blue and red triangle represents the top and bottom boundary of the Xiahuayuan Formation by well drilling. In the absence of drilling data within the (<b>AA’</b>,<b>BB’</b>) profiles, it is inferred that the low-velocity structure is attributable to the muddy shale of the Xiahuayuan Formation. In the (<b>CC’</b>,<b>DD’</b>) profiles, the deeper low-velocity structures are more likely to be caused by the muddy shale of the Xiamaling Formation. Please refer to <a href="#remotesensing-17-00183-f001" class="html-fig">Figure 1</a> for the (<b>AA’</b>–<b>FF’</b>) position.</p>
Full article ">Figure 13
<p>Binary stratigraphic model: (<b>a</b>) actual strata; (<b>b</b>) equivalent model.</p>
Full article ">Figure 14
<p>Velocity distribution with depth characteristics of the drilling area.</p>
Full article ">Figure 15
<p>Shale content distribution map (<b>a</b>) and rock B percentage prediction (<b>b</b>). Based on the comparison images, the trend in large-scale prediction results is consistent. However, due to the current resolution of this method, there is a significant deviation in predicting details.</p>
Full article ">Figure 16
<p>Prediction of mudstone shale thickness with actual mudstone shale thickness.</p>
Full article ">
19 pages, 3088 KiB  
Article
Predicting the Spatial Distribution of Soil Organic Carbon in the Black Soil Area of Northeast Plain, China
by Yunfeng Li, Zhuo Chen, Yang Chen, Taotao Li, Cen Wang and Chaoteng Li
Sustainability 2025, 17(2), 396; https://doi.org/10.3390/su17020396 - 7 Jan 2025
Abstract
The accurate prediction of the spatial distribution of soil organic carbon (SOC) and the identification of the mechanisms underlying its spatial differentiation are of paramount significance for the conservation and utilization of land and regional sustainable development. A total of 512 soil samples [...] Read more.
The accurate prediction of the spatial distribution of soil organic carbon (SOC) and the identification of the mechanisms underlying its spatial differentiation are of paramount significance for the conservation and utilization of land and regional sustainable development. A total of 512 soil samples were collected from Wuchang and Shuangcheng County in Harbin City, Heilongjiang Province, China, which served as the study area. Six machine learning models, including Random Forest (RF), AdaBoost, Support Vector Regression (SVR), weighted average, Stacking, and Blending, were utilized to predict the spatial distribution of SOC and analyze its spatial differentiation. The result reveals that 12 environmental variables, including soil type, bulk density, pH, average annual precipitation, average annual temperature, net primary productivity (NPP), land use type, normalized difference vegetation index (NDVI), slope, elevation, soil parent material, and distance to rivers, are effective influencing factors on SOC in the study area. It turns out that the Stacking model, with an R2 of 0.4327, performed the best in this study, followed by the weighted average, Blending, RF, AdaBoost, and SVR models; a heterogeneous integrated learning model may be more robust than an individual learner. The predicted SOC content is generally lower in the northwestern arable land and higher in the southeastern forest land. In addition, SOC differentiation shows that forest land and grass land with dark brown soil or swamp soil, soil covering igneous and metamorphic rocks with various minerals, higher elevation and slope, and suitable water-thermal and soil intrinsic conditions for aerobic microbial activity benefit the enrichment of SOC in the study area. The enrichment and depletion of SOC are jointly influenced by pedogenesis, microbial activity, and biodiversity. Full article
30 pages, 10463 KiB  
Article
Enhancing Soil Moisture Prediction in Drought-Prone Agricultural Regions Using Remote Sensing and Machine Learning Approaches
by Xizhuoma Zha, Shaofeng Jia, Yan Han, Wenbin Zhu and Aifeng Lv
Remote Sens. 2025, 17(2), 181; https://doi.org/10.3390/rs17020181 - 7 Jan 2025
Abstract
The North China Plain is a crucial agricultural region in China, but irregular precipitation patterns have led to significant water shortages. To address this, analyzing the high-resolution dynamics of root-zone soil moisture transport is essential for optimizing irrigation strategies and improving water resource [...] Read more.
The North China Plain is a crucial agricultural region in China, but irregular precipitation patterns have led to significant water shortages. To address this, analyzing the high-resolution dynamics of root-zone soil moisture transport is essential for optimizing irrigation strategies and improving water resource efficiency. The Richards equation is a robust model for describing soil moisture transport dynamics across multiple soil layers, yet its application at large spatial scales is hindered by its sensitivity to boundary conditions and model parameters. This study introduces a novel approach that, for the first time, employs a continuous time series of near-surface soil moisture as the upper boundary condition in the Richards equation to estimate high-resolution root-zone soil moisture in the North China Plain, thus enabling its large-scale application. Singular spectrum analysis (SSA) was first applied to reconstruct site-specific time series, filling in missing and singular values. Leveraging observational data from 617 monitoring sites across the North China Plain and multiple spatial covariates, we developed a machine learning model to estimate near-surface soil moisture at a 1 km resolution. This high-resolution, continuous near-surface soil moisture series then served as the upper boundary condition for the Richards equation, facilitating the estimation of root-zone soil moisture across the region. The results indicated that the machine learning model achieved a correlation coefficient (R) of 0.92 for estimating spatial near-surface soil moisture. Analysis of spatial covariates showed that atmospheric forcing factors, particularly temperature and evaporation, had the most substantial impact on model performance, followed by static factors such as latitude, longitude, and soil texture. With a continuous time series of near-surface soil moisture, the Richards equation method accurately predicted multi-layer soil moisture and demonstrated its applicability for large-scale spatial use. The model yielded R values of 0.97, 0.78, 0.618, and 0.43, with RMSEs of 0.024, 0.06, 0.08, and 0.11, respectively, for soil layers at depths of 10 cm, 20 cm, 40 cm, and 100 cm across the North China Plain. Full article
(This article belongs to the Special Issue Mapping Essential Elements of Agricultural Land Using Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Distribution of vegetation types and administrative divisions in the NPC.</p>
Full article ">Figure 2
<p>Flowchart for estimating soil moisture at different spatial layers in the North China Plain based on the Richards equation.</p>
Full article ">Figure 3
<p>Definition sketch of <math display="inline"><semantics> <mrow> <mi mathvariant="normal">α</mi> <mo>(</mo> <mi mathvariant="normal">ψ</mi> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>(<b>a</b>) Comparison between SSA-predicted time series and observed trends; (<b>b</b>) time series comparison for data imputation at the start and end of the year; (<b>c</b>) scatter plot of SSA results.</p>
Full article ">Figure 4 Cont.
<p>(<b>a</b>) Comparison between SSA-predicted time series and observed trends; (<b>b</b>) time series comparison for data imputation at the start and end of the year; (<b>c</b>) scatter plot of SSA results.</p>
Full article ">Figure 5
<p>(<b>a</b>) Scatter plot for RF validation; (<b>b</b>) SHAP value interpretation of RF model outputs. Note: In the figure, Tair represents the air temperature 2 m above the surface of land, ocean, or inland water bodies; LST_diff denotes the daily difference in land surface temperature.</p>
Full article ">Figure 5 Cont.
<p>(<b>a</b>) Scatter plot for RF validation; (<b>b</b>) SHAP value interpretation of RF model outputs. Note: In the figure, Tair represents the air temperature 2 m above the surface of land, ocean, or inland water bodies; LST_diff denotes the daily difference in land surface temperature.</p>
Full article ">Figure 6
<p>Comparison between RF model calculations and site observations.</p>
Full article ">Figure 7
<p>(<b>a</b>) Histogram of R value over time series; (<b>b</b>) Histogram of RMSE value over time series.</p>
Full article ">Figure 8
<p>Spatial distribution of R values and RMSE values.</p>
Full article ">Figure 9
<p>(<b>a</b>) Spatial distribution of SSM; (<b>b</b>) Spatial distribution of RZSM.</p>
Full article ">Figure 9 Cont.
<p>(<b>a</b>) Spatial distribution of SSM; (<b>b</b>) Spatial distribution of RZSM.</p>
Full article ">Figure 10
<p>Comparison of soil moisture in each layer between SSMRE model simulation and site measured values.</p>
Full article ">Figure 11
<p>(<b>a</b>) SSM RMSE value; (<b>b</b>) Root layer soil moisture RMSE value changing trend.</p>
Full article ">Figure 12
<p>(<b>a</b>) R value of each layer verified at different sites; (<b>b</b>) RMSE value in each layer verified at different sites.</p>
Full article ">Figure 13
<p>(<b>a</b>) The R values for the multi-layer soil profile were calculated using the Richards equation, with high-resolution SSM as the upper boundary condition; (<b>b</b>) The RMSE values for the multi-layer soil profile.</p>
Full article ">Figure 14
<p>Box plots of parameter sensitivity analyses of soil moisture model outputs by layer.</p>
Full article ">Figure 15
<p>Histogram of the sensitivity analysis of different parameters.</p>
Full article ">
41 pages, 24290 KiB  
Article
Assessing the Impact of Land Use and Land Cover Change on Environmental Parameters in Khyber Pakhtunkhwa, Pakistan: A Comprehensive Study and Future Projections
by Mehjabeen Khan and Ruishan Chen
Remote Sens. 2025, 17(1), 170; https://doi.org/10.3390/rs17010170 - 6 Jan 2025
Viewed by 231
Abstract
Land use and land cover (LULC) change, driven by environmental and human activities, significantly impacts ecosystems, climate, biodiversity, and socio-economic systems. This study focuses on Khyber Pakhtunkhwa (KPK), Pakistan, a region with sensitive ecosystems and diverse landscapes, to analyze LULC dynamics and their [...] Read more.
Land use and land cover (LULC) change, driven by environmental and human activities, significantly impacts ecosystems, climate, biodiversity, and socio-economic systems. This study focuses on Khyber Pakhtunkhwa (KPK), Pakistan, a region with sensitive ecosystems and diverse landscapes, to analyze LULC dynamics and their environmental consequences. Based on Landsat imagery from 2000, 2010, and 2020, we used the Random Forest algorithm on Google Earth Engine (GEE) to classify LULC, and the CA-ANN model to project future scenarios for 2030, 2050, and 2100. Additional simulations were conducted using the MOLUSCE Plugin in QGIS. The results revealed a 138.02% (4071.98 km2) increase in urban areas from 2000 to 2020, marking urbanization as a major driver of LULC change. Urban expansion strongly correlated with land surface temperature (LST) (R2 = 0.89), amplifying the urban heat island effect. Rising LST showed negative correlations with the key environmental indices NDVI (−0.88), MNDWI (−0.49), and NDMI (−0.62), signaling declining vegetation cover, water resources, and soil moisture, respectively. Projections for 2100 predict LST rising to 55.3 °C, with NDVI, MNDWI, and NDMI dropping to 0.36, 0.17, and 0.21, respectively. Vegetation health, as indicated by the Leaf Area Index (LAI), also declined, with maximum and minimum values falling from 4.66 and −5.75 in 2000 to 2.16 and −2.55 in 2020, reflecting increased barren land and reduced greenness. The spatial analysis highlights significant transitions from vegetated to barren or urban land, leading to declining moisture levels, water stress, soil erosion, and biodiversity. Projections show continued reductions in forests, vegetation, and agricultural lands, replaced by barren and built-up areas. Declines in key indices such as NDVI, MNDWI, and NDMI indicate deteriorating vegetation, water resources, and soil moisture levels. These findings emphasize the need for sustainable urban planning and environmental management. Expanding urban green spaces, using reflective materials, and preserving vegetation and water resources are vital to mitigating heat island effects and maintaining ecological balance. Anticipated declines in LST, NDVI, MNDWI, NDMI, and LAI stress the urgency for climate adaptation strategies to protect human health, ecosystem services, and economic stability in KPK. Full article
(This article belongs to the Special Issue Advances of Remote Sensing in Land Cover and Land Use Mapping)
Show Figures

