[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

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 (2,342)

Search Parameters:
Keywords = groundwater level

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
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 265
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 ">
21 pages, 12105 KiB  
Article
Assessment of Total Mercury Levels Emitted from ASGM into Soil and Groundwater in Chami Town, Mauritania
by Mohamed Mamoune Maha, Akito Matsuyama, Takahiko Arima and Atsushi Sainoki
Sustainability 2024, 16(18), 7992; https://doi.org/10.3390/su16187992 - 12 Sep 2024
Viewed by 315
Abstract
Artisanal and small-scale gold mining (ASGM) is a serious growing concern in Sub-Saharan Africa. In Mauritania, recent gold discoveries in the north and northwest have led to an increase in ASGM centers, reflecting trends across the region and posing considerable risks of mercury [...] Read more.
Artisanal and small-scale gold mining (ASGM) is a serious growing concern in Sub-Saharan Africa. In Mauritania, recent gold discoveries in the north and northwest have led to an increase in ASGM centers, reflecting trends across the region and posing considerable risks of mercury (Hg) contamination. Notwithstanding this fact, the extent of mercury contamination in the region remains unclear due to insufficient knowledge on the mechanisms of Hg dispersion in hyper-arid regions. In light of this, the present study aimed to acquire fundamental knowledge to elucidate the dispersion mechanism of mercury through conducting soil and groundwater sampling in and around Chami town, Mauritania, where ASGM activities have intensified. We analyzed 180 soil samples and 5 groundwater samples for total mercury (total Hg) using cold vapor atomic absorption spectrometry (CVAAS) and atomic fluorescence spectrometry (AFS) methods. The total Hg levels in soil samples ranged from 0.002 to 9.3 ppm, with the highest concentrations found at ASGM sites. Groundwater samples exhibited low total Hg levels (0.25–1.25 ng/L). The total Hg content in soil and groundwater samples was below Japanese standards, yet soil samples from hotspot points exceeded other international standards. Our study emphasizes the Hg dispersion patterns around Chami town, suggesting a gradual decrease in total Hg with increasing distance from ASGM sites and a potential influence of wind dynamics. The knowledge accumulated in this study provides essential insights into the Hg dispersion mechanisms in Chami town, laying the foundation for establishing a predictive model of Hg contamination in hyper-arid regions. Full article
(This article belongs to the Special Issue Sustainable Environmental Analysis of Soil and Water)
Show Figures

Figure 1

Figure 1
<p>Location of the sampling sites within the zone of study.</p>
Full article ">Figure 2
<p>Chart of the total Hg content in soil samples within the PNBA area.</p>
Full article ">Figure 3
<p>Chart of the total Hg in soil samples within the Chami town environs.</p>
Full article ">Figure 4
<p>Distribution map of the total Hg content in soil samples.</p>
Full article ">Figure 5
<p>Chart of the total Hg content in the ASGM sites (Hotspots).</p>
Full article ">Figure 6
<p>Total Hg average concentrations across the columns of the sampling grid, with Column 6 representing the east direction and Column 1 positioned to the west.</p>
Full article ">Figure 7
<p>Total Hg average concentrations across the rows of the sampling grid, where Row 1 represents the north direction and Row 6 is positioned to the south.</p>
Full article ">Figure 8
<p>Correlation between total Hg content and distance from the main ASGM site ’Grillage’ of PNBA and Chami town environs soil samples.</p>
Full article ">Figure 9
<p>Chart for moisture ratio in soil samples.</p>
Full article ">Figure 10
<p>Correlation chart between the total Hg concentrations and moisture content in soil samples.</p>
Full article ">Figure 11
<p>Comparison chart: (<b>a</b>) The percentage of total Hg in leachate versus the remaining total Hg in soil samples; (<b>b</b>) the total Hg content in soil samples compared to soil sample leachates.</p>
Full article ">Figure 12
<p>Distribution chart of the geo-accumulation index (Igeo) for the various sampling locations.</p>
Full article ">
13 pages, 3085 KiB  
Article
Study on the Impact of Groundwater and Soil Parameters on Tunnel Deformation and Sensitivity Analysis
by Yongxin Li, Zhimin Zhang, Jinyu Dong, Bobo Wang and Chuang Wang
Appl. Sci. 2024, 14(18), 8196; https://doi.org/10.3390/app14188196 - 12 Sep 2024
Viewed by 243
Abstract
Based on the Xiaolangdi North Bank Irrigation Area Project, this study combines numerical simulation and BP neural network methods to investigate the sensitivity of tunnel soil and its parameter inversion under continuous heavy rainfall. The research results indicate that changes in water-level and [...] Read more.
Based on the Xiaolangdi North Bank Irrigation Area Project, this study combines numerical simulation and BP neural network methods to investigate the sensitivity of tunnel soil and its parameter inversion under continuous heavy rainfall. The research results indicate that changes in water-level and soil strength parameters have a significant impact on the deformation of tunnel surrounding rock. By comparing the sensitivity factors of different parameters, the main parameter sensitivities affecting the displacement of tunnel surrounding rock were determined to be water level, internal friction angle, and cohesion. The mechanical characteristics of the tunnel construction process were analyzed using finite difference method numerical analysis software FLAC3D, and the results were used as a sample dataset for inversion analysis. Through neural network inverse analysis based on orthogonal design method, the cohesion and internal friction angle of loess layer ④, loess layer ④-1, and loess layer ⑤ were determined, and the data of groundwater level elevation were obtained. Field applications proved the effectiveness and rationality of this method. Full article
Show Figures

Figure 1

Figure 1
<p>Location map of the study area.</p>
Full article ">Figure 2
<p>Stratigraphic information and tunnel 3D model diagram. (<b>a</b>) Tunnel and stratigraphic distribution map (unit: m), (<b>b</b>) 3D computational model.</p>
Full article ">Figure 3
<p>Model calculation results. (<b>a</b>) Vertical direction (Z direction), (<b>b</b>) horizontal direction (X direction).</p>
Full article ">Figure 4
<p>Influence of water level variation on displacement. (<b>a</b>) Horizontal direction (X direction), (<b>b</b>) vertical direction (Z direction).</p>
Full article ">Figure 5
<p>Influence of friction angle variation on displacement. (<b>a</b>) Horizontal direction (X direction), (<b>b</b>) vertical direction (Z direction).</p>
Full article ">Figure 6
<p>Influence of cohesion variation on displacement. (<b>a</b>) Horizontal direction (X direction), (<b>b</b>) vertical direction (Z direction).</p>
Full article ">Figure 7
<p>Schematic diagram of the neural network model.</p>
Full article ">Figure 8
<p>Comparison of predicted results for training and testing sets.</p>
Full article ">Figure 9
<p>Contour plot of the forward calculation of vertical displacement.</p>
Full article ">
13 pages, 5823 KiB  
Article
Emerging Contaminants in Landfill Leachate and Groundwater: A Case Study of Hazardous Waste Landfill and Municipal Solid Waste Landfill in Northeastern China
by Nan Zhang, Zhihao Zhang, Chunyang Li, Jiani Yue, Yan Su, Weiguo Cheng, Shoushan Sun, Xi Chen, Deyu Shi and Bo Liu
Water 2024, 16(18), 2575; https://doi.org/10.3390/w16182575 - 11 Sep 2024
Viewed by 442
Abstract
Emerging contaminants (ECs) present a significant risk to both the ecological environment and human health. Landfill leachate (LL) often contains elevated EC levels, posing a potential risk to localized groundwater. This study aimed to characterize ECs in municipal solid waste landfills (MSWLs) and [...] Read more.
Emerging contaminants (ECs) present a significant risk to both the ecological environment and human health. Landfill leachate (LL) often contains elevated EC levels, posing a potential risk to localized groundwater. This study aimed to characterize ECs in municipal solid waste landfills (MSWLs) and hazardous waste landfills (HWLs) in northeast (NE) China. One and three HWLs and MSWLs in NE China with varying types, operational years, and impermeable layers were selected as case studies, respectively. Statistical analysis of 62 indicators of nine ECs in leachate and the groundwater environment indicated the presence of perfluorinated compounds (PFCs), antibiotics, alkylphenols (APs), and bisphenol A (BPA). The leachates of the four landfills exhibited elevated concentrations of ECs of 21.03 μg/L, 40.04 μg/L, 14.54 μg/L, and 43.05 μg/L for PFCs, antibiotics, Aps, and BPA, respectively. There was a positive correlation between the highest concentrations of ECs in groundwater and those in leachate as well as with operational duration of the landfill; in contrast, groundwater EC was negatively correlated with the degree of impermeability. This study can guide future management of ECs in landfills and hazardous waste sites in China, particularly in NE China. Full article
(This article belongs to the Special Issue Management of Solid Waste and Landfill Leachate)
Show Figures

