[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

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

Article Types

Countries / Regions

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

Search Results (21,897)

Search Parameters:
Keywords = rivers

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 5161 KiB  
Article
Underwater Acoustic Orthogonal Frequency-Division Multiplexing Communication Using Deep Neural Network-Based Receiver: River Trial Results
by Sabna Thenginthody Hassan, Peng Chen, Yue Rong and Kit Yan Chan
Sensors 2024, 24(18), 5995; https://doi.org/10.3390/s24185995 (registering DOI) - 15 Sep 2024
Abstract
In this article, a deep neural network (DNN)-based underwater acoustic (UA) communication receiver is proposed. Conventional orthogonal frequency-division multiplexing (OFDM) receivers perform channel estimation using linear interpolation. However, due to the significant delay spread in multipath UA channels, the frequency response often exhibits [...] Read more.
In this article, a deep neural network (DNN)-based underwater acoustic (UA) communication receiver is proposed. Conventional orthogonal frequency-division multiplexing (OFDM) receivers perform channel estimation using linear interpolation. However, due to the significant delay spread in multipath UA channels, the frequency response often exhibits strong non-linearity between pilot subcarriers. Since the channel delay profile is generally unknown, this non-linearity cannot be modeled precisely. A neural network (NN)-based receiver effectively tackles this challenge by learning and compensating for the non-linearity through NN training. The performance of the DNN-based UA communication receiver was tested recently in river trials in Western Australia. The results obtained from the trials prove that the DNN-based receiver performs better than the conventional least-squares (LS) estimator-based receiver. This paper suggests that UA communication using DNN receivers holds great potential for revolutionizing underwater communication systems, enabling higher data rates, improved reliability, and enhanced adaptability to changing underwater conditions. Full article
(This article belongs to the Special Issue Advanced Acoustic Sensing Technology)
20 pages, 20184 KiB  
Article
Snow Cover Extraction from Landsat 8 OLI Based on Deep Learning with Cross-Scale Edge-Aware and Attention Mechanism
by Zehao Yu, Hanying Gong, Shiqiang Zhang and Wei Wang
Remote Sens. 2024, 16(18), 3430; https://doi.org/10.3390/rs16183430 (registering DOI) - 15 Sep 2024
Abstract
Snow cover distribution is of great significance for climate change and water resource management. Current deep learning-based methods for extracting snow cover from remote sensing images face challenges such as insufficient local detail awareness and inadequate utilization of global semantic information. In this [...] Read more.
Snow cover distribution is of great significance for climate change and water resource management. Current deep learning-based methods for extracting snow cover from remote sensing images face challenges such as insufficient local detail awareness and inadequate utilization of global semantic information. In this study, a snow cover extraction algorithm integrating cross-scale edge perception and an attention mechanism on the U-net model architecture is proposed. The cross-scale edge perception module replaces the original jump connection of U-net, enhances the low-level image features by introducing edge detection on the shallow feature scale, and enhances the detail perception via branch separation and fusion features on the deep feature scale. Meanwhile, parallel channel and spatial attention mechanisms are introduced in the model encoding stage to adaptively enhance the model’s attention to key features and improve the efficiency of utilizing global semantic information. The method was evaluated on the publicly available CSWV_S6 optical remote sensing dataset, and the accuracy of 98.14% indicates that the method has significant advantages over existing methods. Snow extraction from Landsat 8 OLI images of the upper reaches of the Irtysh River was achieved with satisfactory accuracy rates of 95.57% (using two, three, and four bands) and 96.65% (using two, three, four, and six bands), indicating its strong potential for automated snow cover extraction over larger areas. Full article
Show Figures

Figure 1

Figure 1
<p>True color CSWV_S6 data synthesized from the red, green, and blue bands (the numbering in the figure corresponds to the original naming in the acquired files).</p>
Full article ">Figure 2
<p>(<b>a</b>) RGB composite of Landsat 8 imagery (red: band 4, green: band 3, blue: band 2). (<b>b</b>) Land cover types.</p>
Full article ">Figure 3
<p>CEFCSAU-net network model architecture. The input size was (512, 512, C), where C denotes the number of channels, and experiments in this paper utilized either 3 or 4; during the model’s operation on a GPU, intermediate feature maps were stored as tensors.</p>
Full article ">Figure 4
<p>Attention mechanism module for channel and space mixing. Here, (H, W, C) represent the height, width, and number of channels of the feature data, respectively, with values determined by input features at different stages. CF and CF’ denote feature maps from various intermediate operations within the channel attention mechanism. Cat Sf, Sf, Sf’ represent feature maps from different intermediate operations of the spatial attention mechanism. The SA feature denotes the feature map post-spatial attention mechanism, the CA feature represents those post-channel attention mechanisms, and the CSA feature illustrates feature maps following the CSA module.</p>
Full article ">Figure 5
<p>Cross-scale edge-aware feature fusion module. Sobelx F, Sobely F, and Laplacian F denote feature maps resulting from various edge detection operations. Shallow F refers to feature maps following shallow feature convolution. Fusion F illustrates feature maps resulting from the fusion of shallow and deep features. Deep F’ represents feature maps after a series of operations on deep features.</p>
Full article ">Figure 6
<p>Snow extraction results of CSWV_S6 data on different segmentation models (a set of two rows is the same image data, and rows two, four, and six are zoomed-in images of local details corresponding to rows one, three, and five. The blue area is snow, the white area is non-snow, and the red area is false detection).</p>
Full article ">Figure 7
<p>Snow extraction results from different deep learning models for Landsat 8 OLI imagery (blue areas are snow, white areas are non-snow, and red areas are false detections).</p>
Full article ">Figure 8
<p>(<b>a</b>) CSWV_S6 test set scores on different models for each type of metrics and (<b>b</b>) Landsat 8 OLI test set scores for various metrics on different models.</p>
Full article ">Figure 9
<p>Score results of the three CSWV_S6 test sets’ example data on the evaluation metrics on each model, with 0.08% of snow image elements in the first row of data, 0.95% of snow image elements in the second row of data, and 1.73% of data in the third row of data.</p>
Full article ">Figure 10
<p>Map of CEFCSAU-net model’s snow extraction in the cloud–snow confusion scenario of CSWV_S6 test set.</p>
Full article ">Figure 11
<p>Heat map comparing the mean values of ablation experiments on the test set with different data sets: (<b>a</b>) CSWV_6 dataset, (<b>b</b>) Landsat8 OLI dataset.</p>
Full article ">Figure 12
<p>(<b>a</b>) Input data image; (<b>b</b>) feature map of the first 8 channels before the intermediate feature data first pass through the CSA module; (<b>c</b>) feature map of the first 8 channels after the feature data first pass through the CSA module.</p>
Full article ">Figure 13
<p>This figure displays average feature maps following skip connections at various stages under different configurations of the CEFCSAU-net model. In this figure, (<b>a1</b>–<b>a4</b>) represent the model configuration without both the CSA and CEF modules; (<b>b1</b>–<b>b4</b>) indicate configurations without the CSA module yet including the CEF module; and (<b>c1</b>–<b>c4</b>) depict configurations featuring both CSA and CEF modules. The dimensions of the four columns of feature maps are sequentially 512 × 512, 256 × 256, 128 × 128, and 64 × 64.</p>
Full article ">
17 pages, 9162 KiB  
Article
Estimating Cadmium Concentration in Agricultural Soils with ZY1-02D Hyperspectral Data: A Comparative Analysis of Spectral Transformations and Machine Learning Models
by Junwei Lv, Jing Geng, Xuanhong Xu, Yong Yu, Huajun Fang, Yifan Guo and Shulan Cheng
Agriculture 2024, 14(9), 1619; https://doi.org/10.3390/agriculture14091619 (registering DOI) - 15 Sep 2024
Abstract
The accumulation of cadmium (Cd) in agricultural soils presents a significant threat to crop safety, emphasizing the critical necessity for effective monitoring and management of soil Cd levels. Despite technological advancements, accurately monitoring soil Cd concentrations using satellite hyperspectral technology remains challenging, particularly [...] Read more.
The accumulation of cadmium (Cd) in agricultural soils presents a significant threat to crop safety, emphasizing the critical necessity for effective monitoring and management of soil Cd levels. Despite technological advancements, accurately monitoring soil Cd concentrations using satellite hyperspectral technology remains challenging, particularly in efficiently extracting spectral information. In this study, a total of 304 soil samples were collected from agricultural soils surrounding a tungsten mine located in the Xiancha River basin, Jiangxi Province, Southern China. Leveraging hyperspectral data from the ZY1-02D satellite, this research developed a comprehensive framework that evaluates the predictive accuracy of nine spectral transformations across four modeling approaches to estimate soil Cd concentrations. The spectral transformation methods included four logarithmic and reciprocal transformations, two derivative transformations, and three baseline correction and normalization transformations. The four models utilized for predicting soil Cd were partial least squares regression (PLSR), support vector machine (SVM), bidirectional recurrent neural networks (BRNN), and random forest (RF). The results indicated that these spectral transformations markedly enhanced the absorption and reflection features of the spectral curves, accentuating key peaks and troughs. Compared to the original spectral curves, the correlation analysis between the transformed spectra and soil Cd content showed a notable improvement, particularly with derivative transformations. The combination of the first derivative (FD) transformation with the RF model yielded the highest accuracy (R2 = 0.61, RMSE = 0.37 mg/kg, MAE = 0.21 mg/kg). Furthermore, the RF model in multiple spectral transformations exhibited higher suitability for modeling soil Cd content compared to other models. Overall, this research highlights the substantial applicative potential of the ZY1-02D satellite hyperspectral data for detecting soil heavy metals and provides a framework that integrates optimal spectral transformations and modeling techniques to estimate soil Cd contents. Full article
(This article belongs to the Section Digital Agriculture)
Show Figures