Figure 1

Figure 1
<p>Study area map of Khyber Pakht-unkhwa showing major road networks and major rivers.</p>
Full article ">Figure 2
<p>LULC calculate framework diagram.</p>
Full article ">Figure 3
<p>Methodology for predicting LULC, NDVI, MNDWI, LST, LAI, and NDMI and for the years 2030, 2050, and 2100.</p>
Full article ">Figure 4
<p>Methodology for calculating environmental parameters (LST, NDVI, MNDWI, LAI, and NDMI) in KPK, Pakistan.</p>
Full article ">Figure 5
<p>Land use and land cover (LULC) classification in Khyber Pakhtunkhwa, Pakistan for the years (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 6
<p>Area of each land cover class for the years 2000, 2010, and 2020.</p>
Full article ">Figure 7
<p>Land surface temperature (LST) distribution in Khyber Pakht-unkhwa, Pakistan in (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 8
<p>Digital elevation model (DEM) of Khyber Pakht-unkhwa (KPK) region highlighting topographical variation.</p>
Full article ">Figure 9
<p>Modified normalized difference water index (MNDWI) in Khyber Pakht-unkhwa, Pakistan for the years (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 10
<p>Normalized difference vegetation index (NDVI) in Khyber Pakht-unkhwa, Pakistan for the years (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 11
<p>Leaf area index (LAI) analysis in Khyber Pakh-tunkhwa, Pakistan for the years (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 12
<p>Normalized difference moisture index (NDMI) analysis in Khyber Pakh-tunkhwa, Pakistan for the years (<b>a</b>) 2000, (<b>b</b>) 2010, (<b>c</b>) 2020.</p>
Full article ">Figure 13
<p>Land use and land cover (LULC) classification in Khyber Pakh-tunkhwa, Pakistan in (<b>a</b>) 2030, (<b>b</b>) 2050, (<b>c</b>) 2100.</p>
Full article ">Figure 14
<p>Areas in square kilometers for each land cover class for the years 2030, 2050, and 2100.</p>
Full article ">Figure 15
<p>Land surface temperature (LST) distribution in Khyber Pakh-tunkhwa, Pakistan in (<b>a</b>) 2030, (<b>b</b>) 2050, (<b>c</b>) 2100.</p>
Full article ">Figure 16
<p>Normalized difference vegetation index (NDVI) in Khyber Pakh-tunkhwa, Pakistan in (<b>a</b>) 2030, (<b>b</b>) 2050, (<b>c</b>) 2100.</p>
Full article ">Figure 17
<p>Modified normalized difference water index (MNDWI) in Khyber Pakh-tunkhwa, Pakistan in (<b>a</b>) 2030, (<b>b</b>) 2050, and (<b>c</b>) 2100.</p>
Full article ">Figure 18
<p>Leaf area index (LAI) analysis in Khyber Pakh-tunkhwa, Pakistan for the years (<b>a</b>) 2030, (<b>b</b>) 2050, and (<b>c</b>) 2100.</p>
Full article ">Figure 19
<p>Prediction of the normalized difference moisture index (NDMI) values in Khyber Pakhtunkhwa, Pakistan for the years (<b>a</b>) 2030, (<b>b</b>) 2050, and (<b>c</b>) 2100.</p>
Full article ">Figure 20
<p>Correlation between LAI-LST.</p>
Full article ">Figure 21
<p>Correlation between LAI-NDVI.</p>
Full article ">Figure 22
<p>Correlation between LST-NDVI.</p>
Full article ">Figure 23
<p>Correlation between MNDWI-NDVI.</p>
Full article ">Figure 24
<p>Correlation between MNDWI-LST.</p>
Full article ">Figure 25
<p>Correlation between NDMI-LAI.</p>
Full article ">Figure 26
<p>Correlation between LST-NDMI.</p>
Full article ">Figure 27
<p>Correlation between NDVI-NDMI.</p>
Full article ">
26 pages, 1539 KiB  
Article
An Adaptive Spatio-Temporal Traffic Flow Prediction Using Self-Attention and Multi-Graph Networks
by Basma Alsehaimi, Ohoud Alzamzami, Nahed Alowidi and Manar Ali
Sensors 2025, 25(1), 282; https://doi.org/10.3390/s25010282 - 6 Jan 2025
Viewed by 271
Abstract
Traffic flow prediction is a pivotal element in Intelligent Transportation Systems (ITSs) that provides significant opportunities for real-world applications. Capturing complex and dynamic spatio-temporal patterns within traffic data remains a significant challenge for traffic flow prediction. Different approaches to effectively modeling complex spatio-temporal [...] Read more.
Traffic flow prediction is a pivotal element in Intelligent Transportation Systems (ITSs) that provides significant opportunities for real-world applications. Capturing complex and dynamic spatio-temporal patterns within traffic data remains a significant challenge for traffic flow prediction. Different approaches to effectively modeling complex spatio-temporal correlations within traffic data have been proposed. These approaches often rely on a single model to capture temporal dependencies, which neglects the varying influences of different time periods on traffic flow. Additionally, these models frequently utilize either static or dynamic graphs to represent spatial dependencies, which limits their ability to address complex and overlapping spatial relationships. Moreover, some approaches struggle to fully capture spatio-temporal variations, leading to the exclusion of critical information and ultimately resulting in suboptimal prediction performance. Thus, this paper introduces the Adaptive Spatio-Temporal Attention-Based Multi-Model (ASTAM), an architecture designed to capture spatio-temporal dependencies within traffic data. The ASTAM employs multi-temporal gated convolution with multi-scale temporal input segments to model complex non-linear temporal correlations. It utilizes static and dynamic parallel multi-graphs to facilitate the modeling of complex spatial dependencies. Furthermore, this model incorporates a spatio-temporal self-attention mechanism to adaptively capture the dynamic and long-term spatio-temporal variations in traffic flow. Experiments conducted on four real-world datasets reveal that the proposed architecture outperformed 13 baseline approaches, achieving average reductions of 5.0% in MAE, 13.28% in RMSE, and 6.46% in MAPE across four datasets. Full article
(This article belongs to the Section Intelligent Sensors)
Show Figures