Figure 1

Figure 1
<p>Maps showing the locations and sampling points at the four case-study landfill sites.</p>
Full article ">Figure 2
<p>A comparative analysis of the concentrations of emerging contaminants (ECs) in leachate from four landfills in northeast China.</p>
Full article ">Figure 3
<p>Concentrations and detection frequencies of emerging contaminants (ECs) in groundwater at four landfill sites in northeast China.</p>
Full article ">Figure 4
<p>Emerging contaminants (ECs) in leachate and groundwater at Landfill A.</p>
Full article ">Figure 5
<p>Comparison of concentrations of emerging contaminants (ECs) in leachate and groundwater of Landfill B.</p>
Full article ">Figure 6
<p>Comparison of concentrations of emerging contaminants (ECs) in leachate and groundwater of Landfill C.</p>
Full article ">Figure 7
<p>Comparison of concentrations of emerging contaminants (ECs) in leachate and groundwater of Landfill D.</p>
Full article ">
20 pages, 3519 KiB  
Article
The Implementation of Multimodal Large Language Models for Hydrological Applications: A Comparative Study of GPT-4 Vision, Gemini, LLaVa, and Multimodal-GPT
by Likith Anoop Kadiyala, Omer Mermer, Dinesh Jackson Samuel, Yusuf Sermet and Ibrahim Demir
Hydrology 2024, 11(9), 148; https://doi.org/10.3390/hydrology11090148 - 11 Sep 2024
Viewed by 563
Abstract
Large Language Models (LLMs) combined with visual foundation models have demonstrated significant advancements, achieving intelligence levels comparable to human capabilities. This study analyzes the latest Multimodal LLMs (MLLMs), including Multimodal-GPT, GPT-4 Vision, Gemini, and LLaVa, with a focus on hydrological applications such as [...] Read more.
Large Language Models (LLMs) combined with visual foundation models have demonstrated significant advancements, achieving intelligence levels comparable to human capabilities. This study analyzes the latest Multimodal LLMs (MLLMs), including Multimodal-GPT, GPT-4 Vision, Gemini, and LLaVa, with a focus on hydrological applications such as flood management, water level monitoring, agricultural water discharge, and water pollution management. We evaluated these MLLMs on hydrology-specific tasks, testing their response generation and real-time suitability in complex real-world scenarios. Prompts were designed to enhance the models’ visual inference capabilities and contextual comprehension from images. Our findings reveal that GPT-4 Vision demonstrated exceptional proficiency in interpreting visual data, providing accurate assessments of flood severity and water quality. Additionally, MLLMs showed potential in various hydrological applications, including drought prediction, streamflow forecasting, groundwater management, and wetland conservation. These models can optimize water resource management by predicting rainfall, evaporation rates, and soil moisture levels, thereby promoting sustainable agricultural practices. This research provides valuable insights into the potential applications of advanced AI models in addressing complex hydrological challenges and improving real-time decision-making in water resource management Full article
Show Figures

Figure 1

Figure 1
<p>The workflow for MLLM benchmarking in hydrological tasks.</p>
Full article ">
18 pages, 6180 KiB  
Article
Field Study and Numerical Modeling to Assess the Impact of On-Site Septic Systems on Groundwater Quality of Jeju Island, South Korea
by Mijin Kim, Eun-Hee Koh and Jinkeun Kim
Hydrology 2024, 11(9), 146; https://doi.org/10.3390/hydrology11090146 - 10 Sep 2024
Viewed by 329
Abstract
Septic-derived nitrogen (N) sources have harmful effects on water resources, humans, and ecosystems in several countries. On Jeju Island, South Korea, the rapid increase in personal sewage treatment facilities (PSTFs, also known as on-site septic systems) raises concerns regarding the deterioration of groundwater [...] Read more.
Septic-derived nitrogen (N) sources have harmful effects on water resources, humans, and ecosystems in several countries. On Jeju Island, South Korea, the rapid increase in personal sewage treatment facilities (PSTFs, also known as on-site septic systems) raises concerns regarding the deterioration of groundwater quality, as groundwater is the sole water resource on the island. Therefore, this study employed a field study and numerical modeling to assess the impact of PSTF effluents on groundwater quality in the Jocheon area of northeastern Jeju. Water quality analysis revealed that the total nitrogen (T-N) concentrations in the effluent exceeded the effluent standards (75–92% PSTFs). The numerical model simulated the transport of N species, showing limited NH4+ and NO2 plume migration near the surface due to nitrification and adsorption. However, NO3 concentrations increased and stabilized over time, leaching on the water table with higher levels in lowland areas and clustered PSTFs. The predictive model estimated a 79% reduction in NO3 leaching when the effluents followed standards, indicating the necessity of effective PSTF management. This study highlights the importance of managing improperly operated septic systems to mitigate groundwater contamination based on an understanding of the behavior of N species in subsurface hydrologic systems. Full article
Show Figures

