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

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

Search Results (2,406)

Search Parameters:
Keywords = meteorological variables

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
19 pages, 7060 KiB  
Article
A Comparison between Radar Variables and Hail Pads for a Twenty-Year Period
by Tomeu Rigo and Carme Farnell
Climate 2024, 12(10), 158; https://doi.org/10.3390/cli12100158 - 4 Oct 2024
Abstract
The time and spatial variability of hail events limit the capability of diagnosing the occurrence and stones’ size in thunderstorms using weather radars. The bibliography presents multiple variables and methods with different pros and cons. The studied area, the Lleida Plain, is annually [...] Read more.
The time and spatial variability of hail events limit the capability of diagnosing the occurrence and stones’ size in thunderstorms using weather radars. The bibliography presents multiple variables and methods with different pros and cons. The studied area, the Lleida Plain, is annually hit by different hailstorms, which have a high impact on the agricultural sector. A rectangular distributed hail pad network in this plain has worked operationally since 2000 to provide information regarding different aspects of hail impact. Since 2002, the Servei Meteorològic de Catalunya (SMC) has operated a single-pol C-band weather radar network that volumetrically covers the region of interest. During these years, the SMC staff has been working on improving the capability of detecting hail, adapting some parameters and searching for thresholds that help to identify the occurrence and size of the stones in thunderstorms. The current research analyzes a twenty-year period (2004–2023) to provide a good picture of the hailstorms occurring in the region of interest. The main research result is that VIL (Vertically Integrated Liquid) density is a better indicator for hailstone size than VIL, which presents more uncertainty in discriminating different hail categories. Full article
(This article belongs to the Special Issue Applications of Smart Technologies in Climate Risk and Adaptation)
Show Figures