Figure 1

Figure 1
<p>Distribution of soil sampling sites in the study area: (<b>a</b>) Jiangxi Province, China; (<b>b</b>) Geographic location of the study area; (<b>c</b>) Distribution of sampling points and elevation within the study area. The top-right image shows the coverage of the study area by the original ZY1-02D imagery.</p>
Full article ">Figure 2
<p>(<b>a</b>) Original spectral curves and (<b>b</b>) Savitzky–Golay (SG) smoothed spectral curves of soil samples from hyperspectral images. Note: Each color represents a sampling point.</p>
Full article ">Figure 3
<p>The correlation coefficients between soil Cd and original soil spectral data, and after Savitzky–Golay (SG) smoothed spectral data.</p>
Full article ">Figure 4
<p>Nine spectral transformation curves of soil samples from hyperspectral images. (<b>a</b>) logarithmic transformation (LT), (<b>b</b>) reciprocal transformation (RT), (<b>c</b>) first derivative (FD), (<b>d</b>) logarithm of reciprocal transformation (LR), (<b>e</b>) reciprocal of logarithmic transformation (RL), (<b>f</b>) reciprocal of logarithmic and first derivative (RLFD), (<b>g</b>) standard normal variate (SNV), (<b>h</b>) continuum removal (CR), and (<b>i</b>) multiplicative scatter correction (MSC). Note: Each color represents a sampling point.</p>
Full article ">Figure 5
<p>The correlation coefficient curves between the spectra derived from nine spectral transformation methods and the soil Cd content. (<b>a</b>) logarithmic transformation (LT), (<b>b</b>) reciprocal transformation (RT), (<b>c</b>) first derivative (FD), (<b>d</b>) logarithm of reciprocal transformation (LR), (<b>e</b>) reciprocal of logarithmic transformation (RL), (<b>f</b>) reciprocal of logarithmic and first derivative (RLFD), (<b>g</b>) standard normal variate (SNV), (<b>h</b>) continuum removal (CR), and (<b>i</b>) multiplicative scatter correction (MSC).</p>
Full article ">Figure 6
<p>Spatial distribution of soil Cd content in the study area driven by the RF model constructed with first derivative-transformed spectral data. Note that this Cd distribution map has been masked with a cropland layer derived from the GlobeLand30 dataset (<a href="http://www.globallandcover.com/" target="_blank">http://www.globallandcover.com/</a>, accessed on 20 December 2022).</p>
Full article ">Figure 7
<p>Relative proportional and spatial extents of three soil pollution categories based on soil Cd contents.</p>
Full article ">
15 pages, 19478 KiB  
Article
Source Apportionment and Human Health Risks of Potentially Toxic Elements in the Surface Water of Coal Mining Areas
by Yuting Yan, Yunhui Zhang, Zhan Xie, Xiangchuan Wu, Chunlin Tu, Qingsong Chen and Lanchu Tao
Toxics 2024, 12(9), 673; https://doi.org/10.3390/toxics12090673 (registering DOI) - 15 Sep 2024
Abstract
Contamination with potentially toxic elements (PTEs) frequently occurs in surface water in coal mining areas. This study analyzed 34 surface water samples collected from the Yunnan–Guizhou Plateau for their hydrochemical characteristics, spatial distribution, source apportionment, and human health risks. Our statistical analysis showed [...] Read more.
Contamination with potentially toxic elements (PTEs) frequently occurs in surface water in coal mining areas. This study analyzed 34 surface water samples collected from the Yunnan–Guizhou Plateau for their hydrochemical characteristics, spatial distribution, source apportionment, and human health risks. Our statistical analysis showed that the average concentrations of PTEs in the surface water ranked as follows: Fe > Al > Zn > Mn > Ba > B> Ni > Li > Cd > Mo > Cu > Co > Hg > Se > As > Pb > Sb. The spatial analysis revealed that samples with high concentrations of Fe, Al, and Mn were predominantly distributed in the main stream, Xichong River, and Yangchang River. Positive matrix factorization (PMF) identified four sources of PTEs in the surface water. Hg, As, and Se originated from wastewater discharged by coal preparation plants and coal mines. Mo, Li, and B originated from the dissolution of clay minerals in coal seams. Elevated concentrations of Cu, Fe, Al, Mn, Co, and Ni were attributed to the dissolution of kaolinite, illite, chalcopyrite, pyrite, and minerals associated with Co and Ni in coal seams. Cd, Zn, and Pb were derived from coal melting and traffic release. The deterministic health risks assessment showed that 94.12% of the surface water samples presented non-carcinogenic risks below the health limit of 1. Meanwhile, 73.56% of the surface water samples with elevated As posed level III carcinogenic risk to the local populations. Special attention to drinking water safety for children is warranted due to their lower metabolic capacity for detoxifying PTEs. This study provides insight for PTE management in sustainable water environments. Full article
(This article belongs to the Section Metals and Radioactive Substances)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Yunnan–Guizhou Plateau in China, (<b>b</b>) location of the study area in the Yunnan–Guizhou Plateau, and (<b>c</b>) location of surface water, groundwater, and mine water sampling sites in the study area (sample size = 34).</p>
Full article ">Figure 2
<p>Box plots with the standard of potentially toxic elements for drinking surface water (sample size = 34).</p>
Full article ">Figure 3
<p>Spatial distribution map of concentrations of potentially toxic elements: (<b>a</b>) Fe, (<b>b</b>) Mn, (<b>c</b>) Cu, (<b>d</b>) Zn, (<b>e</b>) Al, (<b>f</b>) Hg, (<b>g</b>) As, (<b>h</b>) Se, (<b>i</b>) Cd, (<b>j</b>) Pb, (<b>k</b>) Li, (<b>l</b>) B, (<b>m</b>) Ba, (<b>n</b>) Sb, (<b>o</b>) Ni, (<b>p</b>) Co, and (<b>q</b>) Mo (sample size = 34).</p>
Full article ">Figure 4
<p>Source contributions of PTEs based on the PMF model: (<b>a</b>) relative contributions of PTEs to PMF factors and (<b>b</b>) average contributions of PMF factors (sample size = 34).</p>
Full article ">Figure 5
<p>Non-carcinogenic health risks of surface water to children, men, and women: (<b>a</b>) Fe, (<b>b</b>) Mn, (<b>c</b>) Cu, (<b>d</b>) Zn, (<b>e</b>) Al, (<b>f</b>) As, (<b>g</b>) Se, (<b>h</b>) Cd, (<b>i</b>) Li, (<b>j</b>) B, (<b>k</b>) Ba, (<b>l</b>) Sb, (<b>m</b>) Ni, (<b>n</b>) Co, and (<b>o</b>) Mo (sample size = 34).</p>
Full article ">Figure 6
<p>Sensitive non-carcinogenic PTE ranking for HI.</p>
Full article ">Figure 7
<p>Non-carcinogenic and carcinogenic health risks of PTEs in surface water to children, men, and women: (<b>a</b>) hazard index (HI) and (<b>b</b>) cancer risk (CR) (sample size = 34).</p>
Full article ">Figure 8
<p>Spatial distribution map of hazard index (HI) and cancer risk (CR) of PTEs in surface water. HI to (<b>a</b>) children, (<b>b</b>) men, and (<b>c</b>) women and CR to (<b>d</b>) children, (<b>e</b>) men, and (<b>f</b>) women (sample size = 34).</p>
Full article ">
24 pages, 5994 KiB  
Article
Mapping Natural Populus euphratica Forests in the Mainstream of the Tarim River Using Spaceborne Imagery and Google Earth Engine
by Jiawei Zou, Hao Li, Chao Ding, Suhong Liu and Qingdong Shi
Remote Sens. 2024, 16(18), 3429; https://doi.org/10.3390/rs16183429 (registering DOI) - 15 Sep 2024
Abstract
Populus euphratica is a unique constructive tree species within riparian desert areas that is essential for maintaining oasis ecosystem stability. The Tarim River Basin contains the most densely distributed population of P. euphratica forests in the world, and obtaining accurate distribution data in [...] Read more.
Populus euphratica is a unique constructive tree species within riparian desert areas that is essential for maintaining oasis ecosystem stability. The Tarim River Basin contains the most densely distributed population of P. euphratica forests in the world, and obtaining accurate distribution data in the mainstream of the Tarim River would provide important support for its protection and restoration. We propose a new method for automatically extracting P. euphratica using Sentinel-1 and 2 and Landsat-8 images based on the Google Earth Engine cloud platform and the random forest algorithm. A mask of the potential distribution area of P. euphratica was created based on prior knowledge to save computational resources. The NDVI (Normalized Difference Vegetation Index) time series was then reconstructed using the preferred filtering method to obtain phenological parameter features, and the random forest model was input by combining the phenological parameter, spectral index, textural, and backscattering features. An active learning method was employed to optimize the model and obtain the best model for extracting P. euphratica. Finally, the map of natural P. euphratica forests with a resolution of 10 m in the mainstream of the Tarim River was obtained. The overall accuracy, producer’s accuracy, user’s accuracy, kappa coefficient, and F1-score of the map were 0.96, 0.98, 0.95, 0.93, and 0.96, respectively. The comparison experiments showed that simultaneously adding backscattering and textural features improved the P. euphratica extraction accuracy, while textural features alone resulted in a poor extraction effect. The method developed in this study fully considered the prior and posteriori information and determined the feature set suitable for the P. euphratica identification task, which can be used to quickly obtain accurate large-area distribution data of P. euphratica. The method can also provide a reference for identifying other typical desert vegetation. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location of the study area and the distribution of sample points. (<b>a</b>): location of the study area in Xinjiang province in China; (<b>b</b>): training dataset distribution; (<b>c</b>): detailed sample area showing <span class="html-italic">P. euphratica</span> and non–<span class="html-italic">P. euphratica</span> in a Sentinel-2 false-color image.</p>
Full article ">Figure 2
<p>Distribution of validation dataset. The black solid line represents the range of the study area; the red and yellow points represent <span class="html-italic">P. euphratica</span> and non–<span class="html-italic">P. euphratica</span>, respectively.</p>
Full article ">Figure 3
<p>Workflow of the research.</p>
Full article ">Figure 4
<p>Threshold segmentation effect of MNDWI and NDVI. (<b>a</b>): false color image of Jieran Lik Reservoir in Xinjiang Province; (<b>b</b>): statistical result of the corresponding frequency distribution of MNDWI values of water and other ground objects in area (<b>a</b>); (<b>c</b>): false color image of Pazili Tamu in Xinjiang; (<b>d</b>): statistical result for the corresponding frequency distribution of NDVI values of desert bare land and other ground objects in region (<b>c</b>).</p>
Full article ">Figure 5
<p>Comparison of NDVI data before and after spatiotemporal fusion: (<b>a</b>) NDVI data derived from Sentinel-2 before fusion, (<b>b</b>) NDVI data after fusion.</p>
Full article ">Figure 6