Figure 1

Figure 1
<p>Map of (<b>a</b>) Jeju Island and (<b>b</b>) the Jocheon study area in the northeastern Jeju Island, South Korea showing PSTFs (personal sewage treatment facilities), agricultural land, and sampling locations (PSTFs and groundwater wells), and (<b>c</b>) a geological cross-section along line A−A′ in (<b>b</b>).</p>
Full article ">Figure 2
<p>(<b>a</b>) A three-dimensional model domain, (<b>b</b>) cross−sectional geologic layers (B−B′ in (<b>a</b>)) in the model (Z exaggeration: 5) and (<b>c</b>) boundary conditions (BCs) for numerical modeling.</p>
Full article ">Figure 3
<p>Relationships of T-N and NH<sub>4</sub><sup>+</sup> in PSTF effluents.</p>
Full article ">Figure 4
<p>Relationships of Cl<sup>−</sup> and NO<sub>3</sub><sup>−</sup> in groundwater. Gray dotted line indicates anthropogenic NO<sub>3</sub><sup>−</sup> contamination in groundwater suggested by Koh et al. [<a href="#B34-hydrology-11-00146" class="html-bibr">34</a>].</p>
Full article ">Figure 5
<p>Comparison between observed and simulated groundwater heads.</p>
Full article ">Figure 6
<p>Simulated (<b>a</b>) head and (<b>b</b>) saturation in steady-state condition (Z exaggeration: 5).</p>
Full article ">Figure 7
<p>Comparison between observed and simulated NO<sub>3</sub><sup>−</sup> concentrations in groundwater. N source loading (<b>a</b>) only from PSTFs (scenario 1), and (<b>b</b>) PSTFs and chemical fertilizer (scenario 2−2).</p>
Full article ">Figure 8
<p>Cross−sectional view showing spatiotemporal changes in simulated (<b>a</b>–<b>c</b>) NH<sub>4</sub><sup>+</sup>, (<b>d</b>–<b>f</b>) NO<sub>2</sub><sup>−</sup>, and (<b>g</b>–<b>i</b>) NO<sub>3</sub><sup>−</sup> concentration (unit: mg/L). Location of the cross−section is provided in <a href="#hydrology-11-00146-f002" class="html-fig">Figure 2</a>a.</p>
Full article ">Figure 9
<p>Simulated results of leaching NO<sub>3</sub><sup>−</sup> on the water table with different N source loading conditions. (<b>a</b>) PSTFs effluents in present condition (scenario 1), (<b>b</b>) chemical fertilizers only (no PSTFs considered; scenario 2−1), (<b>c</b>) PSTFs effluent and chemical fertilizer (scenario 2−2), and (<b>d</b>) PSTFs effluents following the T-N standards (scenario 3).</p>
Full article ">
28 pages, 13830 KiB  
Article
Integrated Geospatial and Geostatistical Multi-Criteria Evaluation of Urban Groundwater Quality Using Water Quality Indices
by Iram Naz, Hong Fan, Rana Waqar Aslam, Aqil Tariq, Abdul Quddoos, Asif Sajjad, Walid Soufan, Khalid F. Almutairi and Farhan Ali
Water 2024, 16(17), 2549; https://doi.org/10.3390/w16172549 - 9 Sep 2024
Viewed by 601
Abstract
Groundwater contamination poses a severe public health risk in Lahore, Pakistan’s second-largest city, where over-exploited aquifers are the primary municipal and domestic water supply source. This study presents the first comprehensive district-wide assessment of groundwater quality across Lahore using an innovative integrated approach [...] Read more.
Groundwater contamination poses a severe public health risk in Lahore, Pakistan’s second-largest city, where over-exploited aquifers are the primary municipal and domestic water supply source. This study presents the first comprehensive district-wide assessment of groundwater quality across Lahore using an innovative integrated approach combining geographic information systems (GIS), multi-criteria decision analysis (MCDA), and water quality indexing techniques. The core objectives were to map the spatial distributions of critical pollutants like arsenic, model their impacts on overall potability, and evaluate targeted remediation scenarios. The analytic hierarchy process (AHP) methodology was applied to derive weights for the relative importance of diverse water quality parameters based on expert judgments. Arsenic received the highest priority weight (0.28), followed by total dissolved solids (0.22) and hardness (0.15), reflecting their significance as health hazards. Weighted overlay analysis in GIS delineated localized quality hotspots, unveiling severely degraded areas with very poor index values (>150) in urban industrial zones like Lahore Cantt, Model Town, and parts of Lahore City. This corroborates reports of unregulated industrial effluent discharges contributing to aquifer pollution. Prospective improvement scenarios projected that reducing heavy metals like arsenic by 30% could enhance quality indices by up to 20.71% in critically degraded localities like Shalimar. Simulating advanced multi-barrier water treatment processes showcased an over 95% potential reduction in arsenic levels, indicating the requirement for deploying advanced oxidation and filtration infrastructure aligned with local contaminant profiles. The integrated decision support tool enables the visualization of complex contamination patterns, evaluation of remediation options, and prioritizing risk-mitigation investments based on the spatial distribution of hazard exposures. This framework equips urban planners and utilities with critical insights for developing targeted groundwater quality restoration policies through strategic interventions encompassing treatment facilities, drainage infrastructure improvements, and pollutant discharge regulations. Its replicability across other regions allows for tackling widespread groundwater contamination challenges through robust data synthesis and quantitative scenario modeling capabilities. Full article
(This article belongs to the Special Issue Groundwater Quality and Human Health Risk, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Map of the study area: Lahore District.</p>
Full article ">Figure 2
<p>Flow chart of methodology.</p>
Full article ">Figure 3
<p>Parameters results: (<b>a</b>) pH, (<b>b</b>) turbidity, (<b>c</b>) TDS, (<b>d</b>) EC, and (<b>e</b>) hardness.</p>
Full article ">Figure 4
<p>Parameter results: (<b>a</b>) calcium, (<b>b</b>) magnesium, (<b>c</b>) alkalinity, (<b>d</b>) chloride, and (<b>e</b>) arsenic.</p>
Full article ">Figure 5
<p>Groundwater quality AHP index.</p>
Full article ">Figure 6
<p>Scenario I (variation in heavy metals): (<b>a</b>) 10%, (<b>b</b>) 20%, and (<b>c</b>) 30% reduction.</p>
Full article ">Figure 7
<p>Scenario II (variation in chemical parameters): (<b>a</b>) Chemical 10%, (<b>b</b>) Chemical 20%, (<b>c</b>) Chemical 30%.</p>
Full article ">Figure 8
<p>Variation in primary treatment. (<b>a</b>) Part A with a 40% reduction in the total dissolved solids (TDSs) and (<b>b</b>) Part B with a 60% reduction in the total dissolved solids (TDSs).</p>
Full article ">Figure 9
<p>Variation in secondary treatment. (<b>a</b>) Part A with a 65% reduction in the total dissolved solids (TDSs) and 45% reduction in hardness and (<b>b</b>) Part B with an 80% reduction in the total dissolved solids (TDSs) and 55% reduction in hardness.</p>
Full article ">Figure 10