Figure 1

Figure 1
<p><b>Top left</b>: Map of Western Europe. The area included in the rectangle is the region of study. <b>Bottom right</b>: Zoom in on the region of interest. The dots indicate the location of the radars and the circles indicate the 50 (dots) and 100 (straight) km range for each radar. The red shaded area marks the region covered by the hail pad network. “LMI”, “CDV”, “PBE”, and “PDA” indicate the locations of the radars of La Miranda, Creu del Vent, Puig Bernat, and Puig d’Arques, respectively.</p>
Full article ">Figure 2
<p>Example of hit hail pad corresponding to the event of 29 August 2023. The image has been filtered to highlight the impacts over the plaque. The striped rectangle in the middle of the pad corresponds to the calibration area (see [<a href="#B19-climate-12-00158" class="html-bibr">19</a>] for more information).</p>
Full article ">Figure 3
<p>(<b>A</b>) CAPPI at 3 km height at 17.12 UTC on 28 July 2028. The black arrow line shows the cross section segment shown in panel (<b>B</b>). (<b>B</b>) Cross section of the thunderstorm over the region of interest at the same time as panel (<b>A</b>). (<b>C</b>) Maximum VIL field for the whole day of 28 July 2028. The dots indicate the maximum hail size registered by the different hail pads. (<b>D</b>) Same as panel (<b>C</b>), but for the maximum VIL density field.</p>
Full article ">Figure 4
<p><b>Top</b>: each dot corresponds to an analyzed hail pad during the event of 5 July 2012. <b>Below</b>: normalized coordinates of the same points centered in (0, 0).</p>
Full article ">Figure 5
<p><b>Top</b>: Hail size distribution (in logarithm) for the whole dataset of hail pad registers. <b>Bottom</b>: Linear fitting of the distribution.</p>
Full article ">Figure 6
<p>From top to bottom: distribution of graupel (grey), hail (blue), and severe hail (red) for the week of the year (<b>A</b>), the month (<b>B</b>), the year (<b>C</b>), and the maximum daily surface temperature (<b>D</b>).</p>
Full article ">Figure 7
<p>As in <a href="#climate-12-00158-f006" class="html-fig">Figure 6</a>, but for the percentage distribution (only for cases with plaques with impacts). (<b>A</b>–<b>C</b>) panels correspond to weekly of the year, monthly, and yearly distributions, respectively.</p>
Full article ">Figure 8
<p>Box plots of the VIL for the four hail categories detected in the hail pads (from left to right: no hail, graupel, hail and severe hail).</p>
Full article ">Figure 9
<p>Same as <a href="#climate-12-00158-f008" class="html-fig">Figure 8</a>, but for VIL density.</p>
Full article ">Figure 10
<p>VIL (<b>top</b>) and VIL density (<b>bottom</b>) distributions for all the four categories (black: no hail; grey: graupel; blue: hail; and red: severe hail).</p>
Full article ">Figure 11
<p>Total sample of hail records (dots) for 2000–2023 (dark grey for null, grey for graupel, cyan for hail, and orange for severe hail). The dashed lines correspond to the 10th percentile of occurrence for each category (black for null, green for graupel, blue for hail, and red for severe hail), showing the usual behavior of the hailfall in the region concerning the center of the event.</p>
Full article ">Figure 12
<p>Graupel (<b>A</b>), hail (<b>B</b>), and severe hail (<b>C</b>) spatial distributions estimated using maximum daily VIL fields for 2013–2023. Dotted, dashed, and straight lines indicate the 10th, 50th, and 90th ground observations percentiles.</p>
Full article ">Figure 13
<p>Same as <a href="#climate-12-00158-f012" class="html-fig">Figure 12</a>, but for VIL density (Graupel, hail, and severe hail in panels (<b>A</b>–<b>C</b>), respectively).</p>
Full article ">
20 pages, 4810 KiB  
Article
Understanding Spatio-Temporal Hydrological Dynamics Using SWAT: A Case Study in the Pativilca Basin
by Yenica Pachac-Huerta, Waldo Lavado-Casimiro, Melania Zapana and Robinson Peña
Hydrology 2024, 11(10), 165; https://doi.org/10.3390/hydrology11100165 - 4 Oct 2024
Abstract
This study investigates the hydrological dynamics of the Pativilca Basin in the Southern Hemisphere using the SWAT (Soil and Water Assessment Tool) model. Seventy-seven watersheds across a mountainous region were analyzed using elevation data, land cover, soil type, and gridded meteorological products (RAIN4PE [...] Read more.
This study investigates the hydrological dynamics of the Pativilca Basin in the Southern Hemisphere using the SWAT (Soil and Water Assessment Tool) model. Seventy-seven watersheds across a mountainous region were analyzed using elevation data, land cover, soil type, and gridded meteorological products (RAIN4PE and PISCO) for hydrological simulations. Watershed delineation, aided by a Digital Elevation Model, enabled the identification of critical drainage points and the definition of Hydrological Response Units (HRUs). The model calibration and validation, performed using the SWAT-CUP with the SUFI-2 algorithm, achieved Nash–Sutcliffe Efficiency (NSE) values of 0.69 and 0.72, respectively. Cluster analysis categorized the watersheds into six distinct groups with unique hydrological and climatic characteristics. The results showed significant spatial variability in the precipitation and temperature, with pronounced seasonality influencing the daily flow patterns. The higher-altitude watersheds exhibited greater soil water storage and more effective aquifer recharge, whereas the lower-altitude watersheds, despite receiving less precipitation, displayed higher flows due to runoff from the upstream areas. These findings emphasize the importance of incorporating seasonality and spatial variability into water resource planning in mountainous regions and demonstrate the SWAT model’s effectiveness in predicting hydrological responses in the Pativilca Basin, laying the groundwork for future research in mountain hydrology. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical map of the Pativilca River Basin (<b>a</b>) study area in Peru; (<b>b</b>) study area in Ancash and Lima regions; (<b>c</b>) study area with elevation and rivers in the basin.</p>
Full article ">Figure 2
<p>Spatial distribution of slope, land cover, and type soil in the Pativilca Basin. (<b>a</b>) Shows how the slope changes, with steeper areas mostly up in the upper part of the basin; (<b>b</b>) maps out the land cover, including vegetation, farms, and urban spots; and (<b>c</b>) highlights the soil types, showing how they affect water retention and erosion throughout the basin.</p>
Full article ">Figure 3
<p>Methodological flowchart.</p>
Full article ">Figure 4
<p>Cluster dendrogram for the regionalization of catchments in the Pativilca Basin. The dendrogram delineates six distinct catchment groups (A–F), represented by color-coded branches. Each group’s representative catchment is highlighted in pink. The vertical axis reflects the degree of dissimilarity between the catchments, with greater heights indicating higher dissimilarity. This regionalization was achieved using hierarchical clustering based on Euclidean distances, facilitating the identification of hydrologically similar catchment groups for further analysis.</p>
Full article ">Figure 5
<p>Regionalization of watersheds in the Pativilca Basin and selection of representative watersheds.</p>
Full article ">Figure 6
<p>Seasonal variations in precipitation, maximum, and minimum temperatures in the Pativilca Basin regions. The first column (blue bars) represents monthly precipitation, while red bars indicate maximum temperatures and orange bars depict minimum temperatures. The groups are arranged vertically from top to bottom, starting with Group A at the uppermost position and concluding with Group F at the lowest. These graphs highlight the temporal distribution and variability in key climatic variables across different seasons, enabling the assessment of seasonal trends and their impact on hydrological processes in the basin.</p>
Full article ">Figure 7
<p>Calibration and validation at the Cahua hydrometric station.</p>
Full article ">Figure 8
<p>Spatial distribution of hydrological components in the Pativilca Basin. The hydrological components include (<b>a</b>) flow out daily mean (<span class="html-italic">Q</span>) and annual precipitation (<span class="html-italic">R<sub>d</sub></span>), (<b>b</b>) evapotranspiration (ET), (<b>c</b>) percolation (<span class="html-italic">W<sub>seep</sub></span>), (<b>d</b>) groundwater contribution to streamflow (<span class="html-italic">Q<sub>gw</sub></span>), (<b>e</b>) average daily soil water storage (SW), and (<b>f</b>) water yield (<span class="html-italic">W<sub>YLD</sub></span>). Each map illustrates the spatial variability across the basin, highlighting the hydrological dynamics. The representative watersheds are bordered in red, indicating their respective groups at the center. Group boundaries are depicted with black dotted lines, enhancing the differentiation between zones. These visual elements allow for a detailed analysis of the distribution and influence of key hydrological processes across the basin’s distinct regions.</p>
Full article ">Figure 9
<p>Temporal distribution of streamflow in the Pativilca Basin. Daily streamflows from 1981 to 2015 show a clear seasonal pattern, with peak flows during the wet season (January to March) and lows in the dry season (June to September). Flow variation is driven by altitude, storage capacity, and watershed connectivity, with lower watersheds redistributing water from upstream areas.</p>
Full article ">Figure 9 Cont.
<p>Temporal distribution of streamflow in the Pativilca Basin. Daily streamflows from 1981 to 2015 show a clear seasonal pattern, with peak flows during the wet season (January to March) and lows in the dry season (June to September). Flow variation is driven by altitude, storage capacity, and watershed connectivity, with lower watersheds redistributing water from upstream areas.</p>
Full article ">
23 pages, 1046 KiB  
Article
Forecasting Short- and Long-Term Wind Speed in Limpopo Province Using Machine Learning and Extreme Value Theory
by Kgothatso Makubyane and Daniel Maposa
Forecasting 2024, 6(4), 885-907; https://doi.org/10.3390/forecast6040044 - 4 Oct 2024
Abstract
This study investigates wind speed prediction using advanced machine learning techniques, comparing the performance of Vanilla long short-term memory (LSTM) and convolutional neural network (CNN) models, alongside the application of extreme value theory (EVT) using the r-largest order generalised extreme value distribution ( [...] Read more.
This study investigates wind speed prediction using advanced machine learning techniques, comparing the performance of Vanilla long short-term memory (LSTM) and convolutional neural network (CNN) models, alongside the application of extreme value theory (EVT) using the r-largest order generalised extreme value distribution (GEVDr). Over the past couple of decades, the academic literature has transitioned from conventional statistical time series models to embracing EVT and machine learning algorithms for the modelling of environmental variables. This study adds value to the literature and knowledge of modelling wind speed using both EVT and machine learning. The primary aim of this study is to forecast wind speed in the Limpopo province of South Africa to showcase the dependability and potential of wind power generation. The application of CNN showcased considerable predictive accuracy compared to the Vanilla LSTM, achieving 88.66% accuracy with monthly time steps. The CNN predictions for the next five years, in m/s, were 9.91 (2024), 7.64 (2025), 7.81 (2026), 7.13 (2027), and 9.59 (2028), slightly outperforming the Vanilla LSTM, which predicted 9.43 (2024), 7.75 (2025), 7.85 (2026), 6.87 (2027), and 9.43 (2028). This highlights CNN’s superior ability to capture complex patterns in wind speed dynamics over time. Concurrently, the analysis of the GEVDr across various order statistics identified GEVDr=2 as the optimal model, supported by its favourable evaluation metrics in terms of Akaike information criteria (AIC) and Bayesian information criteria (BIC). The 300-year return level for GEVDr=2 was found to be 22.89 m/s, indicating a rare wind speed event. Seasonal wind speed analysis revealed distinct patterns, with winter emerging as the most efficient season for wind, featuring a median wind speed of 7.96 m/s. Future research could focus on enhancing prediction accuracy through hybrid algorithms and incorporating additional meteorological variables. To the best of our knowledge, this is the first study to successfully combine EVT and machine learning for short- and long-term wind speed forecasting, providing a novel framework for reliable wind energy planning. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic location of Polokwane within Limpopo province. Source: <a href="https://w.wiki/B2us" target="_blank">https://w.wiki/B2us</a>, accessed on 5 May 2024.</p>
Full article ">Figure 2
<p>Basic architecture of the Vanilla LSTM network. Source: <a href="https://www.researchgate.net/figure/The-structure-of-LSTM-memory-cell_fig5_342998863" target="_blank">https://www.researchgate.net/figure/The-structure-of-LSTM-memory-cell_fig5_342998863</a>, accessed on 30 June 2024.</p>
Full article ">Figure 3
<p>Basic architecture of a CNN. Source: <a href="https://learnopencv.com/wp-content/uploads/2023/01/tensorflow-keras-cnn-vgg-architecture.png" target="_blank">https://learnopencv.com/wp-content/uploads/2023/01/tensorflow-keras-cnn-vgg-architecture.png</a>, accessed on 18 September 2024.</p>
Full article ">Figure 4
<p>Seasonal wind speed trend analysis for 2023.</p>
Full article ">Figure 5
<p>Trend slope plot.</p>
Full article ">Figure 6
<p>Training and testing predictions of Vanilla LSTM and CNN models.</p>
Full article ">Figure 7
<p>Boxplot of the residual errors of each method with different ranges of the wind speed.</p>
Full article ">Figure 8
<p>Training and validation loss of Vanilla LSTM and CNN models.</p>
Full article ">Figure 9
<p>(<b>Top left panel</b>): Vanilla LSTM daily wind speed predictions; (<b>Top right panel</b>): CNN daily wind speed predictions; (<b>Bottom left panel</b>): Vanilla LSTM yearly wind speed predictions; (<b>Bottom right panel</b>): CNN yearly wind speed predictions.</p>
Full article ">Figure 10
<p><math display="inline"><semantics> <mrow> <mi>G</mi> <mi>E</mi> <mi>V</mi> <msub> <mi>D</mi> <mrow> <mi>r</mi> <mo>=</mo> <mn>2</mn> </mrow> </msub> </mrow> </semantics></math> diagnostic plots.</p>
Full article ">
21 pages, 3496 KiB  
Article
Study of the Gendered Impacts of Climate Change in Bol, Lake Province, Chad
by Exaucé Gali Djako, Evelyne Mendy, Semingar Ngaryamgaye, Komi Sélom Klassou and Jérôme Chenal
Climate 2024, 12(10), 157; https://doi.org/10.3390/cli12100157 - 4 Oct 2024
Abstract
Climate change is a global phenomenon impacting ecosystems, economies, and livelihoods. This research carried out in Bol in the Lake Province of Chad, a region heavily affected by climate change, aims to analyze the gender-differentiated impacts of this phenomenon. It was carried out [...] Read more.
Climate change is a global phenomenon impacting ecosystems, economies, and livelihoods. This research carried out in Bol in the Lake Province of Chad, a region heavily affected by climate change, aims to analyze the gender-differentiated impacts of this phenomenon. It was carried out using the rapid analysis and participatory planning (RAPP) method and structural analysis for social systems (SAS2). Meteorological and socioeconomic data were collected through interviews, household surveys, and focus groups. The results indicate variability in rainfall, with a slight downward trend and an increase in temperature. The women identified an increase in the cost of living, human and material losses, warmer housing, and health problems as socioeconomic socioeconomic consequences of climate change. Their coping strategies include community self-help, humanitarian aid, and welfare activities. Obstacles to full participation in the search for solutions include access to education, low decision-making power, and political representation. This research enriches our understanding of the interactions between gender, climate change, adaptation, and inclusive policy importance. Full article
Show Figures