<p>Comparison of the effects of different filter functions for: (<b>a</b>) <span class="html-italic">P. euphratica</span>; (<b>b</b>) <span class="html-italic">Tamarix</span>; (<b>c</b>) allee tree; (<b>d</b>) farmland; (<b>e</b>) wetland; (<b>f</b>) urban tree.</p>
Full article ">Figure 7
<p>Comparison between phenological curves of six typical vegetation species. Phenology parameters of (<b>a</b>) <span class="html-italic">P. euphratica</span>, (<b>b</b>) <span class="html-italic">Tamarix</span>, (<b>c</b>) allee tree, (<b>d</b>) farmland, (<b>e</b>) wetland, and (<b>f</b>) urban tree.</p>
Full article ">Figure 8
<p>Importance of different features in the RF classification.</p>
Full article ">Figure 9
<p>Natural <span class="html-italic">P. euphratica</span> forest maps extracted using four feature combinations: (<b>a</b>) PS, (<b>b</b>) PSB, (<b>c</b>) PST, and (<b>d</b>) PSBT.</p>
Full article ">Figure 10
<p>Comparison of <span class="html-italic">P. euphratica</span> extraction results using different feature combinations on Sentinel-2 standard false color images. Rows 1 to 4 show the identification of <span class="html-italic">P. euphratica</span> in desert areas, <span class="html-italic">P. euphratica</span>-dense areas, agricultural areas, and large river areas, respectively. The green area represents the classification result of <span class="html-italic">P. euphratica</span>. The yellow circle corresponding to each row is the area where the extraction results of different feature combinations are quite different.</p>
Full article ">Figure 11
<p>(<b>a</b>) Distribution of natural <span class="html-italic">P. euphratica</span> forest in the mainstream of the Tarim River. (<b>b</b>): UAV image of healthy <span class="html-italic">P. euphratica</span>, (<b>c</b>): classification result of healthy <span class="html-italic">P. euphratica</span>, (<b>d</b>): UAV image of unhealthy <span class="html-italic">P. euphratica</span>, (<b>e</b>): classification result of unhealthy <span class="html-italic">P. euphratica</span>, (<b>f</b>): UAV image of dense <span class="html-italic">P. euphratica</span>, (<b>g</b>): classification result of dense <span class="html-italic">P. euphratica</span>, (<b>h</b>): UAV image of sparse <span class="html-italic">P. euphratica</span>, (<b>i</b>): classification result of sparse <span class="html-italic">P. euphratica</span>. The green area represents the classification results of <span class="html-italic">P. euphratica</span>.</p>
Full article ">Figure 12
<p>Mixed pixel problems associated with <span class="html-italic">P. euphratica</span>: (<b>a</b>) <span class="html-italic">P. euphratica</span> occupying less than one pixel; (<b>b</b>) sandy soil interfering with the reflected signal of <span class="html-italic">P. euphratica</span>. The red box represents a pixel on the images for clearer observation. Basemaps of row 1-2 are UAV images while row 3 are Sentinel-2 standard false color images.</p>
Full article ">
18 pages, 6210 KiB  
Article
Research on Glacier Changes and Their Influencing Factors in the Yigong Zangbo River Basin of the Tibetan Plateau, China, Based on ICESat-2 Data
by Wei Nie, Qiqi Du, Xuepeng Zhang, Kunxin Wang, Yang Liu, Yongjie Wang, Peng Gou, Qi Luo and Tianyu Zhou
Water 2024, 16(18), 2617; https://doi.org/10.3390/w16182617 (registering DOI) - 15 Sep 2024
Abstract
The intense changes in glaciers in the southeastern Tibetan Plateau (SETP) have essential impacts on regional water resource management. In order to study the seasonal fluctuations of glaciers in this region and their relationship with climate change, we focus on the Yigong Zangbo [...] Read more.
The intense changes in glaciers in the southeastern Tibetan Plateau (SETP) have essential impacts on regional water resource management. In order to study the seasonal fluctuations of glaciers in this region and their relationship with climate change, we focus on the Yigong Zangbo River Basin in the SETP, extract the annual and seasonal variations of glaciers in the basin during 2018–2023, and analyze their spatio-temporal characteristics through the seasonal-trend decomposition using the LOESS (STL) method. Finally, combining the Extreme Gradient Boosting (XGBoost) model and the Shapley additive explanations (SHAP) model, we assess the comprehensive impact of meteorological factors such as temperature and snowfall on glacier changes. The results indicate that glaciers in the Yigong Zangbo River Basin experienced remarkable mass loss during 2018–2023, with an average annual melting rate of −0.83 ± 0.12 m w.e.∙yr−1. The glacier mass exhibits marked seasonal fluctuations, with increases in January–March (JFM) and April–June (AMJ) and noticeable melting in July–September (JAS) and October–December (OND). The changes over these four periods are 2.12 ± 0.04 m w.e., 0.93 ± 0.15 m w.e., −1.58 ± 0.19 m w.e., and −1.32 ± 0.17 m w.e., respectively. Temperature has been identified as the primary meteorological driver of glacier changes in the study area, surpassing the impact of snowfall. This study uses advanced altimetry data and meteorological data to monitor and analyze glacier changes, which provides valuable data for cryosphere research and also validates a set of replicable research methods, which provides support for future research in related fields. Full article
(This article belongs to the Section Water and Climate Change)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Overview of the study area, (<b>b</b>) distribution of the ICESat-2 points in the study area, and (<b>c</b>) the location of the study area on the Tibetan Plateau.</p>
Full article ">Figure 2
<p>Technical flowchart. “GMC” denotes glacier mass changes, “DEM” the digital elevation model, “RGI” the Randolph Glacier Inventory data, “STL” the seasonal-trend decomposition using LOESS, and “SHAP” the Shapley additive explanations.</p>
Full article ">Figure 3
<p>A schematic diagram of the XGBoost model. The red arrow denotes the selected direction of the tree, the yellow circle denotes the selected node, the green circle denotes the unselected node.</p>
Full article ">Figure 4
<p>Glacier mass change (GMC), seasonal mass difference (<math display="inline"><semantics> <mrow> <mo>∆</mo> <msub> <mrow> <mi mathvariant="normal">M</mi> </mrow> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">L</mi> </mrow> </msub> </mrow> </semantics></math>), and trend and seasonal components after the STL decomposition.</p>
Full article ">Figure 5
<p>Deseasonalized GMC series X <math display="inline"><semantics> <mrow> <mo>(</mo> <msub> <mrow> <mi mathvariant="normal">X</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>=</mo> <msub> <mrow> <mi mathvariant="normal">T</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>+</mo> <msub> <mrow> <mi mathvariant="sans-serif">ε</mi> </mrow> <mrow> <mi mathvariant="normal">i</mi> </mrow> </msub> <mo>)</mo> </mrow> </semantics></math>.</p>
Full article ">Figure 6
<p>(<b>a</b>) Average seasonal GMC relative to ALOS DEM (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi mathvariant="normal">G</mi> <mi mathvariant="normal">M</mi> <mi mathvariant="normal">C</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) and glacial area with altitude, and (<b>b</b>) average seasonal variation in glacier mass (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mo>∆</mo> <mi mathvariant="normal">M</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) at different altitudes.</p>
Full article ">Figure 7
<p>Seasonal variation in glacier mass (<math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mo>∆</mo> <mi mathvariant="normal">M</mi> </mrow> <mo>¯</mo> </mover> </mrow> </semantics></math>) in (<b>a</b>) JFM, (<b>b</b>) AMJ, (<b>c</b>) JAS, and (<b>d</b>) OND.</p>
Full article ">Figure 8
<p>SHAP values of the meteorological drivers for the GMC.</p>
Full article ">Figure 9
<p>Scatter plots of the SHAP values of (<b>a</b>) temperature and (<b>b</b>) snowfall and their functional relationships with the GMC.</p>
Full article ">Figure 10
<p>(<b>a</b>) Monthly average temperature and accumulated precipitation within the study region (2018–2023); seasonal average (<b>b</b>) temperature and (<b>c</b>) precipitation. “Tem” denotes temperature, “TP” denotes total precipitation, and “SF” denotes snowfall.</p>
Full article ">Figure 11
<p>Seasonal averages of (<b>a</b>–<b>d</b>) temperature and (<b>e</b>–<b>h</b>) snowfall from 2018 to 2023.</p>
Full article ">Figure 12
<p>Variations in the standard deviations of the (<b>a</b>) temperature (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="normal">T</mi> <mi mathvariant="normal">e</mi> <mi mathvariant="normal">m</mi> </mrow> </msub> </mrow> </semantics></math>), snowfall (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="normal">S</mi> <mi mathvariant="normal">F</mi> </mrow> </msub> </mrow> </semantics></math>), and (<b>b</b>) seasonal variation in glaciers (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi mathvariant="sans-serif">σ</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> <mi mathvariant="normal">M</mi> </mrow> </msub> </mrow> </semantics></math>) with altitude.</p>
Full article ">
9 pages, 1639 KiB  
Article
Drinking Water Quality in Delta and Non-Delta Counties along the Mississippi River
by Emily V. Pickering, Chunrong Jia and Abu Mohd Naser
Water 2024, 16(18), 2622; https://doi.org/10.3390/w16182622 (registering DOI) - 15 Sep 2024
Abstract
The Mississippi Delta region has worse population health outcomes, including higher overall cardiovascular and infant mortality rates. Water quality has yet to be considered as a factor in these health disparities. The objective of this paper is to determine overall differences in basic [...] Read more.
The Mississippi Delta region has worse population health outcomes, including higher overall cardiovascular and infant mortality rates. Water quality has yet to be considered as a factor in these health disparities. The objective of this paper is to determine overall differences in basic water quality indicators, electrolytes of cardiovascular importance, trace elements, heavy metals, and radioactive ions of groundwater in delta and non-delta counties in states along the Mississippi River. Data were sourced from the major-ions dataset of the U.S. Geological Survey. We used the Wilcoxon rank sum test to determine the difference in water quality parameters. Overall, delta counties had lower total dissolved solids (TDS) (47 and 384 mg/L, p-value < 0.001), calcium (7 and 58 mg/L; p-value < 0.001), magnesium (2 and 22 mg/L; p-value < 0.001), and potassium (1.57 and 1.80 mg/L; p-value < 0.001) and higher sodium (38 mg/L and 22 mg/L; p-value < 0.001) compared to non-delta counties. Overall, there were no statistical differences in trace elements, heavy metals, and radioactive ions across delta versus non-delta counties. These results underscore the need for further epidemiological studies to understand if worse health outcomes in delta counties could be partially explained by these parameters. Full article
(This article belongs to the Section Water Quality and Contamination)
Show Figures