<p>Variation in tertiary treatment. (<b>a</b>) Part A with 90% reduction in arsenic and (<b>b</b>) Part B with 95% reduction in arsenic.</p>
Full article ">
23 pages, 4665 KiB  
Article
Natural Water Sources and Small-Scale Non-Artisanal Andesite Mining: Scenario Analysis of Post-Mining Land Interventions Using System Dynamics
by Mohamad Khusaini, Rita Parmawati, Corinthias P. M. Sianipar, Gatot Ciptadi and Satoshi Hoshino
Water 2024, 16(17), 2536; https://doi.org/10.3390/w16172536 - 7 Sep 2024
Viewed by 410
Abstract
Small-scale open-pit, non-artisanal mining of low-value ores is an understudied practice despite its widespread occurrence and potential impact on freshwater resources due to mining-induced land-use/cover changes (LUCCs). This research investigates the long-term impacts of andesite mining in Pasuruan, Indonesia, on the Umbulan Spring’s [...] Read more.
Small-scale open-pit, non-artisanal mining of low-value ores is an understudied practice despite its widespread occurrence and potential impact on freshwater resources due to mining-induced land-use/cover changes (LUCCs). This research investigates the long-term impacts of andesite mining in Pasuruan, Indonesia, on the Umbulan Spring’s water discharge within its watershed. System Dynamics (SD) modeling captures the systemic and systematic impact of mining-induced LUCCs on discharge volumes and groundwater recharge. Agricultural and reservoir-based land reclamation scenarios then reveal post-mining temporal dynamics. The no-mining scenario sees the spring’s discharge consistently decrease until an inflection point in 2032. With mining expansion, reductions accelerate by ~1.44 million tons over two decades, or 65.31 thousand tons annually. LUCCs also decrease groundwater recharge by ~2.48 million tons via increased surface runoff. Proposed post-mining land interventions over reclaimed mining areas influence water volumes differently. Reservoirs on reclaimed land lead to ~822.14 million extra tons of discharge, 2.75 times higher than the agricultural scenario. Moreover, reservoirs can restore original recharge levels by 2039, while agriculture only reduces the mining impact by 28.64% on average. These findings reveal that small-scale non-artisanal andesite mining can disrupt regional hydrology despite modest operating scales. Thus, evidence-based guidelines are needed for permitting such mines based on environmental risk and site water budgets. Policy options include discharge or aquifer recharge caps tailored to small-scale andesite mines. The varied outputs of rehabilitation scenarios also highlight evaluating combined land and water management interventions. With agriculture alone proving insufficient, optimized mixes of revegetation and water harvesting require further exploration. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>Systems scheme of the impact of small-scale mining on natural water discharge.</p>
Full article ">Figure 2
<p>Research design.</p>
Full article ">Figure 3
<p>Location of the mining area (<b>a</b>) in Pasuruan (<b>b</b>), East Java (<b>c</b>), Indonesia (<b>d</b>).</p>
Full article ">Figure 4
<p>Original conditions of the mining area.</p>
Full article ">Figure 5
<p>The Causal-Loop Diagram (CLD).</p>
Full article ">Figure 6
<p>The Stock-and-Flow Diagram (SFD).</p>
Full article ">Figure 7
<p>Projected water discharge of the Umbulan Spring without mining operations.</p>
Full article ">
31 pages, 29333 KiB  
Article
VARS and HDMR Sensitivity Analysis of Groundwater Flow Modeling through an Alluvial Aquifer Subject to Tidal Effects
by Javier Samper, Brais Sobral, Bruno Pisani, Alba Mon, Carlos López-Vázquez and Javier Samper-Pilar
Water 2024, 16(17), 2526; https://doi.org/10.3390/w16172526 - 5 Sep 2024
Viewed by 487
Abstract
Groundwater flow and transport models are essential tools for assessing and quantifying the migration of organic contaminants at polluted sites. Uncertainties in the hydrodynamic and transport parameters of the aquifer have a significant effect on model predictions. Uncertainties can be quantified with advanced [...] Read more.
Groundwater flow and transport models are essential tools for assessing and quantifying the migration of organic contaminants at polluted sites. Uncertainties in the hydrodynamic and transport parameters of the aquifer have a significant effect on model predictions. Uncertainties can be quantified with advanced sensitivity methods such as Sobol’s High Dimensional Model Reduction (HDMR) and Variogram Analysis of Response Surfaces (VARS). Here we present the application of VARS and HDMR to assess the global sensitivities of the outputs of a transient groundwater flow model of the Gállego alluvial aquifer which is located downstream of the Sardas landfill in Huesca (Spain). The aquifer is subject to the tidal effects caused by the daily oscillations of the water level in the Sabiñánigo reservoir. Global sensitivities are analyzed for hydraulic heads, aquifer/reservoir fluxes, groundwater Darcy velocity, and hydraulic head calibration metrics. Input parameters include aquifer hydraulic conductivities and specific storage, aquitard vertical hydraulic conductivities, and boundary inflows and conductances. VARS, HDMR, and graphical methods agree to identify the most influential parameters, which for most of the outputs are the hydraulic conductivities of the zones closest to the landfill, the vertical hydraulic conductivity of the most permeable zones of the aquitard, and the boundary inflow coming from the landfill. The sensitivity of heads and aquifer/reservoir fluxes with respect to specific storage change with time. The aquifer/reservoir flux when the reservoir level is high shows interactions between specific storage and aquitard conductivity. VARS and HDMR parameter rankings are similar for the most influential parameters. However, there are discrepancies for the less relevant parameters. The efficiency of VARS was demonstrated by achieving stable results with a relatively small number of simulations. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>Flowchart of the methodology used in this study.</p>
Full article ">Figure 2
<p>(<b>a</b>) Location of the study area; (<b>b</b>) enlargement showing the model domain, the Sabiñánigo reservoir, the Sardas landfill, the Gállego River course, and the INQUINOSA (Sabiñánigo, Spain) former production site. The arrows along the Gállego River course indicate the flow direction.</p>
Full article ">Figure 3
<p>Cross-sectional geological profile of the Sabiñánigo reservoir and the Gállego River alluvial plain as reported by Sobral et al. [<a href="#B38-water-16-02526" class="html-bibr">38</a>]. Alluvial deposits include a shallow silt layer (green) and a deep layer of sand and gravel.</p>
Full article ">Figure 4