Figure 1

Figure 1
<p>The geographical location of the city of Bol.</p>
Full article ">Figure 2
<p>Umbrothermal diagram of Bol.</p>
Full article ">Figure 3
<p>Standardized Precipitation Indices of Bol from 1970 to 2023.</p>
Full article ">Figure 4
<p>Trends in annual rainfall between 1970 and 2023.</p>
Full article ">Figure 5
<p>Changes in mean annual temperature from 1970 to 2023.</p>
Full article ">Figure 6
<p>Knowledge of the climate change concept.</p>
Full article ">Figure 7
<p>Channels through which climate change knowledge is acquired.</p>
Full article ">Figure 8
<p>Perception of climate change manifestation.</p>
Full article ">Figure 9
<p>Socioeconomic consequences of climate change.</p>
Full article ">Figure 10
<p>Drought adaptation strategies. MAM: Mutual aid between members of the community; NLF: Non-Local Food Using; INC: Introduction of New Crops; SRA: Soil Restoration by Amendment; ACP: Abandonment of Cultivated Plots; EAC: Economic Activity Changing; PCD: Practice of flood crops; IPU: Irrigated Polders Using; LFA: Livestock Feed Adaptation; MGT: Migration; SLS: Sale of live cattle; PAR: Trees Protection and Reforestation.</p>
Full article ">Figure 11
<p>Urban heat adaptation strategies. STY: Stay Hydrated (water and tea); WAC: Wearing Appropriate Clothing, particularly white boubous and turbans.; CHC: Care in Health Centers; TPR: Tree Protection and Reforestation; CLH: Construction of straw and rammed earth roofs (houses known locally as Dourdour); CSS: Construction of straw sheds; ROD: Rest Outside Dwellings (days and nights).</p>
Full article ">Figure 12
<p>Flood coping strategies. MAC: Mutual aid between members of the community; BMH: Building Makeshift Housing; MFU: Manufactured Food Using; NLF: Non-Local Food Using; MPA: Moving to Peri-urban Areas.SPP: Searching for pasture by pirogue; RDW: Reconstruction of dwellings; RHA: Recourse of humanitarian assistance.</p>
Full article ">
19 pages, 4910 KiB  
Article
The Dynamics of Vegetation Evapotranspiration and Its Response to Surface Meteorological Factors in the Altay Mountains, Northwest China
by Aishajiang Aili, Xu Hailiang, Abdul Waheed, Zhao Wanyu, Xu Qiao, Zhao Xinfeng and Zhang Peng
Sustainability 2024, 16(19), 8608; https://doi.org/10.3390/su16198608 - 3 Oct 2024
Abstract
The Altay Mountains’ forests are vital to Xinjiang’s terrestrial ecosystem, especially water regulation and conservation. This study evaluates vegetation evapotranspiration (ET) from 2000 to 2017 using temperature, precipitation, and ET data from the China Meteorological Data Sharing Service. The dataset underwent quality control [...] Read more.
The Altay Mountains’ forests are vital to Xinjiang’s terrestrial ecosystem, especially water regulation and conservation. This study evaluates vegetation evapotranspiration (ET) from 2000 to 2017 using temperature, precipitation, and ET data from the China Meteorological Data Sharing Service. The dataset underwent quality control and was interpolated using the inverse distance weighted (IDW) method. Correlation analysis and climate trend methodologies were applied to assess the impacts of temperature, precipitation, drought, and extreme weather events on ET. The results indicate that air temperature had a minimal effect on ET, with 68.34% of the region showing weak correlations (coefficients between −0.2 and 0.2). Conversely, precipitation exhibited a strong positive correlation with ET across 98.91% of the area. Drought analysis, using the standardized precipitation evapotranspiration index (SPEI) and the Temperature Vegetation Dryness Index (TVDI), showed that ET was significantly correlated with the SPEI in 96.47% of the region, while the TVDI displayed both positive and negative correlations. Extreme weather events also significantly influenced ET, with reductions in the Simple Daily Intensity Index (SDII), heavy precipitation days (R95p, R10), and increases in indicators like growing season length (GSL) and warm spell duration index (WSDI) leading to variations in ET. Based on the correlation coefficients and their significance, it was confirmed that the SII (precipitation intensity) and R95p (heavy precipitation) are the main factors causing vegetation ET increases. These findings offer crucial insights into the interactions between meteorological variables and ET, essential information for sustainable forest management, by highlighting the importance of optimizing water regulation strategies, such as adjusting species composition and forest density to enhance resilience against drought and extreme weather, thereby ensuring long-term forest health and productivity in response to climate change. Full article
14 pages, 829 KiB  
Article
Optimization of an N2O Emission Flux Model Based on a Variable-Step Drosophila Algorithm
by Lixia Dong, Shujia Mu and Guang Li
Agronomy 2024, 14(10), 2279; https://doi.org/10.3390/agronomy14102279 - 3 Oct 2024
Abstract
The application of intelligent process-based crop model parameter optimization algorithms can effectively improve both the model simulation accuracy and applicability. Based on measured values of soil N2O emission flux in wheat fields from 2020 to 2022, and meteorological data from 1971 [...] Read more.
The application of intelligent process-based crop model parameter optimization algorithms can effectively improve both the model simulation accuracy and applicability. Based on measured values of soil N2O emission flux in wheat fields from 2020 to 2022, and meteorological data from 1971 to 2022, five parameters of the N2O emission flux module in the APSIM model were optimized using the variable step Fruit Fly algorithm (VSS-FOA). The optimized parameters were the soil nitrification potential, the range of concentrated KNH4 of ammonia and nitrogen at semi-maximum utilization efficiency, the proportion of nitrogen loss to N2O during the nitrification process, the denitrification coefficient, and the Power term P for calculating the denitrification water coefficient. Contrasting the optimized parameters using the VSS-FOA algorithm versus the default values supplied with the model substantially improved the goodness-of-fit to field measurements with the overall R2 increasing from 0.41 to 0.74, and a decrease in NRMSE from 17.1% to 11.4%. This work demonstrates that the VSS-FOA algorithm affords a straightforward mechanism for the optimization of parameters in models such as APSIM to enhance the accuracy of model N2O emission flux estimates. Full article
(This article belongs to the Section Precision and Digital Agriculture)
16 pages, 8241 KiB  
Article
Tracking the Development of Lit Fisheries by Using DMSP/OLS Data in the Open South China Sea
by Jiajun Li, Zhixin Zhang, Kui Zhang, Jiangtao Fan, Huaxue Liu, Yongsong Qiu, Xi Li and Zuozhi Chen
Remote Sens. 2024, 16(19), 3678; https://doi.org/10.3390/rs16193678 - 2 Oct 2024
Abstract
Nightly images offer a special data source for monitoring fishing activities. This study used images from the Defense Meteorological Satellite Program (DMSP)/Operational Linescan System (OLS) to analyze the early development of lit fisheries in the open South China Sea (SCS), which mainly occurred [...] Read more.
Nightly images offer a special data source for monitoring fishing activities. This study used images from the Defense Meteorological Satellite Program (DMSP)/Operational Linescan System (OLS) to analyze the early development of lit fisheries in the open South China Sea (SCS), which mainly occurred around the Zhong Sha and Xi Sha Islands. Based on peak detection and a fixed threshold, lit fishing positions were extracted well from filtered, high-quality DMSP/OLS images. The results indicated that fisheries experienced an apparent rise and fall from 2005 to 2012, with the numbers of lit fishing boats rising to a maximum of ~60 from 2005 to 2008, almost disappearing in 2009, peaking at ~130 from 2010 to 2011, and starting to decline in 2012. The fish price of major fishing targets declined by ~60% in 2009, which obviously impacted the year’s fishing operations. The reason for declined fishing operations in 2012 was that most of the lit fishing operations shifted farther south to fishing grounds around the Nan Sha Islands. We also explored factors shaping the distribution patterns of lit fisheries by using MaxEnt models to relate fishing positions to environmental variables. Major environmental factors influencing the distribution of lit fishing boats varied with years, of which water depth was the most important factor across years, with an optimal depth range of 1000–2000 m. In addition to depth, the distribution of lit fisheries was also influenced by SST, especially for the years 2005–2008, and a suitable SST was found between 26 and 28 °C. This study fills the knowledge gaps of the inception of lit fisheries and their dynamic changes in the SCS. Full article
Show Figures