Figure 1

Figure 1
<p>The comprehensive architecture of the ASTAM.</p>
Full article ">Figure 2
<p>Creating multi-scale temporal input segments based on hourly, daily, and weekly cycles.</p>
Full article ">Figure 3
<p>(<b>a</b>) TCN architecture illustration. (<b>b</b>) A structural layout of TGC.</p>
Full article ">Figure 4
<p>An architectural diagram of MGSCM.</p>
Full article ">Figure 5
<p>A structure diagram of GAT.</p>
Full article ">Figure 6
<p>A diagram for STSAM.</p>
Full article ">Figure 7
<p>ASTAM versus different baselines for (<b>a</b>) PeMS03, (<b>b</b>) PeMS04, (<b>c</b>) PeMS07, and (<b>d</b>) PeMS08.</p>
Full article ">Figure 8
<p>The ablation experiments conducted on all datasets.</p>
Full article ">Figure 9
<p>Influence of varying the number of layers in TCN on (<b>a</b>) MAE, (<b>b</b>) RMSE, and (<b>c</b>) MAPE.</p>
Full article ">Figure 10
<p>Effect of the various embedding dimensions on (<b>a</b>) MAE, (<b>b</b>) RMSE, and (<b>c</b>) MAPE.</p>
Full article ">Figure 11
<p>Heatmap representation of the self-adaptive matrix for (<b>a</b>) PeMS03, (<b>b</b>) PeMS04, (<b>c</b>) PeMS07, and (<b>d</b>) PeMS08.</p>
Full article ">Figure 12
<p>Effectiveness of the self-adaptive matrix. (<b>a</b>) Pre-defined matrix; (<b>b</b>) self-adaptive matrix; (<b>c</b>) 48 h traffic flow on sensor pairs 15–16.</p>
Full article ">Figure 13
<p>Traffic flow prediction result visualization of (<b>a</b>) PeMS03, (<b>b</b>) PeMS04, (<b>c</b>) PeMS07, and (<b>d</b>) PeMS08.</p>
Full article ">
21 pages, 7042 KiB  
Article
Partial Discharge Recognition of Transformers Based on Data Augmentation and CNN-BiLSTM-Attention Mechanism
by Zhongjun Fu, Yuhui Wang, Lei Zhou, Keyang Li and Hang Rao
Electronics 2025, 14(1), 193; https://doi.org/10.3390/electronics14010193 - 5 Jan 2025
Viewed by 292
Abstract
Partial discharge (PD) is a commonly encountered discharge-related fault in transformers. Due to the unique characteristics of the environment where PD occurs, challenges such as difficulty in data acquisition and scarcity of samples arise. Convolutional neural networks (CNNs) are widely used in pattern [...] Read more.
Partial discharge (PD) is a commonly encountered discharge-related fault in transformers. Due to the unique characteristics of the environment where PD occurs, challenges such as difficulty in data acquisition and scarcity of samples arise. Convolutional neural networks (CNNs) are widely used in pattern recognition because of their strong feature extraction capabilities. To improve the recognition accuracy of PD models, this paper integrates CNN, bidirectional long short-term memory (BiLSTM), and an attention mechanism. In the proposed model, CNN is employed to extract local spatial and temporal features, BiLSTM is utilized to extract global bidirectional spatial and temporal features, and the attention mechanism assigns adaptive weights to the features. Additionally, to address the issues of sample scarcity and data imbalance, an improved GAN is introduced to augment the data. The experimental results demonstrate that the CNN-BiLSTM-attention method proposed in this paper significantly improves the prediction accuracy. With the help of GAN, the proposed method achieves a recognition accuracy of 97.36%, which is 1.8% higher than that of the CNN+CGAN(Conditional Generative Adversarial Network) method and 5.8% higher than that of thetraditional recognition model, SVM, making it the best-performing method among several comparable methods. Full article
Show Figures

Figure 1

Figure 1
<p>The structure of a typical convolutional neural network.</p>
Full article ">Figure 2
<p>Single pulse waveform diagram of partial discharge signal.</p>
Full article ">Figure 3
<p>Time–frequency spectrum diagram after transformation.</p>
Full article ">Figure 4
<p>Structure of generating adversarial network.</p>
Full article ">Figure 5
<p>The flowchart of the CNN model.</p>
Full article ">Figure 6
<p>Implementation process of BiLSTM model framework.</p>
Full article ">Figure 7
<p>Improved CNN-BiLSTM-attention architecture model.</p>
Full article ">Figure 8
<p>Process of signal processing.</p>
Full article ">Figure 9
<p>Waveform signal diagram before and after processing.</p>
Full article ">Figure 10
<p>Schematic of the experimental wiring setup.</p>
Full article ">Figure 11
<p>Electrode model of a typical PD defect.</p>
Full article ">Figure 12
<p>Original high-frequency waveform.</p>
Full article ">Figure 13
<p>Accuracy and loss curves of model training.</p>
Full article ">Figure 14
<p>Model training results.</p>
Full article ">Figure 15
<p>The confusion matrix of the test set trained on the model.</p>
Full article ">Figure 16
<p>Model performance evaluation method.</p>
Full article ">Figure 17
<p>Comparison of the accuracy of several fault diagnosis models.</p>
Full article ">
26 pages, 880 KiB  
Review
Diffuse Noxious Inhibitory Controls in Chronic Pain States: Insights from Pre-Clinical Studies
by Raquel Pereira-Silva, Fani L. Neto and Isabel Martins
Int. J. Mol. Sci. 2025, 26(1), 402; https://doi.org/10.3390/ijms26010402 - 5 Jan 2025
Viewed by 287
Abstract
Diffuse noxious inhibitory control (DNIC), also known as conditioned pain modulation (CPM) in humans, is a paradigm wherein the heterotopic application of a noxious stimulus results in the attenuation of another spatially distant noxious input. The pre-clinical and clinical studies show the involvement [...] Read more.
Diffuse noxious inhibitory control (DNIC), also known as conditioned pain modulation (CPM) in humans, is a paradigm wherein the heterotopic application of a noxious stimulus results in the attenuation of another spatially distant noxious input. The pre-clinical and clinical studies show the involvement of several neurochemical systems in DNIC/CPM and point to a major contribution of the noradrenergic, serotonergic, and opioidergic systems. Here, we thoroughly review the latest data on the monoaminergic and opioidergic studies, focusing particularly on pre-clinical models of chronic pain. We also conduct an in-depth analysis of these systems by integrating the available data with the descending pain modulatory circuits and the neurochemical systems therein to bring light to the mechanisms involved in the regulation of DNIC. The most recent data suggest that DNIC may have a dual outcome encompassing not only analgesic effects but also hyperalgesic effects. This duality might be explained by the underlying circuitry and the receptor subtypes involved therein. Acknowledging this duality might contribute to validating the prognostic nature of the paradigm. Additionally, DNIC/CPM may serve as a robust paradigm with predictive value for guiding pain treatment through more effective targeting of descending pain modulation. Full article
(This article belongs to the Special Issue New Insights into the Molecular Mechanisms of Chronic Pain)
Show Figures