<p>2D finite element mesh, monitoring wells, material zones, boundary conditions, and GSA input parameters (<b>top plot</b>) and enlargement showing the area downstream of the Sardas landfill (<b>bottom plot</b>). The confined storage coefficient (S<sub>S</sub>) is the same in the four material zones. The sands and gravels are assumed to be confined in the alluvial (r<sub>c</sub>), except in the wooded areas (r<sub>u</sub>). Unconfined areas are shown with a back-hashed polygon.</p>
Full article ">Figure 5
<p>Map showing the reservoir tail area (hashed blue polygon) where aquifer/reservoir fluxes were calculated at times t1, t2, and t3, the monitoring wells whose piezometric data were used to calculate the calibration metrics, monitoring wells ST1C, PS19B, SPN1, and PS16C (where the average Darcy velocity is computed).</p>
Full article ">Figure 6
<p>Measured reservoir hydrograph and piezometric heads in well ST1C from 18–20 September 2020. The computed piezometric heads in monitoring wells ST1C, PS19B, and SPN1 and the aquifer/reservoir fluxes are analyzed at the following times: (1) t1, 18 September 2020, 20:00 (low reservoir water level), (2) t2, 18 September 2020, 22:30 (peak reservoir water level) and (3) t3, 19 September 2020, 04:30 (descending reservoir water level).</p>
Full article ">Figure 7
<p>Scatterplots of the computed piezometric heads in wells ST1Ct2 (<b>upper left plot</b>), PS19Bt2 (<b>upper right plot</b>), SPN1t2 (<b>lower left plot</b>), and Qt2 (<b>lower right plot</b>) versus the vertical hydraulic conductivity of the silting sediments in the former river course (Kvs1). The sample of 16384 points was generated with a Sobol sequence. The clouds of plots are shown for the following three ranges of percentiles, p, of the specific storage coefficient (S<sub>S</sub>): (1) p &lt; 30%; (2) 30% &lt; p &lt; 70%, and (3) p &gt; 70%.</p>
Full article ">Figure 8
<p>CUSUNORO curves of computed head in wells ST1C and PS19B at times t1, t2 and t3; and well SPN1 at times t1 and t2.</p>
Full article ">Figure 9
<p>CUSUNORO curves of the computed head in well SPN1 at time t3, MAEg, NRMSEg, NSEg, Q<sub>t1</sub>, Q<sub>t2</sub>, Q<sub>t3</sub>, and q<sub>av</sub>.</p>
Full article ">Figure 10
<p>IVARS<sub>50</sub> indexes of input parameters as a function of the number of star centers for MAEg (<b>upper plot</b>), and robustness of ranking as a function of the number of star centers (<b>bottom plot</b>).</p>
Full article ">Figure 11
<p>IVARS<sub>50</sub> indexes of input parameters as a function of the number of star centers for the average Darcy velocity (q<sub>av</sub>) (<b>upper plot</b>), and robustness of ranking as a function of the number of star centers (<b>bottom plot</b>).</p>
Full article ">Figure 12
<p>Sample variograms of the computed heads in monitoring wells ST1C and PS19B at times t1, t2, and t3 and monitoring well SPN1 at times t1 and t2. Only the variograms of the five most influential parameters are shown in the plots.</p>
Full article ">Figure 13
<p>Sample variograms of the computed head in well SPN1 at time t3, MAEg, NRMSEg, NSEg, Q<sub>t1</sub>, Q<sub>t2</sub>, Q<sub>t3</sub>, and q<sub>av</sub>. Only the variograms of the five most influential parameters are shown in the plots.</p>
Full article ">Figure 14
<p>VARS-TO, IVARS<sub>50</sub>, and VARS-ABE indexes for the computed heads in wells ST1C and PS19B at times t1, t2, and t3 and well SPN1 at times t1 and t2.</p>
Full article ">Figure 15
<p>VARS-TO, IVARS<sub>50</sub>, and VARS-ABE indexes for the computed head in well SPN1 at time t3, MAEg, NRMSEg, NSEg, Q<sub>t1</sub>, Q<sub>t2</sub>, Q<sub>t3</sub>, and q<sub>av</sub>.</p>
Full article ">Figure 16
<p>IVARS<sub>50</sub> sensitivity indexes for computed heads in wells ST1C (<b>top left plot</b>), PS19B (<b>top right plot</b>), and SPN1 (<b>bottom left plot</b>) and aquifer/reservoir flow (<b>bottom right plot</b>) at times t1, t2, and t3.</p>
Full article ">Figure 17
<p>IVARS<sub>50</sub> sensitivity indexes for calibration metrics MAEg, NRMSEg, and NSEg.</p>
Full article ">
23 pages, 16575 KiB  
Article
Remote Sensing of Floodwater-Induced Subsurface Halite Dissolution in a Salt Karst System, with Implications for Landscape Evolution: The Western Shores of the Dead Sea
by Gidon Baer, Ittai Gavrieli, Iyad Swaed and Ran N. Nof
Remote Sens. 2024, 16(17), 3294; https://doi.org/10.3390/rs16173294 - 4 Sep 2024
Viewed by 637
Abstract
We study the interrelations between salt karst and landscape evolution at the Ze’elim and Hever alluvial fans, Dead Sea (DS), Israel, in an attempt to characterize the ongoing surface and subsurface processes and identify future trends. Using light detection and ranging, interferometric synthetic [...] Read more.
We study the interrelations between salt karst and landscape evolution at the Ze’elim and Hever alluvial fans, Dead Sea (DS), Israel, in an attempt to characterize the ongoing surface and subsurface processes and identify future trends. Using light detection and ranging, interferometric synthetic aperture radar, drone photography, time-lapse cameras, and direct measurements of floodwater levels, we document floodwater recharge through riverbed sinkholes, subsurface salt dissolution, groundwater flow, and brine discharge at shoreline sinkholes during the years 2011–2023. At the Ze’elim fan, most of the surface floodwater drains into streambed sinkholes and discharges at shoreline sinkholes, whereas at the Hever fan, only a small fraction of the floodwater drains into sinkholes, while the majority flows downstream to the DS. This difference is attributed to the low-gradient stream profiles in Ze’elim, which enable water accumulation and recharge in sinkholes and their surrounding depressions, in contrast with the higher-gradient Hever profiles, which yield high-energy floods capable of carrying coarse gravel that eventually fill the sinkholes. The rapid drainage of floodwater into sinkholes also involves slope failure due to pore-pressure drop and cohesion loss within hours after each drainage event. Surface subsidence lineaments detected by InSAR indicate the presence of subsurface dissolution channels between recharge and discharge sites in the two fans and in the nearby Lynch straits. Subsidence and streambed sinkholes occur in most other fans and streams that flow to the DS; however, with the exception of Ze’elim, all other streams show only minor or no recharge along their course. This is due to either the high-gradient profiles, the gravelly sediments, the limited floods, or the lack of conditions for sinkhole development in the other streambeds. Thus, understanding the factors that govern the flood-related karst formation is of great importance for predicting landscape evolution in the DS region and elsewhere and for sinkhole hazard assessment. Full article
(This article belongs to the Special Issue Remote Sensing of the Dead Sea Region)
Show Figures