Figure 1

Figure 1
<p>Boxplots comparing basic water quality parameters in delta versus non-delta counties across eight delta states. Statistical significance is represented above each corresponding boxplot where one star (*) represents <span class="html-italic">p</span>-value &lt; 0.05, two stars (**) represent <span class="html-italic">p</span>-value &lt; 0.01, three stars (***) represent <span class="html-italic">p</span>-value &lt; 0.001, four stars (****) represents <span class="html-italic">p</span>-value &lt; 0.0001, and ns represents non-significance. The dots above the solid lines (whiskers) indicate outliers.</p>
Full article ">Figure 2
<p>Boxplots comparing major electrolytes concentrations in delta versus non-delta counties across eight delta states. Statistical significance is represented above each corresponding boxplot where one star (*) represents <span class="html-italic">p</span>-value &lt; 0.05, two stars (**) represent <span class="html-italic">p</span>-value &lt; 0.01, three stars (***) represent <span class="html-italic">p</span>-value &lt; 0.001, four stars (****) represent <span class="html-italic">p</span>-value &lt; 0.0001, and ns represents non-significance.</p>
Full article ">Figure 3
<p>Boxplots comparing trace element concentrations in delta versus non-delta counties across eight delta states. Statistical significance is represented above each corresponding boxplot where one star (*) represents <span class="html-italic">p</span>-value &lt; 0.05, two stars (**) represent <span class="html-italic">p</span>-value &lt; 0.01, three stars (***) represent <span class="html-italic">p</span>-value &lt; 0.001, and ns represents non-significance.</p>
Full article ">Figure 4
<p>Boxplots comparing heavy metals and radioactive ion concentrations in delta versus non-delta counties across eight delta states. Statistical significance is represented above each corresponding boxplot where one star (*) represents <span class="html-italic">p</span>-value &lt; 0.05, two stars (**) represents <span class="html-italic">p</span>-value &lt; 0.01, three stars (***) represents <span class="html-italic">p</span>-value &lt; 0.001, four stars (****) represents <span class="html-italic">p</span>-value &lt; 0.0001, and ns represents non-significance.</p>
Full article ">
20 pages, 1648 KiB  
Article
Exploring the Formation of Sustainable Entrepreneurial Intentions among Chinese University Students: A Dual Path Moderated Mediation Model
by Jinjin He, Zhongming Wang, Honghao Hu and Zengguang Fan
Sustainability 2024, 16(18), 8069; https://doi.org/10.3390/su16188069 (registering DOI) - 15 Sep 2024
Abstract
As Sustainable Development Goals (SDGs) gain traction in Chinese society, fostering sustainable entrepreneurship among university students has emerged as a key priority for universities and governments. Methods for increasing students’ sustainable entrepreneurship skills and knowledge for the creation of sustainable startups have attracted [...] Read more.
As Sustainable Development Goals (SDGs) gain traction in Chinese society, fostering sustainable entrepreneurship among university students has emerged as a key priority for universities and governments. Methods for increasing students’ sustainable entrepreneurship skills and knowledge for the creation of sustainable startups have attracted substantial attention. This study constructs a moderated mediation model based on entrepreneurial cognition theory to investigate the mediating roles of opportunity identification and attitude in the relationship between sustainable entrepreneurship education and sustainable entrepreneurial intention among university students, in addition to the moderating effect of empathy. The study surveyed 307 students from universities in the Yangtze River Delta region and employed hierarchical regression analysis to test the hypotheses. The results indicate that sustainable entrepreneurship education enhances students’ sustainable entrepreneurial intention by fostering their opportunity identification and attitude, and this enhancement effect is stronger when their level of empathy is higher. These findings enrich entrepreneurial cognition and empathy theories within the context of sustainable entrepreneurship and offer valuable insights for universities and policymakers in developing strategies to support sustainable entrepreneurship among university students. Full article
Show Figures

Figure 1

Figure 1
<p>Research framework.</p>
Full article ">Figure 2
<p>Summary of study results. Note(s): * <span class="html-italic">p</span> &lt; 0.05; ** <span class="html-italic">p</span> &lt; 0.01; *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Moderating effect of sustainable entrepreneurship education on opportunity identification with respect to empathy.</p>
Full article ">Figure 4
<p>Moderating effect of sustainable entrepreneurship education on attitude with respect to empathy.</p>
Full article ">
29 pages, 5760 KiB  
Review
Concentrations of Organochlorine, Organophosphorus, and Pyrethroid Pesticides in Rivers Worldwide (2014–2024): A Review
by Acela López-Benítez, Alfredo Guevara-Lara, Miguel A. Domínguez-Crespo, José A. Andraca-Adame and Aidé M. Torres-Huerta
Sustainability 2024, 16(18), 8066; https://doi.org/10.3390/su16188066 (registering DOI) - 15 Sep 2024
Viewed by 18
Abstract
The extensive use of pesticides has led to the contamination of natural resources, sometimes causing significant and irreversible damage to the environment and human health. Even though the use of many pesticides is banned, these compounds are still being found in rivers worldwide. [...] Read more.
The extensive use of pesticides has led to the contamination of natural resources, sometimes causing significant and irreversible damage to the environment and human health. Even though the use of many pesticides is banned, these compounds are still being found in rivers worldwide. In this review, 205 documents have been selected to provide an overview of pesticide contamination in rivers over the last 10 years (2014–2024). After these documents were examined, information of 47 river systems was organized according to the types of pesticides most frequently detected, including organochloride, organophosphorus, and pyrethroid compounds. A total of 156 compounds were classified, showing that 46% of these rivers contain organochlorine compounds, while 40% exhibit organophosphorus pesticides. Aldrin, hexachlorocyclohexane, and endosulfan were the predominant organochlorine pesticides with concentration values between 0.4 and 37 × 105 ng L−1. Chlorpyrifos, malathion, and diazinon were the main organophosphorus pesticides with concentrations between 1 and 11 × 105 ng L−1. Comparing the pesticide concentrations with standard guidelines, we found that the Ganga River in India (90 ng L−1), the Owan and Okura Rivers in Nigeria (210 and 9 × 103 ng L−1), and the Dong Nai River in Vietnam (68 ng L−1) exceed the permissible levels of aldrin (30 ng L−1). Full article
(This article belongs to the Section Hazards and Sustainability)
Show Figures