Figure 1

Figure 1
<p>Typical DMSP/OLS nighttime low-light imaging with a bright stripe. A stripe usually appears at a specific location along the scan line.</p>
Full article ">Figure 2
<p>Flow chart for extraction of lit fishing boats from DMSP/OLS images.</p>
Full article ">Figure 3
<p>Correlations among the eight marine predictors.</p>
Full article ">Figure 4
<p>Boxplots of number of lit fishing boats derived from nighttime images of Zhong Sha and Xi Sha fishing ground. Development of lit fisheries was classified into four stages (I–IV).</p>
Full article ">Figure 5
<p>Maps showing habitat predictions by the best models for lit fisheries from 2005–2011, where subfigures (<b>A</b>–<b>F</b>) correspond to the years 2005, 2006, 2007, 2008, 2010, and 2011, respectively. The maps display the habitat suitability index, with fishing boat positions for April indicated by the overlaid dots.</p>
Full article ">Figure 6
<p>Relative importance of environmental variables determined by MaxEnt models, with subfigures (<b>A</b>–<b>F</b>) corresponding to the years 2005, 2006, 2007, 2008, 2010, and 2011, respectively.</p>
Full article ">Figure 7
<p>Frequency distribution of number of fishing positions against water depth for Zhong Sha and Xi Sha fishing ground, with subfigures (<b>A</b>–<b>F</b>) corresponding to the years 2005, 2006, 2007, 2008, 2010, and 2011, respectively.</p>
Full article ">Figure 8
<p>Frequency distribution of number of fishing positions against SST for Zhong Sha and Xi Sha fishing ground, with subfigures (<b>A</b>–<b>F</b>) corresponding to the years 2005, 2006, 2007, 2008, 2010, and 2011, respectively.</p>
Full article ">Figure 9
<p>Main targets of lit fisheries in SCS by two types of gear (purse seine and falling net) and three fishing grounds (northern shelf, oceanic waters in Zhong Sha and Xi Sha Islands, and in Nan Sha Islands) [<a href="#B10-remotesensing-16-03678" class="html-bibr">10</a>,<a href="#B13-remotesensing-16-03678" class="html-bibr">13</a>,<a href="#B14-remotesensing-16-03678" class="html-bibr">14</a>].</p>
Full article ">Figure 10
<p>Typical nighttime imagery of fishing in the Zhong Sha and Xi Sha fishing ground in the spring season of 2008 (<b>A</b>) and 2009 (<b>B</b>).</p>
Full article ">
15 pages, 2402 KiB  
Article
Assessment of Hydrometeorological Impacts of Climate Change on Water Bodies in Northern Kazakhstan
by Baurzhan Yessenzholov, Abilzhan Khussainov, Anuarbek Kakabayev, Ivan Plachinta, Zulfiya Bayazitova, Gulmira Kyzdarbekova, Uzak Zhamkenov and Makhabbat Ramazanova
Water 2024, 16(19), 2794; https://doi.org/10.3390/w16192794 - 1 Oct 2024
Abstract
This article examines the impact of climate change on the hydrometeorological indicators of some lakes and reservoirs in the Akmola and North Kazakhstan regions. Two meteorological variables’ annual and seasonal trends at three weather stations in 1986–2023 were analyzed. The non-parametric Mann–Kendall and [...] Read more.
This article examines the impact of climate change on the hydrometeorological indicators of some lakes and reservoirs in the Akmola and North Kazakhstan regions. Two meteorological variables’ annual and seasonal trends at three weather stations in 1986–2023 were analyzed. The non-parametric Mann–Kendall and Sen’s slope methods were used to determine the presence of a positive or negative trend in weather data and their statistical significance. Hydrometric indicators were studied using the ArcGIS 10.8 program from 1995 to 2023. The results indicate an increasing average spring air temperature, with an annual rise of 0.08–0.09 °C. A significant trend in increasing average annual precipitation was observed in Saumalkol, with a rise of 4.7 mm per year. In contrast, no significant trends were found in the annual and seasonal precipitation data for Sergeyevka. It was also found that the area of Lake Saumalkol increased by 1.6% due to a rise in annual precipitation. In contrast, the area of Lake Kopa decreased by 6.04% because of an increase in the annual average temperature. Full article
(This article belongs to the Section Water and Climate Change)
Show Figures