Figure 1

Figure 1
<p>LiDAR topography of the two study areas draped upon hill−shaded DSMs. (<b>a</b>) Ze’elim fan. Gully numbers (in white) are after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]. (<b>b</b>) Hever fan. (<b>c</b>) Location maps of the study areas.</p>
Full article ">Figure 2
<p>Elevation profiles along Ze’elim and Hever riverbeds, May 2020. Note the low gradients (1−3%) and fine-grained composition of the Ze’elim riverbeds (blue-green profiles, for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>a), in contrast with the high gradients (3−4.5%) and coarse gravel sediments of the Hever riverbeds (red-brown profiles; for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>b), which decrease to 1−1.5% only at their easternmost parts.</p>
Full article ">Figure 3
<p>(<b>a</b>) Time-lapse camera and drone, overlooking gully 14 recharge sinkhole. The blue arrow marks the flow direction from west to east. (<b>b</b>) View south at the DSW canal as floodwater crosses the overpasses. Photo courtesy of DSW. (<b>c</b>) Locations of hydrometers (marked by white arrows) that are installed at an overpass. E and W mark eastern and western hydrometers.</p>
Full article ">Figure 4
<p>Photos of recharge sinkholes at Ze’elim fan streambeds. Blue arrows mark the flow direction. Ab—abandoned gullies, overhanging downstream of the recharge sinkholes. For location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>a. (<b>a</b>) Gully 6. (<b>b</b>) Gully 7. (<b>c</b>) Gully 13. (<b>d</b>) Gully 14. Drone picture was taken by Liran Ben Moshe.</p>
Full article ">Figure 5
<p>(<b>a</b>) Floodwater recharge (red circles) and discharge (blue rectangles) sites at the Ze’elim fan. Red and yellow triangles mark locations and operation intervals of the TLCs. Yellow numbers denote gully numbers (after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]). (<b>b</b>) Drone photograph, 2 January 2020, showing the discharge sinkholes (rectangles) and TLCs (triangles) within and around the shoreline sinkholes of gully 10. Note that not all TLCs operated simultaneously.</p>
Full article ">Figure 6
<p>Discharge sinkhole 10a (see location in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b). (<b>a</b>) View east, February 2024. (<b>b</b>) TLC picture showing water discharge following the 25 March 2019 flood. (<b>c</b>) The nested sinkhole at the northern wall of the major sinkhole, February 2024, exposing the “Sinkhole Salt” layer (white layers with small cavities), the dissolution channel openings, and groundwater flow.</p>
Full article ">Figure 7
<p>Sinkholes and subsidence along the course of gully 3. (<b>a</b>) LiDAR DSM, July 2023 (see location in <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>). The location of profile A−A’ (panel <b>d</b>) is shown in a dashed white line. The area around the recharge sinkhole is marked by a circle. (<b>b</b>) Drone photograph of the recharge area (sinkhole marked by white circle), January 2024, taken by Liran Ben Moshe. (<b>c</b>) Streambed sinkhole recharging floodwater after the 15.2.2024 flood. (<b>d</b>) Elevation profile A−A’ along the gully, September 2023 (location shown in panel <b>a</b>). Note that this recharge sinkhole does not appear in 2020 (see Ze’elim 3 profile in <a href="#remotesensing-16-03294-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 8
<p>Water levels at five overpasses during the 21–22 November 2021 flood in Ze’elim. See inset for location. The overpasses are marked by white numbers, and streams are marked by yellow numbers. E and W stand for eastern and western hydrographs, respectively (<a href="#remotesensing-16-03294-f003" class="html-fig">Figure 3</a>c).</p>
Full article ">Figure 9
<p>An interferogram of the Ze’elim fan spanning 44 days in early 2024, showing subsidence lineaments that are interpreted as surface manifestations of subsurface dissolution channels. The two acquisition times are 13 January 2024 and 26 February 2024. WZSL, EZSL, and NZSL stand for western, eastern, and northern Ze’elim subsidence lineaments, shown by white, orange, and yellow arrows, respectively. Gully numbers are marked in white (after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>]).</p>
Full article ">Figure 10
<p>Surface (dashed white lines) and proposed subsurface water pathways (dashed yellow lines) in Ze’elim: a southern pathway from gully 14 to the western side of sinkhole 10 (10a in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b), and central pathways from gullies 5, 6, and 7 to the eastern side of sinkhole 10 (10f in <a href="#remotesensing-16-03294-f005" class="html-fig">Figure 5</a>b). Gully numbers are after [<a href="#B3-remotesensing-16-03294" class="html-bibr">3</a>].</p>
Full article ">Figure 11
<p>(<b>a</b>,<b>b</b>) Annual surface elevation changes in the Hever fan, draped upon LiDAR DSMs. River incision, riverbank collapse, subsidence, and sinkholes are displayed by negative (blue) values. Aggradation of alluvial material along the streambeds and within sinkholes is displayed by positive (red) values and by white arrows. White ellipses mark subsidence around sinkhole clusters. The black arrow in (<b>a</b>) points at a meandering subsidence lineament, interpreted as the surface manifestation of a subsurface dissolution channel. (<b>c</b>) Interferogram showing sinkhole-related subsidence (semi-circular fringe colors) and a meandering subsidence lineament (marked by white arrows) that is interpreted as the surface manifestation of a subsurface dissolution channel between the western cluster of recharging sinkholes and the eastern subsidence zone (similar to the lineament in panel <b>a</b>). The acquisition times of the two images are 7 June 2018 and 18 June 2018.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) TLC photos showing recharge of floodwater at sinkholes in the northern branch of Hever fan during the 20 February 2015 flood. Blue arrows mark the braided streambed flow direction. Note the water overflow and the filling of the sinkhole with gravel at the final hours of the flood (panels (<b>b</b>) and (<b>c</b>), respectively). (<b>d</b>) Drone picture of 7 February 2019 floodwater drained into recharge sinkholes along the northern Hever branch with overflowing water continuing downstream.</p>
Full article ">Figure 13
<p>(<b>a</b>) Discharge sinkholes at the lower part of the southern Hever branch. (<b>b</b>) Offshore discharge sites at the Hever shoreline (white arrows). (<b>c</b>) A small salt chimney 2 m offshore in Hever. (<b>d</b>) <span class="html-italic">Anabasis setifera</span> vegetation at the lower Hever southern streambed. For location, see the black arrow in <a href="#remotesensing-16-03294-f011" class="html-fig">Figure 11</a>a.</p>
Full article ">Figure 14
<p>Linear and meandering subsidence patterns at the Lynch straits (for location, see <a href="#remotesensing-16-03294-f001" class="html-fig">Figure 1</a>c). (<b>a</b>) Subtraction map of LiDAR DSMs between 2023 and 2014. (<b>b</b>) Interferogram between 2 and 13 April 2018. The white arrows point to subsidence lineaments that are interpreted to form above subsurface salt dissolution channels.</p>
Full article ">Figure 15
<p>Stream gradients along the western shoreline of the DS. Color legend distinguishes between low-gradient, mud-dominated streams (green); high-gradient, gravel-dominated streams (red); and intermediate-gradient mixed mud-gravel streams (yellow).</p>
Full article ">
11 pages, 1221 KiB  
Article
The Seasonal Characterization and Temporal Evolution of Nitrogen, Phosphorus and Potassium in the Surface and Groundwater of an Agricultural Hydrographic Basin in the Midwestern Brazilian Savanna
by Nayara Luiz Pires, Daphne Heloisa de Freitas Muniz, Luane Souza de Araújo, Jorge Enoch Furquim Werneck Lima, Roberto Arnaldo Trancoso Gomes, Eloisa Dutra Caldas and Eduardo Cyrino Oliveira-Filho
Sustainability 2024, 16(17), 7659; https://doi.org/10.3390/su16177659 - 3 Sep 2024
Viewed by 531
Abstract
The Brazilian savanna (Cerrado Biome) is one of the most important regions in the world in terms of food production, with the use of fertilizers based on nitrogen, phosphorus and potassium (NPK). When not applied properly, fertilizers can alter and affect water [...] Read more.
The Brazilian savanna (Cerrado Biome) is one of the most important regions in the world in terms of food production, with the use of fertilizers based on nitrogen, phosphorus and potassium (NPK). When not applied properly, fertilizers can alter and affect water quality. The objective of this study was to evaluate the presence of these compounds in surface and groundwater in the Upper Jardim River Hydrographic Unit, Federal District, thus characterizing seasonal variations during the dry and rainy seasons in two periods. A total of 207 groundwater samples and 23 surface water samples were collected in the years 2014, 2015, 2019 and 2020. The parameters analyzed were pH and nitrate, nitrite, ammonium, phosphate and potassium ions. In groundwater samples, pH values were significantly higher and ion levels lower in samples collected during the early years (except for nitrate), and the ammonium concentrations were lower in the dry season than the rainy (in 2014 and 2019). In surface samples, total phosphorus levels were significantly higher in the rainy/2019 compared to the rainy/2020 season, while this tendency was inverted for potassium during the dry season. The use of NPK-based fertilizers has increased considerably in recent years in the region due to the expansion of the agricultural area, and although the results of the study show that concentrations in water are much lower than the maximum values allowed by Brazilian legislation, continuous monitoring is necessary to guarantee water quality. Full article
(This article belongs to the Special Issue Lakes and Rivers Ecological Protection and Water Quality)
Show Figures