Figure 1

Figure 1
<p>Research methodology followed in this review.</p>
Full article ">Figure 2
<p>Classification of the documents mentioned in this review.</p>
Full article ">Figure 3
<p>Chemical structures and LD<sub>50</sub> values of some organochlorine pesticides.</p>
Full article ">Figure 4
<p>Concentrations of organochlorine pesticides detected in rivers around the world: (<b>a</b>) highest concentration values; (<b>b</b>) intermediate concentration values.</p>
Full article ">Figure 5
<p>Chemical structures and LD<sub>50</sub> values of some organophosphorus pesticides.</p>
Full article ">Figure 6
<p>Concentrations of organophosphorus pesticides detected in rivers around the world: (<b>a</b>) highest concentration values; (<b>b</b>) intermediate concentration values.</p>
Full article ">Figure 7
<p>Chemical structures and LD<sub>50</sub> values of some pyrethroid pesticides.</p>
Full article ">Figure 8
<p>Concentrations of pyrethroid pesticides detected in rivers around the world: (<b>a</b>) highest concentration values; (<b>b</b>) intermediate concentration values.</p>
Full article ">Figure 9
<p>Occurrence of pesticides in rivers worldwide (2014–2024).</p>
Full article ">Figure 10
<p>Geographical distribution of rivers across continents (2014–2024).</p>
Full article ">Figure 11
<p>Summary of the pesticides detected in the waters of 47 rivers from 2014 to 2024.</p>
Full article ">Figure 12
<p>Pesticide concentrations in rivers, considering the reported sampling year. (A: aldrin, AC: acephate, BHC: benzene hexachloride, CP: chlorpyrifos, CY: cypermethrin, D: deltamethrin, DVC: dichlorvos, DN: dieldrin, DT: dichlorodiphenyltrichloroethane, DZ: diazinon, E: endosulfan, EN: endrin, G: glyphosate, H: heptachlor, HCH: hexachlorocyclohexane, M: malathion, MT: metamidophos, P: parathion, PF: profenofos, Q: quinalphos, and T: triazophos).</p>
Full article ">
18 pages, 5123 KiB  
Article
Spatiotemporal Changes in the Quantity and Quality of Water in the Xiao Bei Mainstream of the Yellow River and Characteristics of Pollutant Fluxes
by Zhenzhen Yu, Xiaojuan Sun, Li Yan, Yong Li, Huijiao Jin and Shengde Yu
Water 2024, 16(18), 2616; https://doi.org/10.3390/w16182616 (registering DOI) - 15 Sep 2024
Viewed by 141
Abstract
The Xiao Bei mainstream, located in the middle reaches of the Yellow River, plays a vital role in regulating the quality of river water. Our study leveraged 73 years of hydrological data (1951–2023) to investigate long-term runoff trends and seasonal variations in the [...] Read more.
The Xiao Bei mainstream, located in the middle reaches of the Yellow River, plays a vital role in regulating the quality of river water. Our study leveraged 73 years of hydrological data (1951–2023) to investigate long-term runoff trends and seasonal variations in the Xiao Bei mainstream and its two key tributaries, the Wei and Fen Rivers. The results indicated a significant decline in runoff over time, with notable interannual fluctuations and an uneven distribution of runoff within the year. The Wei and Fen Rivers contributed 19.75% and 3.59% of the total runoff to the mainstream, respectively. Field monitoring was conducted at 11 locations along the investigated reach of Xiao Bei, assessing eight water quality parameters (temperature, pH, dissolved oxygen (DO), chemical oxygen demand (COD), ammonia nitrogen (NH3-N), total phosphorus (TP), permanganate index (CODMn), and 5-day biochemical oxygen demand (BOD5)). Our long-term results showed that the water quality of the Xiao Bei mainstream during the monitoring period was generally classified as Class III. Water quality parameters at the confluence points of the Wei and Fen Rivers with the Yellow River were higher compared with the mainstream. After these tributaries merged into the mainstream, local sections show increased concentrations, with the water quality parameters exhibiting spatial fluctuations. Considering the mass flux process of transmission of the quantity and quality of water, the annual NH3-N inputs from the Fen and Wei Rivers to the Yellow River accounted for 11.5% and 67.1%, respectively, and TP inputs accounted for 6.8% and 66.18%. These findings underscore the critical pollutant load from tributaries, highlighting the urgent need for effective pollution management strategies targeting these tributaries to improve the overall water quality of the Yellow River. This study sheds light on the spatiotemporal changes in runoff, water quality, and pollutant flux in the Xiao Bei mainstream and its tributaries, providing valuable insights to enhance the protection and management of the Yellow River’s water environment. Full article
Show Figures

Figure 1

Figure 1
<p>Yellow River (<b>left</b>) and map of the study area (<b>right</b>). The mainstream, tributaries, and basin area of the Yellow River are shown on the left. The investigated Xiao Bei mainstream, with the Wei River and Fen River tributaries and the proximal hydrologic and water quality stations are shown on the right.</p>
Full article ">Figure 2
<p>Trends and variations in monthly and annual runoff at Longmen and Tongguan hydrological stations from 1951 to 2023. (<b>a1</b>) Trend of monthly average runoff at Longmen hydrological station. (<b>a2</b>) Differences in runoff from the annual mean at Longmen hydrological station. (<b>b1</b>) Trend of monthly average runoff at Tongguan hydrological station. (<b>b2</b>) Differences in runoff from the annual mean at Tongguan hydrological station.</p>
Full article ">Figure 3
<p>Intra-annual distribution of long-term average runoff for the Xiao Bei mainstream. (<b>a</b>) Graphs of the monthly distribution of density with histograms and rug plots for the Longmen (blue) and Tongguan (orange) stations over 73 years. The smooth curves represent the probability density functions of the runoff data, providing a continuous view of the data’s distribution. The histograms illustrate the frequency of data points within specific intervals, while the rug plots show individual data points along the <span class="html-italic">x</span>-axis. (<b>b</b>) Line plot showing the measured intra-annual average runoff data for Longmen and Tongguan stations.</p>
Full article ">Figure 4
<p>Trends and variations in monthly and annual runoff at Huaxian and Hejin hydrological stations from 1951 to 2023. (<b>a1</b>) Trend of monthly average runoff at Huaxian hydrological station. (<b>a2</b>) Differences in runoff from the annual mean at Huaxian hydrological station. (<b>b1</b>) Trend of monthly average runoff at Hejin hydrological station. (<b>b2</b>) Differences in runoff from the annual mean at Hejin hydrological station.</p>
Full article ">Figure 5
<p>Intra-annual distribution of long-term average runoff from the Wei River (Huaxian hydrological station) and Fen River (Hejin hydrological station). (<b>a</b>) Graphs of the monthly distribution of density with histograms and rug plots for Huaxian (green) and Hejin (purple) stations over 73 years. The smooth curves represent the probability density functions of the runoff data, providing a continuous view of the data’s distribution. The histograms illustrate the frequency of data points within specific intervals, while the rug plots show individual data points along the <span class="html-italic">x</span>-axis. (<b>b</b>) Line plot showing the measured intra-annual average runoff data for Huaxian and Hejin stations.</p>
Full article ">Figure 6
<p>Comparison of average monthly runoff between the tributaries and mainstream of the Yellow River. (<b>a</b>) Contribution of the Fen River to the Yellow River (<b>b</b>) Contribution of the Wei River to the Yellow River.</p>
Full article ">Figure 7
<p>Characterization of the concentrations of factors for monitoring water quality in the Xiao Bei mainstream: Green area display the data distribution and dark area represents the inter quartile range, which spans from the 25th to the 75th percentile of the data.</p>
Full article ">Figure 8
<p>Comparison of water quality factors in the mainstream and tributaries of the Xiao Bei mainstream.</p>
Full article ">Figure 9
<p>Changes in the monitored values of water quality factors over 11 sampling points along the studied reach. (<b>a</b>) Temperature, (<b>b</b>) pH, (<b>c</b>) chemical oxygen demand (COD), (<b>d</b>) ammonia nitrogen (NH<sub>3</sub>-N), (<b>e</b>) total phosphorus (TP), (<b>f</b>) permanganate index (COD<sub>Mn</sub>), (<b>g</b>) 5-day biochemical oxygen demand (BOD<sub>5</sub>).</p>
Full article ">Figure 10
<p>Monthly changes in TP and NH<sub>3</sub>-N fluxes in the Xiao Bei mainstream in 2021.</p>
Full article ">
17 pages, 951 KiB  
Article
Multiphase Partitioning of Estrogens in a River Impacted by Feedlot Wastewater Discharge
by Kuo-Hui Yang, Hao-Shen Hung, Wei-Hsiang Huang, Chi-Ying Hsieh and Ting-Chien Chen
Toxics 2024, 12(9), 671; https://doi.org/10.3390/toxics12090671 (registering DOI) - 14 Sep 2024
Viewed by 204
Abstract
Estrogens in river systems can significantly impact aquatic ecosystems. This study aimed to investigate the multiphase partitioning of estrogens in Wulo Creek, Taiwan, which receives animal feedlot wastewater, to understand their distribution and potential environmental implications. Water samples were separated into suspended particulate [...] Read more.
Estrogens in river systems can significantly impact aquatic ecosystems. This study aimed to investigate the multiphase partitioning of estrogens in Wulo Creek, Taiwan, which receives animal feedlot wastewater, to understand their distribution and potential environmental implications. Water samples were separated into suspended particulate matter (SPM), colloidal, and soluble phases using centrifugation and cross-flow ultrafiltration. Concentrations of estrone (E1), 17β-estradiol (E2), and estriol (E3) in each phase were analyzed using LC/MS/MS. Partition coefficients were calculated to assess estrogen distribution among phases. Estrogens were predominantly found in the soluble phase (85.8–87.3%). The risk assessment of estrogen equivalent (EEQ) values suggests that estrogen concentration in water poses a higher risk compared to SPM, with a majority of the samples indicating a high risk to aquatic organisms. The colloidal phase contained 12.7–14.2% of estrogens. The log KCOC values (4.72–4.77 L/kg-C) were significantly higher than the log KOC and log KPOC values (2.02–3.40 L/kg-C) for all estrogens. Colloids play a critical role in estrogen distribution in river systems, potentially influencing their fate, transport, and biotoxicity. This finding highlights the importance of considering colloidal interactions in assessing estrogen behavior in aquatic environments. Full article
(This article belongs to the Special Issue Environmental Transport and Transformation of Pollutants)
Show Figures