Figure 1

Figure 1
<p>Trends in changes in average annual temperature in Kokshetau.</p>
Full article ">Figure 2
<p>Trends in changes in average annual temperature in Sergeyevka.</p>
Full article ">Figure 3
<p>Trends in changes in average annual temperature in Saumalkol.</p>
Full article ">Figure 4
<p>Annual precipitation from 1986 to 2023.</p>
Full article ">Figure 5
<p>Map of changes in the shorelines of Lake Kopa (<b>A</b>) and Shagalaly reservoir (<b>B</b>).</p>
Full article ">Figure 6
<p>Map of changes in the shorelines of Lake Saumalkol.</p>
Full article ">Figure 7
<p>Map of changes in the shorelines of the Sergeyevka reservoir.</p>
Full article ">
25 pages, 7359 KiB  
Article
Local Weather and Global Climate Data-Driven Long-Term Runoff Forecasting Based on Local–Global–Temporal Attention Mechanisms and Graph Attention Networks
by Binlin Yang, Lu Chen, Bin Yi, Siming Li and Zhiyuan Leng
Remote Sens. 2024, 16(19), 3659; https://doi.org/10.3390/rs16193659 - 30 Sep 2024
Abstract
The accuracy of long-term runoff models can be increased through the input of local weather variables and global climate indices. However, existing methods do not effectively extract important information from complex input factors across various temporal and spatial dimensions, thereby contributing to inaccurate [...] Read more.
The accuracy of long-term runoff models can be increased through the input of local weather variables and global climate indices. However, existing methods do not effectively extract important information from complex input factors across various temporal and spatial dimensions, thereby contributing to inaccurate predictions of long-term runoff. In this study, local–global–temporal attention mechanisms (LGTA) were proposed for capturing crucial information on global climate indices on monthly, annual, and interannual time scales. The graph attention network (GAT) was employed to extract geographical topological information of meteorological stations, based on remotely sensed elevation data. A long-term runoff prediction model was established based on long-short-term memory (LSTM) integrated with GAT and LGTA, referred to as GAT–LGTA–LSTM. The proposed model was compared to five comparative models (LGTA–LSTM, GAT–GTA–LSTM, GTA–LSTM, GAT–GA–LSTM, GA–LSTM). The models were applied to forecast the long-term runoff at Luning and Pingshan stations in China. The results indicated that the GAT–LGTA–LSTM model demonstrated the best forecasting performance among the comparative models. The Nash–Sutcliffe Efficiency (NSE) of GAT–LGTA–LSTM at the Luning and Pingshan stations reached 0.87 and 0.89, respectively. Compared to the GA–LSTM benchmark model, the GAT–LGTA–LSTM model demonstrated an average increase in NSE of 0.07, an average increase in Kling–Gupta Efficiency (KGE) of 0.08, and an average reduction in mean absolute percent error (MAPE) of 0.12. The excellent performance of the proposed model is attributed to the following: (1) local attention mechanism assigns a higher weight to key global climate indices at a monthly scale, enhancing the ability of global and temporal attention mechanisms to capture the critical information at annual and interannual scales and (2) the global attention mechanism integrated with GAT effectively extracts crucial temporal and spatial information from precipitation and remotely-sensed elevation data. Furthermore, attention visualization reveals that various global climate indices contribute differently to runoff predictions across distinct months. The global climate indices corresponding to specific seasons or months should be selected to forecast the respective monthly runoff. Full article
21 pages, 4017 KiB  
Article
A Machine Learning-Based Sustainable Energy Management of Wind Farms Using Bayesian Recurrent Neural Network
by Aisha Blfgeh and Hanadi Alkhudhayr
Sustainability 2024, 16(19), 8426; https://doi.org/10.3390/su16198426 - 27 Sep 2024
Abstract
The sustainable management of energy sources such as wind plays a crucial role in supplying electricity for both residential and industrial purposes. For this, accurate wind data are essential to bring sustainability in energy output estimations for wind stations. The choice of an [...] Read more.
The sustainable management of energy sources such as wind plays a crucial role in supplying electricity for both residential and industrial purposes. For this, accurate wind data are essential to bring sustainability in energy output estimations for wind stations. The choice of an appropriate distribution function significantly affects the actual wind data, directly influencing the estimated energy output. While the Weibull function is commonly used to describe wind speed at various locations worldwide, the variability of weather information across wind sites varies significantly. Probabilistic forecasting offers comprehensive probability information for renewable generation and load, assisting decision-making in power systems under uncertainty. Traditional probabilistic forecasting techniques based on machine learning (ML) rely on prediction uncertainty derived from previous distributional assumptions. This study utilized a Bayesian Recurrent Neural Network (BNN-RNN), incorporating prior distributions for weight variables in the RNN network layer and extending the Bayesian networks. Initially, a periodic RNN processes data for wind energy prediction, capturing trends and correlation characteristics in time-series data to enable more accurate and reliable energy production forecasts. Subsequently, the wind power meteorological dataset was analyzed using the reciprocal entropy approach to reduce dimensionality and eliminate variables with weak connections, thereby simplifying the structure of the prediction model. The BNN-RNN prediction model integrates inputs from RNN-transformed time-series data, dimensionality-reduced weather information, and time categorization feature data. The Winkler index is lower by 3.4%, 32.6%, and 7.2%, respectively, and the overall index of probability forecasting pinball loss is reduced by 51.2%, 22.3%, and 10.7%, respectively, compared with all three approaches. The implications of this study are significant, as they demonstrate the potential for more accurate wind energy forecasting through Bayesian optimization. These findings contribute to more precise decision-making and bring sustainability to the effective management of energy systems by proposing a Bayesian Recurrent Neural Network (BNN-RNN) to improve wind energy forecasts. The model further enhances future estimates of wind energy generation, considering the stochastic nature of meteorological data. The study is crucial in increasing the understanding and application of machine learning by establishing how Bayesian optimization significantly improves probabilistic forecasting models that would revolutionize sustainable energy management. Full article
(This article belongs to the Special Issue Renewable Energy, Electric Power Systems and Sustainability)
Show Figures