Figure 1

Figure 1
<p>Proposed hypothetical circuits involved in the mediation of DNIC analgesia (red arrows; +DNIC), loss of DNIC analgesia (black arrows; −DNIC), and DNIC hyperalgesia (thinner purple arrows). In the <span class="html-italic">Locus coeruleus</span> (LC), two opposing circuits coexist. The first is an excitatory module, originating from dorsal LC neurons projecting to the spinal cord, which mediates DNIC analgesia. The second is an inhibitory module from ventral LC neurons projecting to spinal cord neurons (black dashed arrows), which abolishes DNIC. Both circuits exert their opposing effects through excitatory alpha-1 adrenergic receptors (a1ARs), likely located on excitatory or inhibitory (GABA) spinal cord interneurons, impinging on spinal wide dynamic range (WDR) neurons to mediate either the loss of DNIC or DNIC analgesia, respectively. In spite of the dichotomy of this circuit, functional studies emphasize the participation of the LC in DNIC analgesia. Therefore, the circuit mediating opposite effects in DNIC might contribute to adequately balancing the system in accordance with the organism’s needs. This again reflects the functioning of the descending pain modulatory system. In the A5 region, noradrenergic (NAergic) neurons projecting to the spinal cord contribute to DNIC analgesia by activating inhibitory alpha-2 adrenergic receptors (a2ARs), likely located on spinal WDR neurons. In the rostral ventromedial medulla (RVM), serotoninergic (5-HT) neurons project to either inhibitory GABAergic or excitatory spinal interneurons. Depending on the type of interneurons involved and receptors therein, this results in either DNIC analgesia or its abolishment. DNIC loss is probably mediated by the excitatory 5-HT3 receptor (5-HT3R) population that is most likely expressed in excitatory interneurons, while DNIC analgesia appears to be mediated by GABAergic interneurons that express both 5-HT3Rs and 5-HT7 receptors (5-HT7Rs). In this context, the effect of the 5-HT7R, which acts synergistically with the a2AR, is likely to become more prominent and mediate DNIC analgesia. Pre-synaptic excitatory 5-HT3Rs are also found in peripheral afferent fibers (PFAs) originating from dorsal root ganglia (DRG) neurons, which synapse onto projection neurons in the spinal cord expressing neurokinin-1 receptors (NK1+). These 5-HT3Rs are involved in a bottom-up circuit related to DNIC initiation. In the dorsal reticular nucleus (DRt), the coupling of mu-opioid receptors (MORs) to either the preferred/predominant inhibitory (Gi) proteins or the stimulatory (Gs) proteins, which are less recruited in physiological conditions (thinner purple arrow), determines whether DNIC analgesia or hyperalgesia occurs. This switch from inhibitory to excitatory signaling can disinhibit the descending facilitation from the DRt, contributing to the transition from DNIC analgesia, observed in physiological conditions, to hyperalgesia, as observed in chronic pain and prolonged opioid use. The concept of DNIC hyperalgesia challenges the established DNIC paradigm.</p>
Full article ">
24 pages, 6859 KiB  
Article
Comparative Analysis of Prediction Models for Trawling Grounds of the Argentine Shortfin Squid Illex argentinus in the Southwest Atlantic High Seas Based on Vessel Position and Fishing Log Data
by Delong Xiang, Yuyan Sun, Hanji Zhu, Jianhua Wang, Sisi Huang, Shengmao Zhang, Famou Zhang and Heng Zhang
Biology 2025, 14(1), 35; https://doi.org/10.3390/biology14010035 - 4 Jan 2025
Viewed by 256
Abstract
To evaluate and compare the effectiveness of prediction models for Argentine squid Illex argentinus trawling grounds in the Southwest Atlantic high seas based on vessel position and fishing log data, this study used AIS datasets and fishing log datasets from fishing seasons spanning [...] Read more.
To evaluate and compare the effectiveness of prediction models for Argentine squid Illex argentinus trawling grounds in the Southwest Atlantic high seas based on vessel position and fishing log data, this study used AIS datasets and fishing log datasets from fishing seasons spanning 2019–2024 (December to June each year). Using a spatial resolution of 0.1° × 0.1° and a monthly temporal resolution, we constructed two datasets—one based on vessel positions and the other on fishing logs. Fishing ground levels were defined according to the density of fishing locations, and combined with oceanographic data (sea surface temperature, 50 m water temperature, sea surface salinity, sea surface height, and mixed layer depth). A CNN-Attention deep learning model was applied to each dataset to develop Illex argentinus trawling ground prediction models. Model accuracy was then compared and potential causes for differences were analyzed. Results showed that the vessel position-based model had a higher accuracy (Accuracy = 0.813) and lower loss rate (Loss = 0.407) than the fishing log-based model (Accuracy = 0.727, Loss = 0.513). The vessel-based model achieved a prediction accuracy of 0.763 on the 2024 test set, while the fishing log-based model reached an accuracy of 0.712, slightly lower than the former, indicating the high accuracy and unique advantages of the vessel position-based model in predicting fishing grounds. Using CPUE from fishing logs as a reference, we found that the vessel position-based model performed well from January to April, whereas the CPUE-based model consistently maintained good accuracy across all months. The 2024 fishing season predictions indicated the formation of primary fishing grounds as early as January 2023, initially near the 46° S line of the Argentine Exclusive Economic Zone, with grounds shifting southeastward from March onward and reaching around 42° S by May and June. This study confirms the reliability of vessel position data in identifying fishing ground information and levels, with higher accuracy in some months compared to the fishing log-based model, thereby reducing the data lag associated with fishing logs, which are typically available a year later. Additionally, national-level fishing log data are often confidential, limiting the ability to fully consider fishing activities across the entire fishing ground region, a limitation effectively addressed by AIS vessel position data. While vessel data reflects daily catch volumes across vessels without distinguishing CPUE by species, log data provide a detailed daily CPUE breakdown by species (e.g., Illex argentinus). This distinction resulted in lower accuracy for vessel-based predictions in December 2023 and May–June 2024, suggesting the need to incorporate fishing log data for more precise assessments of fishing ground levels or resource abundance during those months. Given the near-real-time nature of vessel position data, fishing ground dynamics can be monitored in near real time. The successful development of vessel position-based prediction models aids enterprises in reducing fuel and time costs associated with indiscriminate squid searches, enhancing trawling efficiency. Additionally, such models support quota management in global fisheries by optimizing resource use, reducing fishing time, and consequently lowering carbon emissions and environmental impact, while promoting marine environmental protection in the Southwest Atlantic high seas. Full article
Show Figures