Figure 1

Figure 1
<p>Geographical map of sampling site locations (adapted from Hung et al. [<a href="#B37-toxics-12-00671" class="html-bibr">37</a>], and Ministry of Agriculture, Republic of China (Taiwan)).</p>
Full article ">Figure 2
<p>The mass percentages of organic carbon (OC) distributed on colloidal and soluble phases.</p>
Full article ">Figure 3
<p>The mass percentages of estrogens distributed on colloidal and soluble phases: (<b>a</b>) (E1), (<b>b</b>) (E2), and (<b>c</b>) (E3).</p>
Full article ">
22 pages, 19530 KiB  
Article
Cascading Landslide: Kinematic and Finite Element Method Analysis through Remote Sensing Techniques
by Claudia Zito, Massimo Mangifesta, Mirko Francioni, Luigi Guerriero, Diego Di Martire, Domenico Calcaterra and Nicola Sciarra
Remote Sens. 2024, 16(18), 3423; https://doi.org/10.3390/rs16183423 (registering DOI) - 14 Sep 2024
Viewed by 217
Abstract
Cascading landslides are specific multi-hazard events in which a primary movement triggers successive landslide processes. Areas with dynamic and quickly changing environments are more prone to this type of phenomena. Both the kind and the evolution velocity of a landslide depends on the [...] Read more.
Cascading landslides are specific multi-hazard events in which a primary movement triggers successive landslide processes. Areas with dynamic and quickly changing environments are more prone to this type of phenomena. Both the kind and the evolution velocity of a landslide depends on the materials involved. Indeed, rockfalls are generated when rocks fall from a very steep slope, while debris flow and/or mudslides are generated by fine materials like silt and clay after strong water imbibition. These events can amplify the damage caused by the initial trigger and propagate instability along a slope, often resulting in significant environmental and societal impacts. The Morino-Rendinara cascading landslide, situated in the Ernici Mountains along the border of the Abruzzo and Lazio regions (Italy), serves as a notable example of the complexities and devastating consequences associated with such events. In March 2021, a substantial debris flow event obstructed the Liri River, marking the latest step in a series of landslide events. Conventional techniques such as geomorphological observations and geological surveys may not provide exhaustive information to explain the landslide phenomena in progress. For this reason, UAV image acquisition, InSAR interferometry, and pixel offset analysis can be used to improve the knowledge of the mechanism and kinematics of landslide events. In this work, the interferometric data ranged from 3 January 2020 to 24 March 2023, while the pixel offset data covered the period from 2016 to 2022. The choice of such an extensive data window provided comprehensive insight into the investigated events, including the possibility of identifying other unrecorded events and aiding in the development of more effective mitigation strategies. Furthermore, to supplement the analysis, a specific finite element method for slope stability analysis was used to reconstruct the deep geometry of the system, emphasizing the effect of groundwater-level flow on slope stability. All of the findings indicate that major landslide activities were concentrated during the heavy rainfall season, with movements ranging from several centimeters per year. These results were consistent with numerical analyses, which showed that the potential slip surface became significantly more unstable when the water table was elevated. Full article
Show Figures

Figure 1

Figure 1
<p>Aerial and field images of the Morino-Rendinara landslide that are representative of the impact of the landslide on the environment. (<b>a</b>) Overview of phenomenon taken from Google Earth [<a href="#B16-remotesensing-16-03423" class="html-bibr">16</a>] satellite images of 13 June 2022, from the upper sector near Morino Hamlet to the lower sector, Liri River, and deep-seated rotational slide; (<b>b</b>) Details of rockfall/avalanches sector; (<b>c</b>) Debris flow source area; (<b>d</b>) Debris flow transit zone; (<b>e</b>) Lowest debris flow transit zone; (<b>f</b>) Liri River dam; and (<b>g</b>) Effect on Liri River dam.</p>
Full article ">Figure 2
<p>Geographical location of Morino-Rendinara. Green lines indicate the regional boundaries; red lines indicate the municipality of Morino, Castronovo, and San Vincenzo Valle Roveto composing the involved municipality; the light blue square indicates the landslide and the study area.</p>
Full article ">Figure 3
<p>Geological map extract from the CARG (Geological CARtography Map n.220 Sora) Project [<a href="#B22-remotesensing-16-03423" class="html-bibr">22</a>], with indications of the geological formations and tectonic processes present in the area.</p>
Full article ">Figure 4
<p>Maps of the survey and debritic cover layer reconstruction using cross-sections to empathize the heterogeneity of deposits covering the substrate. (a) The section develops on maximum slope line. (b) The section develops on perpendicular direction.</p>
Full article ">Figure 5
<p>Conceptual flow chart of the work phases.</p>
Full article ">Figure 6
<p>Spatiotemporal baseline map of SBAS-InSAR interferometric data of ascending track (<b>a</b>) and descending track (<b>b</b>).</p>
Full article ">Figure 7
<p>The inventory map was drawn using the results of the study. The image identifies three main mechanisms: a rockfall in the upper part, a deep-seated rotational slide in the central part, and a debris flow in the lower part.</p>
Full article ">Figure 8
<p>Details of the landslide inventory map. The various fillings show the different landslide types identified in the study area: the rockfall in the upper part, the deep-seated rotational slide in the central part, and the debris flow in the lower part of the slope.</p>
Full article ">Figure 9
<p>PS velocity along the ascending (<b>a</b>) and descending (<b>b</b>) geometries from 2020 to 2023. Red dots indicate major velocity trends and instability, green and blue dots indicate minor velocity and stable sectors.</p>
Full article ">Figure 10
<p>Selected time series in ascending geometry (<b>a</b>) and descending geometry (<b>b</b>). The analyzed time series illustrates a very unstable sector represented by reflectors P106_60-61 and 102_57 in ascending geometry and P_70_141-141-136 in descending geometry. Additionally, some stable sectors are represented, such as P_83_135-143, P_85_68, and P_86_69.</p>
Full article ">Figure 11
<p>Time series vs. rainfall analyses. (<b>a</b>) Monthly cumulative rainfall for analysis period vs. one ascending and descending representative time series; (<b>b</b>) Daily cumulative rainfall for analysis period vs. one ascending and descending representative time series; (<b>c</b>) Cumulative rainfall for analysis period vs. one ascending and descending representative time series.</p>
Full article ">Figure 12
<p>Shear strain index results of the 2D numerical modelling with different pore pressure conditions. (<b>a</b>) Analysis without pore pressure; (<b>b</b>) Analysis with water table 0.5 m from ground level; (<b>c</b>) Analysis with water table 2.0 m from ground level; (<b>d</b>) Analysis with water table 6.0 m from ground level.</p>
Full article ">
22 pages, 3621 KiB  
Article
Estimating Non-Stationary Extreme-Value Probability Distribution Shifts and Their Parameters Under Climate Change Using L-Moments and L-Moment Ratio Diagrams: A Case Study of Hydrologic Drought in the Goat River Near Creston, British Columbia
by Isaac Dekker, Kristian L. Dubrawski, Pearce Jones and Ryan MacDonald
Hydrology 2024, 11(9), 154; https://doi.org/10.3390/hydrology11090154 (registering DOI) - 14 Sep 2024
Viewed by 168
Abstract
Here, we investigate the use of rolling-windowed L-moments (RWLMs) and L-moment ratio diagrams (LMRDs) combined with a Multiple Linear Regression (MLR) machine learning algorithm to model non-stationary low-flow hydrological extremes with the potential to simultaneously understand time-variant shape, scale, location, and probability distribution [...] Read more.
Here, we investigate the use of rolling-windowed L-moments (RWLMs) and L-moment ratio diagrams (LMRDs) combined with a Multiple Linear Regression (MLR) machine learning algorithm to model non-stationary low-flow hydrological extremes with the potential to simultaneously understand time-variant shape, scale, location, and probability distribution (PD) shifts under climate change. By employing LMRDs, we analyse changes in PDs and their parameters over time, identifying key environmental predictors such as lagged precipitation for September 5-day low-flows. Our findings indicate a significant relationship between total August precipitation L-moment ratios (LMRs) and September 5-day low-flow LMRs (τ2-Precipitation and τ2-Discharge: R2 = 0.675, p-values < 0.001; τ3-Precipitation and τ3-Discharge: R2 = 0.925, p-value for slope < 0.001, intercept not significant with p = 0.451, assuming α = 0.05 and a 31-year RWLM), which we later refine and use for prediction within our MLR algorithm. The methodology, applied to the Goat River near Creston, British Columbia, aids in understanding the implications of climate change on water resources, particularly for the yaqan nuʔkiy First Nation. We find that future low-flows under climate change will be outside the Natural Range of Variability (NROV) simulated from historical records (assuming a constant PD). This study provides insights that may help in adaptive water management strategies necessary to help preserve Indigenous cultural rights and practices and to help sustain fish and fish habitat into the future. Full article
(This article belongs to the Section Hydrological and Hydrodynamic Processes and Modelling)
Show Figures