Figure 1

Figure 1
<p>Electricity transaction flow [<a href="#B14-sustainability-16-08426" class="html-bibr">14</a>].</p>
Full article ">Figure 2
<p>Methodology workflow.</p>
Full article ">Figure 3
<p>Bayesian Neural Network (BNN) model architecture [<a href="#B37-sustainability-16-08426" class="html-bibr">37</a>].</p>
Full article ">Figure 4
<p>Basic structure of BNN.</p>
Full article ">Figure 5
<p>Structure of RNN.</p>
Full article ">Figure 6
<p>Winkler index and pinball loss comparison [<a href="#B29-sustainability-16-08426" class="html-bibr">29</a>].</p>
Full article ">Figure 7
<p>MAE Comparison.</p>
Full article ">Figure 8
<p>RMSE Comparison.</p>
Full article ">Figure 9
<p>MSE Comparison.</p>
Full article ">
27 pages, 13823 KiB  
Article
Application of Remote Sensing and Explainable Artificial Intelligence (XAI) for Wildfire Occurrence Mapping in the Mountainous Region of Southwest China
by Jia Liu, Yukuan Wang, Yafeng Lu, Pengguo Zhao, Shunjiu Wang, Yu Sun and Yu Luo
Remote Sens. 2024, 16(19), 3602; https://doi.org/10.3390/rs16193602 - 27 Sep 2024
Abstract
The ecosystems in the mountainous region of Southwest China are exceptionally fragile and constitute one of the global hotspots for wildfire occurrences. Understanding the complex interactions between wildfires and their environmental and anthropogenic factors is crucial for effective wildfire modeling and management. Despite [...] Read more.
The ecosystems in the mountainous region of Southwest China are exceptionally fragile and constitute one of the global hotspots for wildfire occurrences. Understanding the complex interactions between wildfires and their environmental and anthropogenic factors is crucial for effective wildfire modeling and management. Despite significant advancements in wildfire modeling using machine learning (ML) methods, their limited explainability remains a barrier to utilizing them for in-depth wildfire analysis. This paper employs Logistic Regression (LR), Random Forest (RF), and Extreme Gradient Boosting (XGBoost) models along with the MODIS global fire atlas dataset (2004–2020) to study the influence of meteorological, topographic, vegetation, and human factors on wildfire occurrences in the mountainous region of Southwest China. It also utilizes Shapley Additive exPlanations (SHAP) values, a method within explainable artificial intelligence (XAI), to demonstrate the influence of key controlling factors on the frequency of fire occurrences. The results indicate that wildfires in this region are primarily influenced by meteorological conditions, particularly sunshine duration, relative humidity (seasonal and daily), seasonal precipitation, and daily land surface temperature. Among local variables, altitude, proximity to roads, railways, residential areas, and population density are significant factors. All models demonstrate strong predictive capabilities with AUC values over 0.8 and prediction accuracies ranging from 76.0% to 95.0%. XGBoost outperforms LR and RF in predictive accuracy across all factor groups (climatic, local, and combinations thereof). The inclusion of topographic factors and human activities enhances model optimization to some extent. SHAP results reveal critical features that significantly influence wildfire occurrences, and the thresholds of positive or negative changes, highlighting that relative humidity, rain-free days, and land use land cover changes (LULC) are primary contributors to frequent wildfires in this region. Based on regional differences in wildfire drivers, a wildfire-risk zoning map for the mountainous region of Southwest China is created. Areas identified as high risk are predominantly located in the Northwestern and Southern parts of the study area, particularly in Yanyuan and Miyi, while areas assessed as low risk are mainly distributed in the Northeastern region. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the research region and the distribution of MODIS active fire incidents from 2004 to 2020. Maps at a national scale represent the kernel density of local wildfires for the same time frame.</p>
Full article ">Figure 2
<p>Hierarchical importance of climatic variables.</p>
Full article ">Figure 3
<p>Hierarchical importance of local factors.</p>
Full article ">Figure 4
<p>The SHAP summary plot ranks the top 20 variables affecting model predictions by their mean absolute SHAP values, shown on the <span class="html-italic">y</span>-axis. Subfigure (<b>a</b>) showcases the importance of these features, while subfigure (<b>b</b>) illustrates their positive or negative effects on wildfire predictions through scatter points.</p>
Full article ">Figure 5
<p>The SHAP dependence plots (<b>a</b>) between SHAP values and Da_minRH, with a fitted trend line (red line); (<b>b</b>) between SHAP values and Norainday_avg, with a fitted trend line (red line); (<b>c</b>) between SHAP values and Da_minRH, showing the interaction with Tmax_avg (color scale); (<b>d</b>) between SHAP values and Norainday_avg, showing the interaction with Tmax_avg (color scale). Da_minRH, daily minimum relative humidity; Noraindy_avg, average number of rainless days of fire season.</p>
Full article ">Figure 6
<p>SHAP interaction plot (<b>a</b>) and heatmap analysis (<b>b</b>).</p>
Full article ">Figure 6 Cont.
<p>SHAP interaction plot (<b>a</b>) and heatmap analysis (<b>b</b>).</p>
Full article ">Figure 7
<p>Fire-occurrence probability: analysis using LR, RF, and XGB based on meteorological factors.</p>
Full article ">Figure 8
<p>Fire-occurrence probability: analysis using LR, RF, and XGB based on local factors.</p>
Full article ">Figure 9
<p>Fire-occurrence probability: combined meteorological and local factors analysis with LR, RF, and XGB.</p>
Full article ">Figure 10
<p>ROC curves of the success rate of three models.</p>
Full article ">Figure 11
<p>Comparison of error metrics for different models.</p>
Full article ">Figure 12
<p>Risk-assessment mapping results of XGB model.</p>
Full article ">
27 pages, 25123 KiB  
Article
Evaluation of Reanalysis and Satellite Products against Ground-Based Observations in a Desert Environment
by Narendra Nelli, Diana Francis, Abdulrahman Alkatheeri and Ricardo Fonseca
Remote Sens. 2024, 16(19), 3593; https://doi.org/10.3390/rs16193593 - 26 Sep 2024
Abstract
The Arabian Peninsula (AP) is notable for its unique meteorological and climatic patterns and plays a pivotal role in understanding regional climate dynamics and dust emissions. The scarcity of ground-based observations makes atmospheric data essential, rendering reanalysis and satellite products invaluable for understanding [...] Read more.
The Arabian Peninsula (AP) is notable for its unique meteorological and climatic patterns and plays a pivotal role in understanding regional climate dynamics and dust emissions. The scarcity of ground-based observations makes atmospheric data essential, rendering reanalysis and satellite products invaluable for understanding weather patterns and climate variability. However, the accuracy of these products in the AP’s desert environment has not been extensively evaluated. This study undertakes the first comprehensive validation of reanalysis products—the European Centre for Medium-Range Weather Forecasts’ European Reanalysis version 5 (ERA5) and ERA5 Land (ERA5L), along with Clouds and Earth’s Radiant Energy System (CERES) radiation fluxes—against measurements from the Liwa desert in the UAE. The data, collected during the Wind-blown Sand Experiment (WISE)–UAE field experiment from July 2022 to December 2023, includes air temperature and relative humidity at 2 m, 10 m wind speed, surface pressure, skin temperature, and net radiation fluxes. Our analysis reveals a strong agreement between ERA5/ERA5L and the observed diurnal T2m cycle, despite a warm night bias and cold day bias with a magnitude within 2 K. The wind speed analysis uncovered a bimodal distribution attributed to sea-breeze circulation and the nocturnal low-level jet, with the reanalysis overestimating the nighttime wind speeds by 2 m s−1. This is linked to biases in nighttime temperatures arising from an inaccurate representation of nocturnal boundary layer processes. The daytime cold bias contrasts with the excessive net radiation flux at the surface by about 50–100 W m−2, underscoring the challenges in the physical representation of land–atmosphere interactions. Full article
Show Figures