Figure 1

Figure 1
<p>Location and distribution of <span class="html-italic">Illex argentinus</span> trawling grounds and fishing points in the Southwest Atlantic high seas, 2020–2024. (<b>a</b>) shows the location of the squid trawling grounds in the Southwest Atlantic high seas; (<b>b</b>) is a statistical plot of fishing points in the northern fishing ground; (<b>c</b>) is a statistical plot of fishing points in the southern fishing ground.</p>
Full article ">Figure 2
<p>Illustration of trawler status classification using the threshold method (example of vessel Luqingyuanyu 201 with MMSI 412,329,684 from 00:00 on 1 January to 24:00 on 2 January 2023).</p>
Full article ">Figure 3
<p>Structure of the CNN-Attention model used in this study (revised from [<a href="#B33-biology-14-00035" class="html-bibr">33</a>]).</p>
Full article ">Figure 4
<p>Monthly total fishing duration per vessel during the 2020–2023 fishing seasons based on AIS data.</p>
Full article ">Figure 5
<p>Monthly average duration (hours/day/vessel) on fishing (from AIS and Logs), sailing, and drifting behavior for trawlers during the 2020–2023 fishing seasons.</p>
Full article ">Figure 6
<p>Proportional distribution of monthly trawling occurrences and CPUE values within grid cells in these squid fishing grounds from 2020 to 2023 ((<b>a</b>): proportion of trawling occurrences (%), (<b>b</b>): proportion of CPUE values (%)).</p>
Full article ">Figure 7
<p>Heatmap of Spearman correlation analysis for nine indicators.</p>
Full article ">Figure 8
<p>Accuracy and loss rates for the training and validation sets using vessel position data in the CNN-Attention model. ((<b>a</b>) shows the accuracy trends of the training and validation datasets, (<b>b</b>) shows the loss trends of the training and validation datasets).</p>
Full article ">Figure 9
<p>Accuracy and loss rates for the training and validation sets using fishing log data in the CNN-Attention model. ((<b>a</b>) shows the accuracy trends of the training and validation datasets, (<b>b</b>) shows the loss trends of the training and validation datasets).</p>
Full article ">Figure 10
<p>Predicted distribution of the squid high-seas trawling grounds based on the AIS vessel position data model.</p>
Full article ">Figure 11
<p>Predicted distribution of the squid high-seas trawling grounds based on the fishing log data model.</p>
Full article ">Figure 12
<p>Real Distribution of the squid high-seas trawling grounds based on the 2024 year CPUE determination.</p>
Full article ">Figure 13
<p>Monthly prediction accuracy of the vessel position data model and fishing log model.</p>
Full article ">
20 pages, 12082 KiB  
Article
Mapping Habitat Structures of Endangered Open Grassland Species (E. aurinia) Using a Biotope Classification Based on Very High-Resolution Imagery
by Steffen Dietenberger, Marlin M. Mueller, Andreas Henkel, Clémence Dubois, Christian Thiel and Sören Hese
Remote Sens. 2025, 17(1), 149; https://doi.org/10.3390/rs17010149 - 4 Jan 2025
Viewed by 345
Abstract
Analyzing habitat conditions and mapping habitat structures are crucial for monitoring ecosystems and implementing effective conservation measures, especially in the context of declining open grassland ecosystems in Europe. The marsh fritillary (Euphydryas aurinia), an endangered butterfly species, depends heavily on specific [...] Read more.
Analyzing habitat conditions and mapping habitat structures are crucial for monitoring ecosystems and implementing effective conservation measures, especially in the context of declining open grassland ecosystems in Europe. The marsh fritillary (Euphydryas aurinia), an endangered butterfly species, depends heavily on specific habitat conditions found in these grasslands, making it vulnerable to environmental changes. To address this, we conducted a comprehensive habitat suitability analysis within the Hainich National Park in Thuringia, Germany, leveraging very high-resolution (VHR) airborne, red-green-blue (RGB), and color-infrared (CIR) remote sensing data and deep learning techniques. We generated habitat suitability models (HSM) to gain insights into the spatial factors influencing the occurrence of E. aurinia and to predict potential habitat suitability for the whole study site. Through a deep learning classification technique, we conducted biotope mapping and generated fine-scale spatial variables to model habitat suitability. By employing various modeling techniques, including Generalized Additive Models (GAM), Generalized Linear Models (GLM), and Random Forest (RF), we assessed the influence of different modeling parameters and pseudo-absence (PA) data generation on model performance. The biotope mapping achieved an overall accuracy of 81.8%, while the subsequent HSMs yielded accuracies ranging from 0.69 to 0.75, with RF showing slightly better performance. The models agree that homogeneous grasslands, paths, hedges, and areas with dense bush encroachment are unsuitable habitats, but they differ in their identification of high-suitability areas. Shrub proximity and density were identified as important factors influencing the occurrence of E. aurinia. Our findings underscore the critical role of human intervention in preserving habitat suitability, particularly in mitigating the adverse effects of natural succession dominated by shrubs and trees. Furthermore, our approach demonstrates the potential of VHR remote sensing data in mapping small-scale butterfly habitats, offering applicability to habitat mapping for various other species. Full article
Show Figures

Figure 1

Figure 1
<p>Overview of the study area in the southern part of the HNP in central Germany. Two different types of plots (each 1 ha) have been used for collecting the reference data regarding first the biotope types and second <span class="html-italic">E. aurinia</span> occurrences.</p>
Full article ">Figure 2
<p>Overview of the methodological system used for the biotope classification.</p>
Full article ">Figure 3
<p>List of variables used as input layers in the CNN model for biotope classification. The indices are explained in <a href="#remotesensing-17-00149-t002" class="html-table">Table 2</a>.</p>
Full article ">Figure 4
<p>Variables used for the analysis of the habitat structure.</p>
Full article ">Figure 5
<p>Exemplary subsets of the classification results in biotope types (level B) using a CNN.</p>
Full article ">Figure 6
<p>Number of occurrences of <span class="html-italic">E. aurinia</span> for the respective variable value.</p>
Full article ">Figure 7
<p>Response curves of the variables used for the GAM and GLM methods to model the marsh fritillary habitat, each for two pseudo-absence datasets generated randomly.</p>
Full article ">Figure 8
<p>Projection of the habitat suitability for <span class="html-italic">E. aurinia</span> across the entire study site utilizing a RF (<b>left</b>) and GAM (<b>right</b>) model, both generated with an SRE-PA dataset. The projections are represented as probability maps, with blue areas indicating low habitat suitability and orange to red areas representing high habitat suitability.</p>
Full article ">Figure A1
<p>Response curves of the variables used for the GAM and GLM methods to model the marsh fritillary habitat, each for two PA datasets generated with the SRE-PA method.</p>
Full article ">Figure A2
<p>Variable importance in the RF model. Higher values indicate a greater influence of the variable on the model output. The variable importance was calculated using the HSM with the SRE-PA dataset 2.</p>
Full article ">
22 pages, 34247 KiB  
Article
Habitat Quality Dynamics in Urumqi over the Last Two Decades: Evidence of Land Use and Land Cover Changes
by Siying Chen, Ümüt Halik, Lei Shi, Wentao Fu, Lu Gan and Martin Welp
Land 2025, 14(1), 84; https://doi.org/10.3390/land14010084 - 3 Jan 2025
Viewed by 299
Abstract
The integrity of habitat quality is a pivotal cornerstone for the sustainable advancement of local ecological systems. Rapid urbanization has led to habitat degradation and loss of biodiversity, posing severe threats to regional sustainability, particularly in extremely vulnerable arid zones. However, systematic research [...] Read more.
The integrity of habitat quality is a pivotal cornerstone for the sustainable advancement of local ecological systems. Rapid urbanization has led to habitat degradation and loss of biodiversity, posing severe threats to regional sustainability, particularly in extremely vulnerable arid zones. However, systematic research on the assessment indicators, limiting factors, and driving mechanisms of habitat quality in arid regions is notably lacking. This study takes Urumqi, an oasis city in China’s arid region, as a case study and employs the InVEST and PLUS models to conduct a dynamic evaluation of habitat quality in Urumqi from 2000 to 2022 against the backdrop of land use changes. It also simulates habitat quality under different scenarios for the year 2035, exploring the temporal and spatial dynamics of habitat quality and its driving mechanisms. The results indicate a decline in habitat quality. The habitat quality in the southern mountainous areas is significantly superior to that surrounding the northern Gurbantunggut Desert, and it exhibits greater stability. The simulation and prediction results suggest that from 2020 to 2035, habitat degradation will be mitigated under Ecological Protection scenarios, while the decline in habitat quality will be most pronounced under Business-As-Usual scenarios. The spatial distribution of habitat quality changes in Urumqi exhibits significant autocorrelation and clustering, with these patterns intensifying over time. The observed decline in habitat quality in Urumqi is primarily driven by anthropogenic activities, urban expansion, and climate change. These factors have collectively contributed to significant alterations in the landscape, leading to the degradation of ecological conditions. To mitigate further habitat quality loss and support sustainable development, it is essential to implement rigorous ecological protection policies, adopt effective ecological risk management strategies, and promote the expansion of ecological land use. These actions are crucial for stabilizing and improving regional habitat quality in the long term. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch map of study area. (<b>a</b>) illustrates the geographical position of Urumqi, Xinjiang, within China; (<b>b</b>) presents the spatial distribution of different land use categories in Urumqi in 2022.</p>
Full article ">Figure 2
<p>Technology road map for land use modeling and habitat quality evaluation.</p>
Full article ">Figure 3
<p>Development probability of various types of land.</p>
Full article ">Figure 4
<p>Comparing real and simulated land use in 2020. The simulated spatial distribution of land use exhibits discrepancies primarily in the central urban area (<b>A</b>) and two ecological transition zones (<b>B</b>,<b>C</b>). Consequently, (<b>A</b>) depicts an enlarged view of the central urban area, (<b>B</b>) provides an enlarged view of the southwestern region, and (<b>C</b>) presents an enlarged view of the southeastern region.</p>
Full article ">Figure 5
<p>Conversion of land use types, 2000–2022.</p>
Full article ">Figure 6
<p>Spatial distribution of land use in different scenarios for 2035.</p>
Full article ">Figure 7
<p>Transformation of land use in 2022–2035.</p>
Full article ">Figure 8
<p>Quantitative changes in habitat quality level.</p>
Full article ">Figure 9
<p>Spatial distribution of habitat quality from 2000 to 2022.</p>
Full article ">Figure 10
<p>Spatial variation in habitat quality from 2000 to 2022. (<b>a</b>) Spatial pattern of habitat quality grade shift. (<b>b</b>) Spatial changes in habitat quality. (<b>c</b>) Spatiotemporal characterization of HQ cold spots and hot spots.</p>
Full article ">Figure 11
<p>Habitat quality in different scenarios for 2035. (<b>A</b>) represents the magnified view of the southwestern area, while (<b>B</b>) depicts the magnified view of the southeastern area.</p>
Full article ">Figure 12
<p>Contribution of driving factors by land use type.</p>
Full article ">
24 pages, 106853 KiB  
Article
Assessment of Vegetation Dynamics in Xinjiang Using NDVI Data and Machine Learning Models from 2000 to 2023
by Nan Ma, Shanshan Cao, Tao Bai, Zhihao Yang, Zhaozhao Cai and Wei Sun
Sustainability 2025, 17(1), 306; https://doi.org/10.3390/su17010306 - 3 Jan 2025
Viewed by 379
Abstract
This study utilizes NASA’s Normalized Difference Vegetation Index (NDVI) data from the Google Earth Engine (GEE) platform and employs methods such as mean analysis, trend analysis, and the Hurst index to assess NDVI dynamics in Xinjiang, with a particular focus on desert, meadow, [...] Read more.
This study utilizes NASA’s Normalized Difference Vegetation Index (NDVI) data from the Google Earth Engine (GEE) platform and employs methods such as mean analysis, trend analysis, and the Hurst index to assess NDVI dynamics in Xinjiang, with a particular focus on desert, meadow, and grassland vegetation. Furthermore, multiple linear regression, random forest, support vector machines, and XGBoost models are applied to construct and evaluate the NDVI prediction models. The key driving forces are identified and ranked based on the results of the optimal model. Changes in the vegetation cover in response to these driving forces are analyzed using the Mann–Kendall test and partial correlation analysis. The results indicate the following: (1) From 2000 to 2023, the annual variation in NDVI in Xinjian fluctuates at a rate of 0.0012 per year. The intra-annual trend follows an inverted U shape, with meadow vegetation exhibiting the highest monthly NDVI fluctuations. (2) During this period, the annual average NDVI in Xinjiang ranges from 0 to 0.3, covering 74.74% of the region. Spatially, higher NDVI values are observed in the north and northwest, while lower values are concentrated in the south and southeast. (3) The overall slope of the variation in NDVI in Xinjiang between 2000 and 2023 ranges between −0.034 and 0.047, indicating no significant upward trend. According to the Hurst index, future projections suggest a shift from vegetation improvement to potential degradation. (4) Machine learning models are developed to predict NDVI, with random forest and XGBoost showing the highest precision. Soil moisture, runoff, and potential evaporation are identified as key drivers. In the last 24 years, the temperatures in Xinjiang have generally increased, while precipitation, soil moisture, and runoff have declined. There is a significant negative correlation between NDVI and both temperature and potential evaporation, while the correlation between NDVI and precipitation, soil moisture, and runoff is positive and significant, with distinct spatial variations throughout the region. The overall trend of vegetation cover in Xinjiang has been increasing, but the future outlook is less promising. Enhanced environmental monitoring and protective measures are essential moving forward. Full article
Show Figures