Figure 1

Figure 1
<p>(<b>A</b>) Brazil and Brazilian Federal District in South America. (<b>B</b>) Location of the Upper Jardim River Hydrographic Unit (HU-35) in the Brazilian Federal District. (<b>C</b>) Water sampling points. Prepared using MAPBIOMAS [<a href="#B23-sustainability-16-07659" class="html-bibr">23</a>] and SIEG [<a href="#B24-sustainability-16-07659" class="html-bibr">24</a>].</p>
Full article ">Figure 2
<p>pH and ion concentration ranges in groundwater samples collected from the Upper Jardim River Hydrographic Unit (HU-35) during the dry and rainy seasons of 2014, 2015, 2019 and 2020. Kruskal–Wallis test (non-parametric, n = 25–27); * &lt; 0.05; ** &lt; 0.01; *** &lt; 0.001; **** &lt; 0.0001.</p>
Full article ">Figure 3
<p>pH and ion concentrations in surface samples collected from the Upper Jardim River Hydrographic Unit (HU-35) during the dry and rainy seasons of 2019/2020. Kruskal–Wallis test (non-parametric, n = 5–6); * &lt; 0.05.</p>
Full article ">
22 pages, 9280 KiB  
Article
Concentrations of F, Na+, and K+ in Groundwater before and after an Earthquake: A Case Study on Tenerife Island, Spain
by Eduardo de Miguel-García and José Francisco Gómez-González
Hydrology 2024, 11(9), 138; https://doi.org/10.3390/hydrology11090138 - 3 Sep 2024
Viewed by 425
Abstract
Freshwater, vital for life and ecosystems, accounts for only 2.5% of Earth’s water, and is primarily located in polar caps, underground reservoirs, and surface water. Its quality varies due to environmental interactions, especially in groundwater. Tenerife, located in the Canary Islands, Spain, relies [...] Read more.
Freshwater, vital for life and ecosystems, accounts for only 2.5% of Earth’s water, and is primarily located in polar caps, underground reservoirs, and surface water. Its quality varies due to environmental interactions, especially in groundwater. Tenerife, located in the Canary Islands, Spain, relies mainly on underground aquifers and tunnels capturing 51.6 cubic hectometers annually. Ensuring safe drinking water is a global challenge due to health risks from poor water quality, including diseases and cancer. Fluoride, sodium, and potassium are essential for health, and are mainly derived from groundwater as fluoride ions (F) and sodium and potassium cations (Na+, K+). However, excessive F, Na+, and K+ in drinking water is harmful. The World Health Organization limits F to 1.5 mg/L, Na+ to 8.70 meq/L, and K+ to 0.31 meq/L. Geological, climatic, and human factors control the presence and transport of F, Na+, and K+ in groundwater. Seismic events can impact water quality, with long-term effects linked to aquifer structure and transient effects from gas and fluid expansion during earthquakes. This study was motivated by a 3.8 mbLg earthquake in Tenerife in 2012, which allowed its impact on groundwater quality, specifically F, Na+, and K concentrations, to be examined. Post-earthquake, F levels alarmingly increased to 8.367 meq/L, while Na+ and K+ showed no significant changes. This research quantifies the influence of earthquakes on increasing F levels and evaluates F reduction during low seismic activity, emphasizing the importance of water management on volcanic islands. Full article
(This article belongs to the Section Surface Waters and Groundwaters)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Geological units of Tenerife. (<b>b</b>) Water bodies of Tenerife Island. (<b>c</b>) Groundwater extraction zones. (<b>d</b>) Georeferenced information on the piezometric surface [<a href="#B20-hydrology-11-00138" class="html-bibr">20</a>].</p>
Full article ">Figure 2
<p>Cross-section of the island of Tenerife: (<b>a</b>) identification of saturated zone in dikes and (<b>b</b>) the tunnel to the aquifer.</p>
Full article ">Figure 3
<p>Tunnel entrance and water sampling point (<b>a</b>), and the end of the tunnel (<b>b</b>).</p>
Full article ">Figure 4
<p>Distribution of groundwater tunnels studied (red points) (tunnels referenced as Ref. 1 to Ref. 45 in <a href="#app1-hydrology-11-00138" class="html-app">Appendix A</a>, <a href="#hydrology-11-00138-t0A1" class="html-table">Table A1</a>, <a href="#hydrology-11-00138-t0A2" class="html-table">Table A2</a> and <a href="#hydrology-11-00138-t0A3" class="html-table">Table A3</a>) on the island of Tenerife, Canary Islands, Spain.</p>
Full article ">Figure 5
<p>Variations in F<sup>−</sup>, Na<sup>+</sup>, and K<sup>+</sup> in water samples taken in 45 tunnels, between the years 2011 and 2021 (each colorful curve represents data from the same tunnel.). Ion concentration (meq/L) vs. normalized value. The dashed vertical line represents the 18 August 2012 earthquake with a magnitude of 3.8 mbLg and an intensity of IV EMS.</p>
Full article ">Figure 6
<p>Precipitation patterns on the island of Tenerife: Geospatial distribution of 15 meteorological stations referenced in <a href="#hydrology-11-00138-t0A4" class="html-table">Table A4</a>, <a href="#hydrology-11-00138-t0A5" class="html-table">Table A5</a> and <a href="#hydrology-11-00138-t0A6" class="html-table">Table A6</a> across the island and the annual precipitation levels recorded at these stations.</p>
Full article ">Figure 7
<p>Normalized average precipitation and F<sup>−</sup> concentration: Black dots represent the F<sup>−</sup> concentration in 45 groundwater tunnels, while the black line illustrates the modeled F<sup>−</sup> concentration in water. The blue line denotes the average precipitation recorded at the meteorological stations.</p>
Full article ">Figure 8
<p>Map of European Macroseismic Scale and location of groundwater tunnels. Distribution of groundwater tunnels (red dots ) studied on the island of Tenerife, Canary Islands, Spain on the map of the European Macroseismic Scale (<a href="#hydrology-11-00138-t0A8" class="html-table">Table A8</a>), provided by the Spanish National Geographic Institute (IGN) on 18 August 2012, with a seism with a magnitude of 3.8 mbLg. The star is where the epicenter of the seism was located.</p>
Full article ">Figure 9
<p>Earthquake distribution and fluorine evolution. (<b>a</b>) Earthquake distribution since 18 August 2012 with intensities between II and IV (EMS). The red dot represents the location of the 18 August 2012 earthquake, which is the only one with a maximum intensity of IV. (<b>b</b>) shows the fit of a linear decreasing model (red line) with fluorine over time, starting from the maximum value reached after the earthquake of 18 August 2012 (blue dots). The green marks are the samples not used for the estimation of the linear decreasing model. The pink triangles represent the earthquake intensity located in (<b>a</b>).</p>
Full article ">Figure 10
<p>Spatial estimation of fluoride concentration variations over time (before, after, 10 years later, and 20 years later). The yellow dot is the location of the 18 August 2012 earthquake.</p>
Full article ">
22 pages, 8611 KiB  
Article
GIS-Based Analytical Hierarchy Process for Identifying Groundwater Potential Zones in Punjab, Pakistan
by Maira Naeem, Hafiz Umar Farid, Muhammad Arbaz Madni, Raffaele Albano, Muhammad Azhar Inam, Muhammad Shoaib, Muhammad Shoaib, Tehmena Rashid, Aqsa Dilshad and Akhlaq Ahmad
ISPRS Int. J. Geo-Inf. 2024, 13(9), 317; https://doi.org/10.3390/ijgi13090317 - 3 Sep 2024
Viewed by 508
Abstract
The quality and level of groundwater tables have rapidly declined because of intensive pumping in Punjab (Pakistan). For sustainable groundwater supplies, there is a need for better management practices. So, the identification of potential groundwater recharge zones is crucial for developing effective management [...] Read more.
The quality and level of groundwater tables have rapidly declined because of intensive pumping in Punjab (Pakistan). For sustainable groundwater supplies, there is a need for better management practices. So, the identification of potential groundwater recharge zones is crucial for developing effective management systems. The current research is based on integrating seven contributing factors, including geology, soil map, land cover/land use, lineament density, drainage density, slope, and rainfall to categorize the area into various groundwater recharge potential zones using remote sensing, geographic information system (GIS), and analytical hierarchical process (AHP) for Punjab, Pakistan. The weights (for various thematic layers) and rating values (for sub-classes) in the overlay analysis were assigned for thematic layers and then modified and normalized using the AHP. The result indicates that about 17.88% of the area falls under the category of very high groundwater potential zones (GWPZs). It was found that only 12.27% of the area falls under the category of very low GWPZs. The results showed that spatial technologies like remote sensing and geographic information system (GIS), when combined with AHP technique, provide a robust platform for studying GWPZs. This will help the public and government sectors to understand the potential zone for sustainable groundwater management. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study area.</p>
Full article ">Figure 2
<p>Flow diagram of the detailed methodology (dotted line differentiate the methodological steps).</p>
Full article ">Figure 3
<p>Relationship between the most influential parameters of GW potential zone.</p>
Full article ">Figure 4
<p>Study area maps of (<b>a</b>) elevation; (<b>b</b>) lineament density; (<b>c</b>) drainage density; (<b>d</b>) geology; (<b>e</b>) soil type; (<b>f</b>) LULC type; (<b>g</b>) rainfall; and (<b>h</b>) slope.</p>
Full article ">Figure 5
<p>Percentage change in land use/land cover (LULC) in the study region.</p>
Full article ">Figure 6
<p>Percentage weighted of each category selected for preparing thematic maps of groundwater recharge zones.</p>
Full article ">Figure 7
<p>Groundwater potential zones (GWPZs) map.</p>
Full article ">Figure 8
<p>Potential groundwater recharge zones in percentage area and actual area in km<sup>2</sup> of Punjab province.</p>
Full article ">
18 pages, 7517 KiB  
Article
Springs of the Arabian Desert: Hydrogeology and Hydrochemistry of Abu Jir Springs, Central Iraq
by John A. Webb, Jaafar Jotheri and Rod J. Fensham
Water 2024, 16(17), 2491; https://doi.org/10.3390/w16172491 - 2 Sep 2024
Viewed by 715
Abstract
The Arabian Desert is characterised by very low rainfall and high evaporation, yet over 210 springs are on its northeastern edge in central Iraq along the Abu Jir lineament, which represents the western depositional margin of a foreland basin infilled by the floodplain [...] Read more.
The Arabian Desert is characterised by very low rainfall and high evaporation, yet over 210 springs are on its northeastern edge in central Iraq along the Abu Jir lineament, which represents the western depositional margin of a foreland basin infilled by the floodplain sediments of the Tigris and Euphrates Rivers; there is little evidence of faulting. The springs discharge from gently east-dipping Paleocene–Eocene limestones, either where groundwater flowpaths intersect the ground surface or where groundwater flow is forced to the surface by confining aquitards. Calculated annual recharge to the aquifer system across the Arabian Desert plateau (130–500 million m3) is significant, largely due to rapid infiltration through karst dolines, such that karst porosity is the primary enabler of groundwater recharge. The recharge is enough to maintain flow at the Abu Jir springs, but active management of groundwater extraction for agriculture is required for their long-term sustainability. The hydrochemistry of the springs is determined by evaporation, rainfall composition (high SO4 concentrations are due to the dissolution of wind-blown gypsum in rainfall), and plant uptake of Ca and K (despite the sparse vegetation). Limestone dissolution has relatively little impact; many of the springs are undersaturated with respect to calcite and lack tufa/travertine deposits. The springs at Hit-Kubaysa contain tar and high levels of H2S that probably seeped upwards along subvertical faults from underlying oil reservoirs; this is the only location along the Abu Jir lineament where deep-seated faults penetrate to the surface. The presence of hydrocarbons reduces the Hit-Kubaysa spring water and converts the dissolved SO4 to H2S. Full article
Show Figures