Figure 1

Figure 1
<p>Diurnal variations in meteorological parameters from ground-based observations and reanalysis products over a hyper-arid region in the United Arab Emirates. Parameters shown are the composite of (<b>a</b>–<b>d</b>) the 2 m temperature (T2m, K), (<b>e</b>–<b>h</b>) 2 m relative humidity (RH2m, %), (<b>i</b>–<b>l</b>) 10 m wind speed (WS10m, m s<sup>−1</sup>), and (<b>m</b>–<b>p</b>) surface pressure (hPa) for four seasons: December–January–February (DJF, first column), March–April–May (MAM, second column), June–July–August (JJA, third column), and September–October–November (SON, fourth column). Ground-based observations, ERA5, and ERA5 Land reanalysis data are depicted with black, blue, and red solid lines, respectively. Error bars indicate one standard deviation.</p>
Full article ">Figure 2
<p>Box plots of bias with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes): (<b>a</b>) 2 m air temperature (T2m), (<b>b</b>) 2 m relative humidity (RH2m), (<b>c</b>) 10 m wind speed (WS10m), and (<b>d</b>) surface pressure. The <b>top</b> (<b>bottom</b>) panel in each subplot indicates the bias in ERA5 (ERA5L) with respect to the observation.</p>
Full article ">Figure 2 Cont.
<p>Box plots of bias with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes): (<b>a</b>) 2 m air temperature (T2m), (<b>b</b>) 2 m relative humidity (RH2m), (<b>c</b>) 10 m wind speed (WS10m), and (<b>d</b>) surface pressure. The <b>top</b> (<b>bottom</b>) panel in each subplot indicates the bias in ERA5 (ERA5L) with respect to the observation.</p>
Full article ">Figure 3
<p>The same as <a href="#remotesensing-16-03593-f001" class="html-fig">Figure 1</a>, but for skin temperature (K).</p>
Full article ">Figure 4
<p>Box plots of bias in skin temperature with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes). (<b>a</b>) ERA5 and (<b>b</b>) ERA5L.</p>
Full article ">Figure 5
<p>The same as <a href="#remotesensing-16-03593-f001" class="html-fig">Figure 1</a>, but for net shortwave radiation (W m<sup>−2</sup>, (<b>a</b>–<b>d</b>)) and net longwave radiation (W m<sup>−2</sup>, (<b>e</b>–<b>h</b>)).</p>
Full article ">Figure 6
<p>Box plots of bias in net shortwave radiation flux with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes). The bias in (<b>a</b>) ERA5 and (<b>b</b>) ERA5L with respect to the observations. (<b>c</b>) Bias in CERES with respect to the observations.</p>
Full article ">Figure 7
<p>The same as <a href="#remotesensing-16-03593-f006" class="html-fig">Figure 6</a>, but for the net longwave radiation flux.</p>
Full article ">Figure 8
<p>Box plots of bias with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes) in daily minima/maxima of (<b>a</b>) 2 m air temperature (K), (<b>b</b>) skin temperature; (<b>c</b>) daily maximum in net shortwave radiation and net longwave radiation.</p>
Full article ">Figure 8 Cont.
<p>Box plots of bias with respect to in situ measurements for winter (blue boxes), spring (red boxes), summer (green boxes), and autumn (magenta boxes) in daily minima/maxima of (<b>a</b>) 2 m air temperature (K), (<b>b</b>) skin temperature; (<b>c</b>) daily maximum in net shortwave radiation and net longwave radiation.</p>
Full article ">Figure 9
<p>Correlation analysis of daily averaged in situ measurements with reanalysis data for key meteorological parameters in a hyper-arid region of the UAE. The first row (<b>a</b>–<b>d</b>) corresponds to ERA5 and the second row (<b>e</b>–<b>h</b>) to ERA5 Land. Each column represents a scatter plot for (<b>a</b>) T2m (K), (<b>b</b>) RH2m (%), (<b>c</b>) WS10m (m s<sup>−1</sup>), and (<b>d</b>) surface pressure (hPa). The solid black line indicates the 1:1 correspondence, serving as a reference for an ideal match.</p>
Full article ">Figure 10
<p>The same as <a href="#remotesensing-16-03593-f009" class="html-fig">Figure 9</a>, but for skin temperature (T<sub>skin</sub>, K). (<b>a</b>) ERA5 and (<b>b</b>) ERA5 Land.</p>
Full article ">Figure 11
<p>Correlation analysis of daily averaged in situ measurements with reanalysis data and remote sensing satellite (CERES) data for net shortwave and net longwave radiation fluxes in a hyper-arid region of the UAE. The first column (<b>a</b>–<b>d</b>) corresponds to ERA5, the second column (<b>b</b>–<b>e</b>) to ERA5 Land, and the third column (<b>c</b>–<b>f</b>) to CERES. The solid black line indicates the 1:1 correspondence, serving as a reference for an ideal match.</p>
Full article ">
4 pages, 1260 KiB  
Proceeding Paper
A Multivariate LSTM Model for Short-Term Water Demand Forecasting
by Aly K. Salem and Ahmed A. Abokifa
Eng. Proc. 2024, 69(1), 167; https://doi.org/10.3390/engproc2024069167 - 25 Sep 2024
Abstract
Accurate water demand forecasting is crucial for the effective operation and management of water distribution networks. Predicting future water demand empowers utilities to optimally operate system components. Various data-driven methodologies have been proposed for water demand forecasting, including artificial neural networks and econometric [...] Read more.
Accurate water demand forecasting is crucial for the effective operation and management of water distribution networks. Predicting future water demand empowers utilities to optimally operate system components. Various data-driven methodologies have been proposed for water demand forecasting, including artificial neural networks and econometric models. Recently, Long Short-Term Memory (LSTM) was shown to be particularly relevant for this application. Nevertheless, few studies have utilized multivariate-LSTM (M-LSTM) models for water demand forecasting. This study introduces an M-LSTM model incorporating historical water demands, meteorological data, and social variables to forecast short-term water demand. The proposed M-LSTM model performance was tested by applying it to the ten district metered areas (DMAs) case study of the Battle of Water Demand Forecasting (BWDF). The results demonstrated the model’s ability to accurately predict the hourly water demand one week in advance. The mean absolute error of the predictions ranged between 0.5 and 2.2 l/s (2.8% to 12.9% of the average demand). The results also showed a strong correlation between the prediction error and the variability of the water demand data. Full article
Show Figures

Figure 1

Figure 1
<p>The workflow of the proposed M-LSTM model.</p>
Full article ">Figure 2
<p>(<b>a</b>) Model results across different DMAs; (<b>b</b>) Correlation analysis results.</p>
Full article ">
12 pages, 4327 KiB  
Article
Effect of Meteorological Variables on Energy Demand in the Northeast and Southeast Regions of Brazil
by Helber Barros Gomes, Dirceu Luís Herdies, Luiz Fernando dos Santos, João Augusto Hackerott, Bruno Ribeiro Herdies, Fabrício Daniel dos Santos Silva, Maria Cristina Lemos da Silva, Mario Francisco Leal de Quadro, Robinson Semolini, Amanda Cortez, Bruna Schatz, Bruno Dantas Cerqueira and Djanilton Henrique Moura Junior
Energies 2024, 17(19), 4776; https://doi.org/10.3390/en17194776 - 24 Sep 2024
Abstract
Energy consumption demand has shown successive records during recent months, primarily associated with heat waves in almost all Brazilian states. The effects of climate change induced by global warming and the increasingly frequent occurrence of extreme events, mainly regarding temperature and precipitation, are [...] Read more.
Energy consumption demand has shown successive records during recent months, primarily associated with heat waves in almost all Brazilian states. The effects of climate change induced by global warming and the increasingly frequent occurrence of extreme events, mainly regarding temperature and precipitation, are associated with this increase in demand. In this sense, the impact of meteorological variables on load demand in some substations in the northeast and southeast of Brazil was analyzed, considering the historical series of energy injected into these substations. Fifteen substations were analyzed: three in the state of São Paulo, six in Bahia, three in Pernambuco, and three in Rio Grande do Norte. Initially, essential quality control was carried out on the energy injection data. The SAMeT data sets were used for the variable temperature, and Xavier was used for precipitation and relative humidity to obtain a homogeneous data series. Daily and monthly data were used for a detailed analysis of these variables in energy demand over the northeast and southeast regions of Brazil. Some regions were observed to be sensitive to the maximum temperature variable, while others were sensitive to the average temperature. On the other hand, few cases showed the highest correlation with the precipitation and relative humidity variables, with most cases being considered slight or close to zero. A more refined analysis was based on the type of consumers associated with each substation. These results showed that where consumption is more residential, the highest correlations were associated with maximum temperature in most stations in the northeast and average temperature in the southeast. In regions where consumption is primarily rural, the correlation observed with precipitation and relative humidity was the highest despite being negative. A more detailed analysis shows that rural production is associated with irrigation in these substations, which partly explains consumption, as when rainfall occurs, the demand for irrigation decreases, and thus energy consumption is reduced. Full article
(This article belongs to the Section B: Energy and Environment)
Show Figures