Figure 1

Figure 1
<p>Distribution of vegetation types of Xinjiang and area percentage charts.</p>
Full article ">Figure 2
<p>Overall flowchart of the research on NDVI changes and their drivers in Xinjiang.</p>
Full article ">Figure 3
<p>Interannual variation in NDVI during the growing season, 2000–2023.</p>
Full article ">Figure 4
<p>Intra-annual variation in monthly mean values of NDVI from 2000 to 2023.</p>
Full article ">Figure 5
<p>Distribution of annual mean values of NDVI for different vegetation types during the growing season from 2000 to 2023. (<b>a</b>) Xinjiang. (<b>b</b>) Desert vegetation. (<b>c</b>) Grassland vegetation. (<b>d</b>) Meadow vegetation.</p>
Full article ">Figure 6
<p>Distribution of slope and <span class="html-italic">p</span>-value for the growing season in Xinjiang from 2000 to 2023. (<b>a</b>) Slope. (<b>b</b>) <span class="html-italic">p</span>-value.</p>
Full article ">Figure 7
<p>Trend distribution of annual average values of NDVI during the growing season from 2000 to 2023. (<b>a</b>) Xinjiang. (<b>b</b>) Desert vegetation. (<b>c</b>) Grassland vegetation. (<b>d</b>) Meadow vegetation.</p>
Full article ">Figure 8
<p>Distribution of H for different vegetation types in Xinjiang from 2000 to 2023. (<b>a</b>) Xinjiang. (<b>b</b>) Desert vegetation. (<b>c</b>) Grassland vegetation. (<b>d</b>) Meadow vegetation.</p>
Full article ">Figure 9
<p>Future trends in vegetation cover in Xinjiang. (<b>a</b>) Xinjiang. (<b>b</b>) Desert vegetation. (<b>c</b>) Grassland vegetation. (<b>d</b>) Meadow vegetation.</p>
Full article ">Figure 10
<p>Future trends in area percentage of different vegetation types.</p>
Full article ">Figure 11
<p>Heat map of correlations between drivers.</p>
Full article ">Figure 12
<p>The importance of each influencing factor calculated by random forest and XGBoost. (<b>a</b>) Feature importance calculated by XGBoost. (<b>b</b>) Feature importance calculated by random forest.</p>
Full article ">Figure 13
<p>Trends in temperature, precipitation, soil moisture, runoff and potential evapotranspiration as a percentage of plot.</p>
Full article ">Figure 14
<p>Results of significance tests for partial correlation coefficients of air temperature, precipitation, soil moisture, runoff, and potential evaporation. (<b>a</b>) Potential evaporation. (<b>b</b>) Precipitation. (<b>c</b>) Runoff. (<b>d</b>) Soil moisture. (<b>e</b>) Temperature.</p>
Full article ">Figure A1
<p>Results of significance tests for air temperature, precipitation, soil moisture, runoff, and potential evaporation.</p>
Full article ">
25 pages, 4894 KiB  
Article
Unsteady Loading on a Tidal Turbine Due to the Turbulent Wake of an Upstream Turbine Interacting with a Seabed Ridge
by Sulaiman Hurubi, Hannah Mullings, Pablo Ouro, Peter Stansby and Tim Stallard
Energies 2025, 18(1), 151; https://doi.org/10.3390/en18010151 - 2 Jan 2025
Viewed by 300
Abstract
Tidal sites can present uneven seabed bathymetry features that induce favourable or adverse pressure gradients and are sources of turbulence, and so are likely to affect the operation, performance, and wake recovery dynamics of deployed tidal-stream turbines. Large-eddy simulations are conducted to analyse [...] Read more.
Tidal sites can present uneven seabed bathymetry features that induce favourable or adverse pressure gradients and are sources of turbulence, and so are likely to affect the operation, performance, and wake recovery dynamics of deployed tidal-stream turbines. Large-eddy simulations are conducted to analyse the unsteady loading of a tidal turbine subjected to the wake of an upstream turbine that interacts with a two-dimensional ridge located between the two turbines. Relative to an isolated turbine, blade fatigue loading is increased by up to 43% when subject to the wake of a turbine located 8 turbine diameters upstream interacting with a ridge located 2 turbine diameters upstream, whereas for the same spacing, the turbine wake led to a limited 6% reduction in loading and the ridge wake only caused a 79% increase. For larger spacings, the trends were similar, but the magnitude of difference reduced. Predictions of fatigue loads with a blade element momentum model (BEMT) provided a good agreement for flat bed conditions. However, the ridge-induced pressure gradient drives rapid spatial change of coherent flow structures, which limits the applicability of Taylor’s frozen turbulence hypothesis adopted in the BEMT. Reasonable prediction of rotor loading with BEMT was found to be obtained using the turbulent onset flow field at a plane one-diameter upstream of the turbine. This is more accurate than use of the planes at the rotor plane or two-diameters upstream, as coherent structures represent those modified by wake recovery and rotor induction in the approach flow to the turbine. Full article
Show Figures