Figure 1

Figure 1
<p>Location of springs (blue dots) along the Abu Jir lineament, and modern cities (squares). For location, see <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 2
<p>Topography of central Iraq showing the three major geomorphic sub-divisions, springs (white dots) located along the Abu Jir lineament, and the locations of the cross-sections in <a href="#water-16-02491-f005" class="html-fig">Figure 5</a>. For location, see <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>.</p>
Full article ">Figure 3
<p>Spring exposed in the bed of Sawa Lake when the lake receded (see <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> for location).</p>
Full article ">Figure 4
<p>Geology of the desert plateau west of Abu Jir lineament, after refs. [<a href="#B23-water-16-02491" class="html-bibr">23</a>,<a href="#B26-water-16-02491" class="html-bibr">26</a>,<a href="#B27-water-16-02491" class="html-bibr">27</a>], showing location of <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> and <a href="#water-16-02491-f002" class="html-fig">Figure 2</a> (box on main figure) and the stratigraphic profiles in <a href="#water-16-02491-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 5
<p>Geological cross-sections along the Abu Jir lineament (for locations, see <a href="#water-16-02491-f002" class="html-fig">Figure 2</a> and <a href="#water-16-02491-f004" class="html-fig">Figure 4</a>), showing the hydrogeology of the springs. Note the vertical exaggeration (x82); the actual westwards dip of the strata is &lt;1°. Stratigraphy derived from the outcrop distribution and bore logs on the following 1:250,000 geological maps: Karbala [<a href="#B33-water-16-02491" class="html-bibr">33</a>]; Al Najaf [<a href="#B34-water-16-02491" class="html-bibr">34</a>]; Baghdad [<a href="#B35-water-16-02491" class="html-bibr">35</a>]; Al Birreet [<a href="#B36-water-16-02491" class="html-bibr">36</a>]; Al Ramadi [<a href="#B37-water-16-02491" class="html-bibr">37</a>]; Shithatha [<a href="#B38-water-16-02491" class="html-bibr">38</a>].</p>
Full article ">Figure 6
<p>Piper plot of compositions of Abu Jir springs (data from) [<a href="#B8-water-16-02491" class="html-bibr">8</a>,<a href="#B10-water-16-02491" class="html-bibr">10</a>,<a href="#B11-water-16-02491" class="html-bibr">11</a>]; note that many available spring compositions, including the Shinafiyah springs, were incomplete as published and could not be plotted.</p>
Full article ">Figure 7
<p>Schoeller plot (standardised to Cl) comparing the median compositions of Abu Jir springs with desert plateau rainfall (see <a href="#water-16-02491-t001" class="html-table">Table 1</a> for data).</p>
Full article ">Figure 8
<p>Groundwater spring at Hit showing floating spongy bitumen (see <a href="#water-16-02491-f001" class="html-fig">Figure 1</a> for location).</p>
Full article ">Figure 9
<p>Stable isotope data for some Abu Jir springs; groundwater data from nearby wells shown for comparison; data from refs. [<a href="#B7-water-16-02491" class="html-bibr">7</a>,<a href="#B48-water-16-02491" class="html-bibr">48</a>].</p>
Full article ">
18 pages, 2730 KiB  
Article
Soil Characteristics and Response Mechanism of the Microbial Community in a Coal–Grain Compound Area with High Groundwater Levels
by Zhichao Chen, Jialiang Luo, Yiheng Jiao, Xiaoxuan Lyu, Shidong Wang and Hebing Zhang
Agronomy 2024, 14(9), 1993; https://doi.org/10.3390/agronomy14091993 - 2 Sep 2024
Viewed by 407
Abstract
Coal mining has led to escalating ecological and environmental issues in significant coal and grain production areas, posing a severe danger to food security. This study examines the disturbance patterns of soil factors and microbial communities in coal and grain production areas, and [...] Read more.
Coal mining has led to escalating ecological and environmental issues in significant coal and grain production areas, posing a severe danger to food security. This study examines the disturbance patterns of soil factors and microbial communities in coal and grain production areas, and attempts to understand the impact of subsidence and water accumulation stress on soil characteristics and microbial communities in coal mining subsidence areas with high subsidence levels. Five specific regions of Zhao Gu Yi Mine, situated in Henan Province and under the ownership of Jiaozuo Coal Group, were chosen. Aside from the control group (CK), the study blocks situated in the coal mining subsidence zones consisted of perennial subsidence ponding (PSP), seasonal subsidence ponding (SSP), the neutral zone (NZ), and the horizontal deformation zone (HDZ). The soil nutrient indices and the stoichiometric properties of soil C, N, and P were assessed on the surface of each block. The organization of the soil microbial community was identified using high-throughput sequencing. The findings indicate that: 1. Substantial disparities exist in soil properties and microbial community structure between the subsidence and non-subsidence zones. The levels of soil organic mater (SOM), total nitrogen (TN), total phosphorus (TP), available nitrogen (AN), and available phosphorus (AP) all decrease to different extents in the subsidence area. Additionally, the coal mining subsidence waterlogged area exhibits higher levels compared to the coal mining subsidence non-waterlogged area. Conversely, the soil water content (SWC), C/N ratio, C/P ratio, and N/P ratio all increase to varying degrees. 2. Regarding the composition of the community, the presence of Proteobacteria is considerably greater in the non-water-logged area of coal mining subsidence (NZ, HDZ) compared to the water-logged area and control group (p < 0.05). The prevalence of Firmicutes in the subsidence water area was substantially greater compared to both the subsidence non-waterlogged area and the control group (p < 0.05). The prevalence of Gemmatimonadota is markedly greater in the waterlogged area of mining subsidence compared to the non-waterlogged area and CK (p < 0.05). The Ascomycota population reached its highest value in the neutral zone (NZ), which was significantly greater than the values observed in the seasonal subsidence ponding (SSP) and perennial subsidence ponding (PSP) regions (p < 0.05). On the other hand, the Rozellomycota population had its highest value in the SSP region, which was significantly greater than the values observed in the other regions (p < 0.05). 3. The abundance and variety of soil bacteria and fungi, as well as their important populations, are associated with different levels of soil characteristics. The primary elements that influence the alteration of microbial communities are soil nutrients and soil water content. The presence of coal mine subsidence and water accumulation has a notable impact on the properties of the soil in the surrounding area. This study offers a scientific foundation for reclaiming land affected by subsidence caused by coal mining in regions where coal and grain production are the dominant industries. Full article
Show Figures

Figure 1

Figure 1
<p>Geographical location and sampling points distribution over the study area. Note: PSP: perennial subsidence ponding; SSP: seasonal subsidence ponding; HDZ: horizontal deformation zone; NZ: neutral zone; CK: CalvinKlein.</p>
Full article ">Figure 2
<p>Diversity of bacteria (<b>A</b>–<b>C</b>) and fungi (<b>D</b>–<b>F</b>) α diversity index in different locations of coal. Note: * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 3
<p>Non-metric Multidimensional Scaling (NMDS) analysis based on the Bray-Curtis distance of bacteria (<b>A</b>) and fungi (<b>B</b>). The ellipses in the figure illustrate grouped ellipses formed between sample points in different study areas;PSP: perennial subsidence ponding; SSP: seasonal subsidence ponding; HDZ: horizontal defor-mation zone; NZ: neutral zone; CK: CalvinKlein.</p>
Full article ">Figure 4
<p>Relative abundance of the top 15 of soil bacterial (<b>A</b>) and fungal (<b>B</b>) communities at the phylum level. Relative abundance bacterial (<b>C</b>) and fungal (<b>D</b>) phylum that showed significant differences in different locations of coal. The Kruskal–Wallis H test was used to evaluate the significance of differences between the indicated groups. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 5
<p>Heat map of correlation between of the top 10 bacterial (<b>A</b>) and fungal (<b>B</b>) phylum of soil microorganisms and soil characteristics. * <span class="html-italic">p</span> &lt; 0.05, ** <span class="html-italic">p</span> &lt; 0.01, and *** <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 6
<p>db-RDA analysis based on the Bray–Curtis distance of bacteria (<b>A</b>) and fungi (<b>B</b>).</p>
Full article ">
Back to TopTop