Figure 1

Figure 1
<p>L-moment ratio diagrams (LMRDs) for: (<b>a</b>) August total precipitation (mm) and (<b>b</b>) September 5-day low-flow (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>). Each panel includes the L-Coefficient of Variation (L-CV;<math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>) versus L-skewness (<math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>) and L-kurtosis versus L-skewness ratios.</p>
Full article ">Figure 2
<p>Relationship between L-moment ratios (LMRs) of August total precipitation (mm) and September 5-day low-flow (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>. <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>-Precipitation and <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>-Discharge: <span class="html-italic">p</span>-values &lt; 0.001; <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>-Precipitation and <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>-Discharge: <span class="html-italic">p</span>-value for slope &lt; 0.001, intercept not significant with <span class="html-italic">p</span> = 0.451, assuming <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.05 and a 31-year rolling-windowed L-moments (RDLMs).</p>
Full article ">Figure 3
<p>Comparison of predicted and observed L-moments (LMs; testing during training) using a 31-year rolling-window. The plots display the predicted (green) and observed (blue) values for the first (<b>a</b>), second (<b>b</b>), third (<b>c</b>), and fourth (<b>d</b>) LMs. Each subplot includes the Overall Mean Squared Error (MSE) between the predicted and observed values computed by summing and averaging the best-fit model Squared Error (SE) for each step in the forward chaining process. The equations plotted alongside the model are derived from the final (best-fit) iteration (index 38), which demonstrated the lowest SE.</p>
Full article ">Figure 4
<p>Location, scale, and shape parameters estimated using [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>]’s method of L-moments (LMs) for the Generalized Extreme Value (GEV) probability distribution (PD) for the observed (blue) and predicted (testing during training; dashed red) LMs under a 31-year rolling-window.</p>
Full article ">Figure 5
<p>LMRDs using 31-year windows under two Representative Concentration Pathways (RCP) scenarios. Panels (<b>a</b>–<b>c</b>) correspond to the RCP 4.5 scenario, while panels (<b>d</b>–<b>f</b>) correspond to the RCP 8.5 scenario. Diagrams show: (<b>a</b>,<b>d</b>) L-CV/L-skewness, (<b>b</b>,<b>e</b>) L-kurtosis/L-skewness, with theoretical PDs described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] (distributions include Generalized Logistic (GLO), Generalized Extreme Value (GEV), Generalized Pareto (GPA), Generalized Normal (GNO), Pearson Type III (PE3), Wakeby Lower Bound (WAK_LB), and All Distribution Lower Bound (ALL_LB)). Plots (<b>c</b>,<b>f</b>) show the distribution count for each window. The observed LMRs for the 5-day September low-flow at the Water Survey of Canada (WSC) Goat River Near Erikson Hydrometric Gauge Station (<a href="https://wateroffice.ec.gc.ca/report/data_availability_e.html?type=historical&amp;station=08NH004&amp;parameter_type=Flow&amp;wbdisable=true" target="_blank">08NH004</a>) are plotted alongside simulated future data derived from Multiple Linear Regression (MLR) driven with total August precipitation LMs. Future data are generated using a splice of six Coupled Model Intercomparison Project Phase 5 (CMIP5) series downscaled climate models (median of “ACCESS1-0”, “CanESM2”, “CCSM4”, “CNRM-CM5”, “HadGEM2-ES”, and “MPI-ESM-LR” from 2018 to 2100) downloaded using the single cell extraction tool from the Pacific Climate Impacts Consortium (<a href="https://pacificclimate.org/data/gridded-hydrologic-model-output" target="_blank">PCIC</a>). Historical climate data are downloaded from Historical Climate Data Online (HCDO) repository for the Creston station (Climate ID <a href="https://climate.weather.gc.ca/climate_data/daily_data_e.html?hlyRange=%7C&amp;dlyRange=1912-06-01%7C2017-12-31&amp;mlyRange=1912-01-01%7C2007-02-01&amp;StationID=1111&amp;Prov=BC&amp;urlExtension=_e.html&amp;searchType=stnName&amp;optLimit=yearRange&amp;StartYear=1840&amp;EndYear=2024&amp;selRowPerPage=25&amp;Line=0&amp;searchMethod=contains&amp;Month=12&amp;Day=2&amp;txtStationName=Creston&amp;timeframe=2&amp;Year=2017" target="_blank">1142160</a>; available from 1996 to 2017).</p>
Full article ">Figure 6
<p>Results of the L-moments (LMs) derived from Multiple Linear Regression (MLR) fit to a Generalized Extreme Value (GEV) probability distribution (PD) to produce shape, scale, and location parameters: (<b>a</b>) GEV parameters (shape, scale, and location) over 144 rolling windowed time units under Representative Concentration Pathway (RCP) 4.5 and (<b>b</b>) RCP 8.5.</p>
Full article ">Figure 7
<p>LMs derived from MLR fit to a GEV PD to produce shape, scale, and location parameters to derive median flows (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) and confidence intervals (CIs) estimated from percentiles: (<b>a</b>) simulated September 5-day low-flows driven by MLR regression using August total precipitation LMs over a 31-year rolling time window under (<b>a</b>) RCP 4.5 and (<b>b</b>) RCP 8.5. The dashed red line denotes a flow of &lt;1 m<sup>3</sup>/s. Note: each simulation is based on <span class="html-italic">n</span> = 1000 iterations for both panels.</p>
Full article ">Figure 8
<p>Results of the LMs derived from MLR fit to the best-fit probability distribution (PD) (distributions described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>]) to produce shape, scale, and location parameters to derive median flows (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) and confidence intervals (CIs) estimated numerically from percentiles: (<b>a</b>) simulated September 5-day low-flows driven by MLR regression using August total precipitation LMs over a 31-year moving window under RCP 4.5 and (<b>b</b>) RCP 8.5. The dashed red line denotes a flow of &lt;1 m<sup>3</sup>/s. Note: each simulation is based on <span class="html-italic">n</span> = 1000 iterations for both panels.</p>
Full article ">Figure 9
<p>Overall standardized Mean Square Error (MSE) across different window sizes during model training.</p>
Full article ">Figure 10
<p>Sensitivity of window size on location, scale, and shape parameters for September 5-day low-flow estimated using the method of LMs described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] derived from MLR driven by total August precipitation LMs for six common distributions described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] (Generalized Extreme Value (GEV; (<b>a</b>–<b>c</b>)), Generalized Logistic (GLO; (<b>d</b>–<b>f</b>)), Generalized Normal (GNO; (<b>g</b>–<b>i</b>)), Pearson Type III (PE3; (<b>j</b>–<b>l</b>)), and Generalized Pareto (GPA; (<b>m</b>–<b>o</b>)). The solid line displays data under the Representative Concentration Pathway (RCP) 4.5 emission scenario, while the dashed line displays the RCP 8.5 emissions scenario.</p>
Full article ">Figure 11
<p>Low-flow exceedance and cumulative exceedance probability for the Goat River near Erikson Gauge Station, showing values less than 2.7 cubic meters per second (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) assuming <span class="html-italic">n</span> = 1000 simulations and a Generalized Extreme Value (GEV) probability distribution (PD). Future data assume a Representative Concentration Pathway (RCP) 4.5 emissions scenario.</p>
Full article ">
20 pages, 2596 KiB  
Article
Integrating Active and Passive Remote Sensing Data for Forest Age Estimation in Shangri-La City, China
by Feng Cheng, Ruijiao Yang and Junen Wu
Forests 2024, 15(9), 1622; https://doi.org/10.3390/f15091622 (registering DOI) - 14 Sep 2024
Viewed by 186
Abstract
The accurate mapping of age structure and access to spatially explicit information are essential to optimal planning and policy-making for forest ecosystems, including forest management and sustainable economic development. Specifically, surveying and mapping the age structure of forests is crucial for calculating the [...] Read more.