Figure 1

Figure 1
<p>Representation of the computational domain includes a seabed ridge with dimensions expressed in turbine diameters <span class="html-italic">D</span>. The two-turbine configurations studied in the presence of a ridge are 6U8, 6U12, and 2U8. The same spacings between turbines are also tested for flat-bed conditions, labelled as F8 and F12.</p>
Full article ">Figure 2
<p>Mean streamwise velocity (<math display="inline"><semantics> <mrow> <mi>U</mi> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) distribution of the cases with a flat-bed (TF for a single turbine and F8 and F12 for two turbines). Horizontal (<b>left</b>) and vertical (<b>right</b>) contours across the turbine centre are shown with the origin of coordinates at the first-row turbine.</p>
Full article ">Figure 3
<p>Mean streamwise velocity (<math display="inline"><semantics> <mrow> <mi>U</mi> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) distribution of the ridge cases (6U and 2U for a single turbine and 6U8, 6U12 and 2U8 for two turbines). Horizontal (<b>left</b>) and vertical (<b>right</b>) contours across the turbine centre are shown with the origin of coordinates at the first-row turbine.</p>
Full article ">Figure 4
<p>Horizontal (<b>left</b>) and vertical (<b>right</b>) contours of normalised velocity deficit (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>U</mi> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) across turbine’s centre in the flat-bed configurations (TF for a single turbine and F8 and F12 for two turbines). The deficit in cases with two turbines is computed by subtracting the mean velocity field from the two-turbine simulation from that with the first-row turbine only. The streamwise origin of coordinates is taken at the first-row turbine rotor location.</p>
Full article ">Figure 5
<p>Horizontal (<b>left</b>) and vertical (<b>right</b>) contours of normalised velocity deficit (<math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi>U</mi> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) across turbine’s centre in the ridge configurations (6U and 2U for a single turbine and 6U8, 6U12, and 2U8 for two turbines). The deficit in cases with two turbines is computed by subtracting the mean velocity field from the two-turbine simulation from that with the first-row turbine only. The streamwise origin of coordinates is taken at the first-row turbine rotor location.</p>
Full article ">Figure 6
<p>Transverse profiles of the velocity deficit at hub height for the flat-bed cases (TF, F8, and F12) and two-turbine configurations (6U8, 6U12, and 2U8) compared with the theoretical Gaussian shape model. The data are normalised by the maximum velocity deficit, <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">Δ</mi> <msub> <mi>U</mi> <mi>max</mi> </msub> </mrow> </semantics></math>, and wake half-width, <math display="inline"><semantics> <msub> <mi>y</mi> <mrow> <mn>1</mn> <mo>/</mo> <mn>2</mn> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 7
<p>Streamwise evolution of the normalised disc-averaged (<b>a</b>) velocity (<math display="inline"><semantics> <mrow> <msub> <mi>U</mi> <mi>D</mi> </msub> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) and (<b>b</b>) turbulence intensity (<math display="inline"><semantics> <mrow> <msubsup> <mi>u</mi> <mi>D</mi> <mo>′</mo> </msubsup> <mo>/</mo> <msub> <mi>U</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) computed at hub height. Data from configurations with the first-row turbines only (dashed lines) and with second-row turbines (solid lines) are compared. Here, <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> <mo>=</mo> <mn>0</mn> </mrow> </semantics></math> corresponds to the rotor location of the second-row turbines. Shaded area and detail (in <b>a</b>) upstream of the second-row turbine represents ±1.5% range of the predicted induction velocities based on applying vortex sheet theory (Equations (<a href="#FD8-energies-18-00151" class="html-disp-formula">8</a>) and (<a href="#FD9-energies-18-00151" class="html-disp-formula">9</a>)) to the base-flow.</p>
Full article ">Figure 8
<p>Turbulence intensity contours in the <math display="inline"><semantics> <mrow> <mi>y</mi> <mi>z</mi> </mrow> </semantics></math>-plane from the base-flow simulations at the turbine locations of interest, with those in the top row being for configurations without upstream turbines and bottom row for scenarios where the upstream turbine is included. Dashed circles indicate the rotor swept area.</p>
Full article ">Figure 9
<p>Comparison of the disc-averaged values (markers) of (<b>a</b>) streamwise velocity and (<b>b</b>) turbulence intensity at turbine positions from the base-flow with their maximum and minimum values over the swept area. Square markers are used for cases 2D and 6D to highlight the absence of first-row turbines while circular markers are adopted for the other cases.</p>
Full article ">Figure 10
<p>(<b>a</b>) Power spectral density (PSD) of rotor thrust load with frequency normalised by the blade passing frequency <math display="inline"><semantics> <msub> <mi>f</mi> <mn>0</mn> </msub> </semantics></math> and (<b>b</b>) probability of exceedance (PoE) for the rotor thrust load normalised by the maximum load of the flat-bed case (<math display="inline"><semantics> <msub> <mi>T</mi> <mn>0</mn> </msub> </semantics></math>).</p>
Full article ">Figure 11
<p>Normalised Damage Equivalent Loads (DEL/DEL<sub>0</sub>) of blade thrust load in second-row turbines, obtained from LES–ALM (black) and BEMT calculated using inflow planes from the base-flow LES at the turbine position (red) and one diameter upstream (blue), normalised by the value of the first-row turbine in flat-bed conditions, DEL<sub>0</sub> = 0.980 N.</p>
Full article ">Figure 12
<p>Contours of two-point correlation of the streamwise velocity <math display="inline"><semantics> <msub> <mi>ρ</mi> <mrow> <mi>u</mi> <mi>u</mi> </mrow> </msub> </semantics></math> in the <math display="inline"><semantics> <mrow> <mi>y</mi> <mi>z</mi> </mrow> </semantics></math>-planes for cases TF and F8 (<b>top</b>), 6U and 6U12 (<b>middle</b>), and 2U and 2U8 (<b>bottom</b>). Dashed circles indicate the rotor swept area.</p>
Full article ">Figure 13
<p>Probability of exceedance (PoE) of the blade thrust loads normalised by the maximum blade thrust obtained for flat-bed conditions (<math display="inline"><semantics> <mrow> <mi>T</mi> <mo>/</mo> <msub> <mi>T</mi> <mn>0</mn> </msub> </mrow> </semantics></math>) from the LES–ALM (black) and BEMT (red) for cases (<b>a</b>) 6U8, (<b>b</b>) 6U12, (<b>c</b>) 2U8, and (<b>d</b>) F8.</p>
Full article ">Figure 14
<p>Power spectral density (PSD) of the blade thrust load from the LES–ALM (black) and BEMT (red) for cases (<b>a</b>) 6U8, (<b>b</b>) 6U12, (<b>c</b>) 2U8, and (<b>d</b>) F8. The frequency is normalised by the blade passing frequency <math display="inline"><semantics> <msub> <mi>f</mi> <mn>0</mn> </msub> </semantics></math>.</p>
Full article ">Figure 15
<p>Comparison of the power spectral density (PSD) profiles of the streamwise (<b>a</b>,<b>b</b>) and spanwise (<b>c</b>,<b>d</b>) velocity time series obtained from LES at one diameter upstream of the second-row turbine location, with and without its presence compared to the ridge only case (2DR). (<b>a</b>,<b>c</b>) 6U12 and 6U cases. (<b>b</b>,<b>d</b>) 2U8 and 2U cases. The frequency is normalised by the blade passing frequency <math display="inline"><semantics> <msub> <mi>f</mi> <mn>0</mn> </msub> </semantics></math> and the location <math display="inline"><semantics> <mrow> <mi>x</mi> <mo>/</mo> <mi>D</mi> </mrow> </semantics></math> of the sampling points is relative to the first-row turbine.</p>
Full article ">Figure A1
<p>Reynolds shear stress contours in the <math display="inline"><semantics> <mrow> <mi>y</mi> <mi>z</mi> </mrow> </semantics></math>-plane from the base-flow simulations at turbine locations, dashed circles indicates the rotor area. The top row shows cases with no upstream turbine, and the bottom row depicts scenarios where an upstream turbine is present.</p>
Full article ">
19 pages, 6481 KiB  
Article
Roughness Evaluation of Bamboo Surfaces Created by Abrasive Belt Sanding
by Jian Zhang, Yunhao Cui, Haibin Yang, Liuting Wang and Jun Qian
Forests 2025, 16(1), 66; https://doi.org/10.3390/f16010066 - 2 Jan 2025
Viewed by 255
Abstract
Mechanical belt sanding is critical in the manufacturing of bamboo and bamboo products, where surface roughness is commonly used to quantitatively evaluate the surface quality. In this study, flattened bamboo workpieces were sanded using P80 and P120 abrasive belts to create different surfaces. [...] Read more.
Mechanical belt sanding is critical in the manufacturing of bamboo and bamboo products, where surface roughness is commonly used to quantitatively evaluate the surface quality. In this study, flattened bamboo workpieces were sanded using P80 and P120 abrasive belts to create different surfaces. The linear roughness parameters, namely Rz, Ra, Rq, Rsk, Rku, and Rmr(c), were measured using both a stylus profilometer and a 3D profilometer. Statistical t-tests were conducted to determine the significance of differences between the two methods. Additionally, roughness profiles were analyzed in the frequency domain using Fast Fourier Transform (FFT) and Power Spectral Density (PSD) methods. A Random Forest (RF) regression model was also developed to predict the roughness values and figure out the dominant factors between granularity and measurement methods. The results revealed that both the stylus and 3D profilometers provided reliable comparisons of Rz, Ra, Rq, and Rmr (50%) for different grit sizes. However, resolution differences between the two methods were found to be critical for accurately interpreting roughness values. Variations in Rsk and Rku highlighted differences in sensitivity and detection range, particularly at finer scales, between the two methods. The stylus profilometer, with its higher spatial resolution and finer sampling density, demonstrated greater sensitivity to finer surface details. This was consistent with the FFT and PSD analyses, which showed that the stylus profilometer captured higher-frequency surface components more effectively. Furthermore, the RF model indicated that the choice of measurement method had negligible impact on the evaluation of the selected roughness parameters, suggesting that standardizing measurement techniques may not be essential for consistent roughness assessments of sanded bamboo surfaces. Full article
(This article belongs to the Section Wood Science and Forest Products)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of the experimental setup.</p>
Full article ">Figure 2
<p>Illustration of surface roughness measurement by stylus profilometer and 3D profilometer.</p>
Full article ">Figure 3
<p>Illustration of the Abbott curve. Note: the dashed lines colored in blue and red refer to different profile heights between the the peak height and the valley bottom depth.</p>
Full article ">Figure 4
<p>Typical roughness profiles obtained from: (<b>a</b>,<b>b</b>) stylus profilometer; (<b>c</b>,<b>d</b>) 3D profilometer.</p>
Full article ">Figure 5
<p>Results of surface roughness parameters selected in this study.</p>
Full article ">Figure 6
<p>Abbott curve for roughness profiles obtained from stylus profilometer.</p>
Full article ">Figure 7
<p>FFT spectra of typical roughness profiles obtained from: (<b>a</b>,<b>b</b>) stylus profilometer; (<b>c</b>,<b>d</b>) 3D profilometer.</p>
Full article ">Figure 8
<p>FFT spectra of typical roughness profiles obtained from: (<b>a</b>,<b>b</b>) stylus profilometer; (<b>c</b>,<b>d</b>) 3D profilometer. Note: the red dashed lines refer to the position of 10<sup>6</sup> <math display="inline"><semantics> <mrow> <mi mathvariant="sans-serif">μ</mi> </mrow> </semantics></math>m<sup>2</sup>·mm, which aims to enhance comparability.</p>
Full article ">Figure 9
<p>Feature importances of grit size and measuring method on surface roughness.</p>
Full article ">
19 pages, 2360 KiB  
Article
Deep-Learning-Driven Insights into Nitrogen Leaching for Sustainable Land Use and Agricultural Practices
by Caixia Hu, Jie Li, Yaxu Pang, Lan Luo, Fang Liu, Wenhao Wu, Yan Xu, Houyu Li, Bingcang Tan and Guilong Zhang
Land 2025, 14(1), 69; https://doi.org/10.3390/land14010069 - 2 Jan 2025
Viewed by 276
Abstract
Nitrate leaching from soil presents a significant threat to soil health, as it can result in nutrient loss, soil acidification, and structural damage. It is crucial to quantify the spatial heterogeneity of nitrate leaching and its drivers. A total of 509 observational data [...] Read more.
Nitrate leaching from soil presents a significant threat to soil health, as it can result in nutrient loss, soil acidification, and structural damage. It is crucial to quantify the spatial heterogeneity of nitrate leaching and its drivers. A total of 509 observational data points regarding nitrate leaching in northern China were collected, capturing the spatial and temporal variations across crops such as winter wheat, maize, and greenhouse vegetables. A machine learning (ML) model for predicting nitrate leaching was then developed, with the random forest (RF) model outperforming the support vector machine (SVM), extreme gradient boosting (XGBoost), and convolutional neural network (CNN) models, achieving an R2 of 0.75. However, the performance improved significantly after integrating the four models with Bayesian optimization (all models had R2 > 0.56), which realized quantitative prediction capabilities for nitrate leaching loss concentrations. Moreover, the XGBoost model exhibited the highest fitting accuracy and the smallest error in estimating nitrate leaching losses, with an R2 value of 0.79 and an average absolute error (MAE) of 3.87 kg/ha. Analyses of the feature importance and SHAP values in the optimal XGBoost model identified soil organic matter, chemical nitrogen fertilizer input, and water input (including rainfall and irrigation) as the main indicators of nitrate leaching loss. The ML-based modeling method developed overcomes the difficulty of the determination of the functional relationship between nitrate loss intensity and its influencing factors, providing a data-driven solution for estimating nitrate–nitrogen loss in farmlands in North China and strengthening sustainable agricultural practices. Full article
Show Figures