Figure 1

Figure 1
<p>The spatial location of the fifteen substations in the Neoenergia concession region.</p>
Full article ">Figure 2
<p>Time series of daily energy load injected into Neoenergia substations (blue), average temperature (orange) at Andradina, Francisco Morato, and Ubatuba I stations, and maximum temperature at other substations. Asa Branca (<b>a</b>), Barreiras Norte (<b>b</b>), Lobato (<b>c</b>), Rio Guara (<b>d</b>), Formoso (<b>e</b>), America Dourada II (<b>f</b>), Andradina (<b>g</b>), Francisco Morato (<b>h</b>), Ubatuba I (<b>i</b>), Pau Amarelo (<b>j</b>), Don Avelar (<b>l</b>), Campus (<b>m</b>), Jiqui (<b>n</b>), Mossoró III (<b>o</b>) and Natal (<b>p</b>). Units: Energy load injected in Megawatts and temperature in Celsius.</p>
Full article ">Figure 3
<p>Correlation between the injected energy load and temperature, daily averages (blue), and monthly averages (orange) for the fifteen substations.</p>
Full article ">Figure 4
<p>Daily time series of the energy load injected into the fifteen Neoenergia substations (blue) and daily precipitation series (orange) from [<a href="#B16-energies-17-04776" class="html-bibr">16</a>]. Asa Branca (<b>a</b>), Barreiras Norte (<b>b</b>), Lobato (<b>c</b>), Rio Guara (<b>d</b>), Formoso (<b>e</b>), America Dourada II (<b>f</b>), Andradina (<b>g</b>), Francisco Morato (<b>h</b>), Ubatuba I (<b>i</b>), Pau Amarelo (<b>j</b>), Don Avelar (<b>l</b>), Campus (<b>m</b>), Jiqui (<b>n</b>), Mossoró III (<b>o</b>) and Natal (<b>p</b>). Units: Energy load injected in Megawatts and precipitation in millimeters.</p>
Full article ">Figure 5
<p>Daily time series of the energy load injected into the Neoenergia substations (blue) and daily relative humidity (orange) for the Rio Guará (<b>a</b>), Formoso (<b>b</b>), and América Dourada II (<b>c</b>) substations. Units: Energy load injected in Megawatts and relative humidity (%).</p>
Full article ">Figure 6
<p>Correlation between the daily injected energy load for all the fifteen substations of Neonergia, daily precipitation (blue), and daily relative humidity (orange) from [<a href="#B16-energies-17-04776" class="html-bibr">16</a>].</p>
Full article ">
23 pages, 3534 KiB  
Article
Global Incidence of Diarrheal Diseases—An Update Using an Interpretable Predictive Model Based on XGBoost and SHAP: A Systematic Analysis
by Dan Liang, Li Wang, Shuang Liu, Shanglin Li, Xing Zhou, Yun Xiao, Panpan Zhong, Yanxi Chen, Changyi Wang, Shan Xu, Juan Su, Zhen Luo, Changwen Ke and Yingsi Lai
Nutrients 2024, 16(18), 3217; https://doi.org/10.3390/nu16183217 - 23 Sep 2024
Abstract
Background: Diarrheal disease remains a significant public health issue, particularly affecting young children and older adults. Despite efforts to control and prevent these diseases, their incidence continues to be a global concern. Understanding the trends in diarrhea incidence and the factors influencing these [...] Read more.
Background: Diarrheal disease remains a significant public health issue, particularly affecting young children and older adults. Despite efforts to control and prevent these diseases, their incidence continues to be a global concern. Understanding the trends in diarrhea incidence and the factors influencing these trends is crucial for developing effective public health strategies. Objective: This study aimed to explore the temporal trends in diarrhea incidence and associated factors from 1990 to 2019 and to project the incidence for the period 2020–2040 at global, regional, and national levels. We aimed to identify key factors influencing these trends to inform future prevention and control strategies. Methods: The eXtreme Gradient Boosting (XGBoost) model was used to predict the incidence from 2020 to 2040 based on demographic, meteorological, water sanitation, and sanitation and hygiene indicators. SHapley Additive exPlanations (SHAP) value was performed to explain the impact of variables in the model on the incidence. Estimated annual percentage change (EAPC) was calculated to assess the temporal trends of age-standardized incidence rates (ASIRs) from 1990 to 2019 and from 2020 to 2040. Results: Globally, both incident cases and ASIRs of diarrhea increased between 2010 and 2019. The incident cases are expected to rise from 2020 to 2040, while the ASIRs and incidence rates are predicted to slightly decrease. During the observed (1990–2019) and predicted (2020–2040) periods, adults aged 60 years and above exhibited an upward trend in incidence rate as age increased, while children aged < 5 years consistently had the highest incident cases. The SHAP framework was applied to explain the model predictions. We identified several risk factors associated with an increased incidence of diarrhea, including age over 60 years, yearly precipitation exceeding 3000 mm, temperature above 20 °C for both maximum and minimum values, and vapor pressure deficit over 1500 Pa. A decreased incidence rate was associated with relative humidity over 60%, wind speed over 4 m/s, and populations with above 80% using safely managed drinking water services and over 40% using safely managed sanitation services. Conclusions: Diarrheal diseases are still serious public health concerns, with predicted increases in the incident cases despite decreasing ASIRs globally. Children aged < 5 years remain highly susceptible to diarrheal diseases, yet the incidence rate in the older adults aged 60 plus years still warrants additional attention. Additionally, more targeted efforts to improve access to safe drinking water and sanitation services are crucial for reducing the incidence of diarrheal diseases globally. Full article
(This article belongs to the Section Nutritional Immunology)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>–<b>C</b>) The trends and projections of ASIRs per 100,000 in diarrheal diseases from 1990 to 2040 at the global level by genders ((<b>A</b>), males, (<b>B</b>) females, and (<b>C</b>) both genders combined). The open spots correspond to the observations between 1990 and 2019, and the pink shadow indicates the 95% UIs of the predictions. The mean of the predictions is displayed as a black line (solid line for 1990–2019 and dashed line for 2020 to 2040), and a vertical dashed line denotes the start year of the prediction. (<b>D</b>–<b>F</b>) The trends in the incident cases of diarrheal diseases from 1990 to 2040 at the global level for males (<b>D</b>), females (<b>E</b>), and both genders combined (<b>F</b>). The black error bar refers to the 95% UIs of the predictions. Abbreviations: ASIRs, age-standardized incidence rates; UIs, 95% uncertainty intervals.</p>
Full article ">Figure 2
<p>(<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) The distribution in ASIRs per 100,000 persons for diarrheal diseases at the national level in 1990 (<b>A</b>), 2019 (<b>B</b>), 2020 (<b>D</b>), and 2040 (<b>E</b>). (<b>C</b>,<b>F</b>) The EAPCs in ASIRs for diarrheal diseases at the national level from 1990 to 2019 (<b>C</b>) and from 2020 to 2040 (<b>F</b>). Abbreviations: ASIRs, age-standardized incidence rates; EAPCs, estimated annual percentage changes.</p>
Full article ">Figure 3
<p>(<b>A</b>–<b>C</b>) ASIRs per 100,000 persons of diarrheal diseases at the global and regional levels by SDIs from 1990 to 2019 for males (<b>A</b>), females (<b>B</b>), and both genders combined (<b>C</b>). The predictions based on SDI and ASIRs in all 21 GBD regions are displayed as the solid black line, and the grey shadow indicates the 95% UIs of the predictions. (<b>D</b>–<b>I</b>) ASIRs per 100,000 in diarrheal diseases at the national level by SDI in 1990 for male (<b>D</b>), female (<b>E</b>), and both genders combined (<b>F</b>) and in 2019 for male (<b>G</b>), female (<b>H</b>), and both genders combined (<b>I</b>). The predictions based on SDI and ASIRs in all 204 countries or territories are demonstrated as the solid black line, and the grey shadow indicates the 95% UIs of the predictions. Abbreviations: ASIRs, age-standardized incidence rates; SDI, socio-demographic index; GBD, global burden of disease.</p>
Full article ">Figure 4
<p>(<b>A</b>–<b>C</b>) The variation trends of incident cases and incidence rates for diarrheal diseases from 1990 to 2040 in all age groups in 20-year intervals at the global level for males (<b>A</b>), females (<b>B</b>), and both genders combined (<b>C</b>). A vertical dashed line denotes the start year of the prediction. (<b>D</b>–<b>G</b>) The variation trends of incident cases and incidence rates for diarrheal diseases in all age groups in 5-year intervals at the global level for both genders combined in 1990 (<b>D</b>), 2019 (<b>E</b>), 2020 (<b>F</b>), and 2040 (<b>G</b>).</p>
Full article ">Figure 5
<p>SHAP summary plot of feature contributions ranked by mean |SHAP| values and SHAP dependence plots for each feature in the XGBoost model predicting diarrheal incidence rate. (<b>A</b>) summary plot. Different colored dots stated the scaled feature values of all instances, with navy blue dots demonstrating high feature values and light sky blue dots expressing low feature values. (<b>B</b>–<b>N</b>) the dependence plot of the contribution to the model for age (<b>B</b>), gender (<b>C</b>), year (<b>D</b>), proportion of population using safely managed drinking water services (<b>E</b>), proportion of population using safely managed sanitation services (<b>F</b>), population (<b>G</b>), relative humidity (<b>H</b>), precipitation (<b>I</b>), mean temperature (<b>J</b>), minimum temperature (<b>K</b>), maximum temperature (<b>L</b>), vapor pressure deficit (<b>M</b>), and wind speed (<b>N</b>). Abbreviations: SHAP, SHapley Additive exPlanations.</p>
Full article ">
Back to TopTop