The accurate mapping of age structure and access to spatially explicit information are essential to optimal planning and policy-making for forest ecosystems, including forest management and sustainable economic development. Specifically, surveying and mapping the age structure of forests is crucial for calculating the carbon sequestration capacity of forest ecosystems. However, spatial heterogeneity and limited accessibility make forest age mapping in mountainous areas challenging. Here, we present a new workflow using ICESat-2 LiDAR data integrated with multisource remote sensing imagery to estimate forest age in Shangri-La, China. Two methods—a climate-driven exponential model and a random forest algorithm—are compared to infer the age structure of the five dominant species in Shangri-La. The climate-driven model, with an R2 of 0.67 and an RMSE of 12.79 years, outperforms the random forest model. The derived wall-to-wall forest age map at 30 m resolution reveals that nearly all forests in Shangri-La are mature or overmature, especially among the high-elevation species Abies fabri (Mast.) Craib and Picea asperata Mast., compared with Pinus yunnanensis Franch., Quercus aquifolioides Rehd. and E.H. Wils. and Pinus densata Mast., where the age structure is more evenly distributed across different elevation ranges. Younger forests are frequently found around human settlements and along the Jinsha River valley, whereas older forests are located in remote and high-elevation areas that are less disturbed. The combined use of active and passive remote sensing data has resulted in substantial improvements in the spatial detail and accuracy of wall-to-wall age mapping, which is expected to be a cost-effective approach for supporting forest management and carbon accounting in this important ecological region. The method developed here can be scaled to other mountain areas both to understand the age patterns and structure of mountain forests and to provide critical information for forestation, reforestation and carbon accounting in surface-to-high mountain areas, which are increasingly crucial for climate mitigation. Full article
(This article belongs to the Section Forest Inventory, Modeling and Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>Overview of the study area in Shangri-La city, Yunnan Province, China: (<b>a</b>) location of Shangri-La city within Yunnan Province; (<b>b</b>) digital elevation model (DEM) showing the complex mountainous terrain of the study area; (<b>c</b>) spatial distributions of the five dominant tree species in Shangri-La city; (<b>d</b>) distribution across different canopy height intervals.</p>
Full article ">Figure 2
<p>Workflow of forest age estimation in Shangri-La based on the integration of ICESat-2/ATLAS data, multisource remote sensing imagery, and field survey samples. RF indicates random forest.</p>
Full article ">Figure 3
<p>Comparison of forest age predictions between the random forest (RF) model and the climate-dependent exponential (Exp) model for all species combined and individual dominant tree species in Shangri-La. (<b>a</b>) All species combined; (<b>b</b>) <span class="html-italic">P. densata</span>; (<b>c</b>) <span class="html-italic">P. asperata</span>; (<b>d</b>) <span class="html-italic">A. fabri</span>; (<b>e</b>) <span class="html-italic">Q. aquifolioides</span>; (<b>f</b>) <span class="html-italic">P. yunnanensis</span> The dashed line represents the 1:1 line, while blue and red points and lines represent predictions from the RF and Exp models, respectively. <span class="html-italic">R</span><sup>2</sup> and RMSE values are provided for each model and species.</p>
Full article ">Figure 4
<p>Spatial patterns and area statistics of forest age groups in Shangri-La city derived from the climate-dependent exponential model: (<b>a</b>) The overall forest age distribution classified into five groups: young, middle-aged, near-mature, mature, and overmature forests. The inset charts show the area proportions of different age groups for each dominant species; (<b>b</b>–<b>f</b>) forest age distributions of the five dominant tree species: (<b>b</b>) <span class="html-italic">P. densata</span>, (<b>c</b>) <span class="html-italic">P. asperata</span>, (<b>d</b>) <span class="html-italic">Q. aquifolioides</span>, (<b>e</b>) <span class="html-italic">A. fabri</span>, (<b>f</b>) <span class="html-italic">P. yunnanensis</span>. The age thresholds and area statistics for each species are provided in <a href="#forests-15-01622-t002" class="html-table">Table 2</a>.</p>
Full article ">
18 pages, 7514 KiB  
Article
Enhancing Evapotranspiration Estimations through Multi-Source Product Fusion in the Yellow River Basin, China
by Runke Wang, Xiaoni You, Yaya Shi and Chengyong Wu
Water 2024, 16(18), 2603; https://doi.org/10.3390/w16182603 (registering DOI) - 14 Sep 2024
Viewed by 200
Abstract
An accurate estimation of evapotranspiration (ET) is critical to understanding the water cycle in watersheds and promoting the sustainable utilization of water resources. Although there are various ET products in the Yellow River Basin, various ET products have many uncertainties due to input [...] Read more.
An accurate estimation of evapotranspiration (ET) is critical to understanding the water cycle in watersheds and promoting the sustainable utilization of water resources. Although there are various ET products in the Yellow River Basin, various ET products have many uncertainties due to input data, parameterization schemes, and scale conversion, resulting in significant uncertainties in regional ET data products. To reduce the uncertainty of a single product and obtain more accurate ET data, more accurate ET data can be obtained by fusing different ET data. Addressing this challenge, by calculating the uncertainty of three ET data products, namely global land surface satellite (GLASS) ET, Penman–Monteith–Leuning (PML)-V2 ET, and reliability-affordable averaging (REA) ET, the weight of each product is obtained to drive the Bayesian three-cornered Hat (BTCH) algorithm to obtain higher quality fused ET data, which are then validated at the site and basin scales, and the accuracy has significantly improved compared to a single input product. On a daily scale, the fused data’s root mean square error (RMSE) is 0.78 mm/day and 1.14 mm/day. The mean absolute error (MAE) is 0.53 mm/day and 0.84 mm/day, respectively, which has a lower RMSE and MAE than the model input data; the correlation coefficients (R) are 0.9 and 0.83, respectively. At the basin scale, the RMSE and MAE of the annual average ET of the fused data are 11.77 mm/year and 14.95 mm/year, respectively, and the correlation coefficient is 0.84. The results show that the BTCH ET fusion data are better than single-input product data. An analysis of the fused ET data on a spatiotemporal scale shows that from 2001 to 2017, the ET increased in 85.64% of the area of the Yellow River Basin. Fluctuations in ET were greater in the middle reaches of the Yellow River than in the upstream and downstream regions. The BTCH algorithm has indispensable reference value for regional ET estimation research, and the ET data after BTCH algorithm fusion have higher data quality than the original input data. The fused ET data can inform the development of management strategies for water resources in the YRB and provide a deeper understanding of the regional water supply and demand balance mechanism. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the research area and distribution of major cities, hydrological stations, and flux observation stations.</p>
Full article ">Figure 2
<p>Technical flowchart.</p>
Full article ">Figure 3
<p>Uncertainty results of different evapotranspiration products, (<b>a</b>) GLASS ET, (<b>b</b>) PML-V2 ET, (<b>c</b>) REA ET, and (<b>d</b>) uncertainty statistics.</p>
Full article ">Figure 4
<p>Accuracy validation of ET products, fusion ET data, and flux tower data at Haibei shrub station from 2003 to 2010. (<b>a</b>) GLASS, (<b>b</b>) REA, (<b>c</b>) PML-V2, (<b>d</b>) BTCH (red: fitting line; dashed: best-fit line, a 1:1 line).</p>
Full article ">Figure 5
<p>Accuracy validation of ET products, fusion ET data, and flux tower data at Haibei wetland station in 2005. (<b>a</b>) GLASS, (<b>b</b>) REA, (<b>c</b>) PML-V2, (<b>d</b>) BTCH (red: fitting line; dashed: best-fit line, a 1:1 line).</p>
Full article ">Figure 6
<p>Energy balance ratio of observed data and fused ET data from the Haibei shrub station. (<b>a</b>) measured data, (<b>b</b>) fused ET data.</p>
Full article ">Figure 7
<p>Comparison of the average and observed values for ET from the different products in the YRB. (<b>a</b>) GLASS, (<b>b</b>) REA, (<b>c</b>) PML-V2, (<b>d</b>) BTCH.</p>
Full article ">Figure 8
<p>Variation characteristics for the average values of different ET products in the YRB.</p>
Full article ">Figure 9
<p>Interannual ET change rate (<b>a</b>) and M-K significance test (<b>b</b>) of BTCH fusion from 2001 to 2017.</p>
Full article ">
Back to TopTop