Figure 1

Figure 1
<p>Raw data and processed data QQ plots of average annual temperature (<b>A</b>), average annual rainfall (<b>B</b>), soil type (<b>C</b>), chemical N fertilizer input (<b>D</b>), organic N fertilizer input (<b>E</b>), irrigation amount (<b>F</b>), irrigation methods (<b>G</b>), soil total N (<b>H</b>), soil organic matter (<b>I</b>), soil pH (<b>J</b>), soil bulk density (<b>K</b>) and soil clay (<b>L</b>).</p>
Full article ">Figure 2
<p>Pearson’s correlation matrix of independent variables.</p>
Full article ">Figure 3
<p>Comparison of R<sup>2</sup>, <span class="html-italic">RMSE</span>, and <span class="html-italic">MAE</span> using the SVM (<b>A</b>), RF (<b>B</b>), XGBoost (<b>C</b>), and CNN (<b>D</b>) models for nitrate–nitrogen loss rate prediction on training and test datasets. Abbreviations: SVM, support vector machine; RF, random forest; XGBoost, extreme gradient boosting; CNN, convolutional neural network. Moving average error (<span class="html-italic">MAE</span>), root mean square error (<span class="html-italic">RMSE</span>), and coefficient of determination (R<sup>2</sup>).</p>
Full article ">Figure 4
<p>Result of Bayesian-optimized hyperparameters in SVM (<b>A</b>), RF (<b>B</b>), XGBoost (<b>C</b>), and CNN (<b>D</b>) models for nitrate–nitrogen loss rate prediction. Abbreviations: SVM, support vector machine; RF, random forest; XGBoost, extreme gradient boosting; CNN, convolutional neural network. Moving average error (<span class="html-italic">MAE</span>), root mean square error (<span class="html-italic">RMSE</span>), and coefficient of determination (R<sup>2</sup>).</p>
Full article ">Figure 5
<p>Ranking of the importance of input features (<b>A</b>) and the SHAP value for a particular variable (<b>B</b>).</p>
Full article ">
Back to TopTop