[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

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 (2,248)

Search Parameters:
Keywords = river discharge

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
15 pages, 19478 KiB  
Article
Source Apportionment and Human Health Risks of Potentially Toxic Elements in the Surface Water of Coal Mining Areas
by Yuting Yan, Yunhui Zhang, Zhan Xie, Xiangchuan Wu, Chunlin Tu, Qingsong Chen and Lanchu Tao
Toxics 2024, 12(9), 673; https://doi.org/10.3390/toxics12090673 (registering DOI) - 15 Sep 2024
Abstract
Contamination with potentially toxic elements (PTEs) frequently occurs in surface water in coal mining areas. This study analyzed 34 surface water samples collected from the Yunnan–Guizhou Plateau for their hydrochemical characteristics, spatial distribution, source apportionment, and human health risks. Our statistical analysis showed [...] Read more.
Contamination with potentially toxic elements (PTEs) frequently occurs in surface water in coal mining areas. This study analyzed 34 surface water samples collected from the Yunnan–Guizhou Plateau for their hydrochemical characteristics, spatial distribution, source apportionment, and human health risks. Our statistical analysis showed that the average concentrations of PTEs in the surface water ranked as follows: Fe > Al > Zn > Mn > Ba > B> Ni > Li > Cd > Mo > Cu > Co > Hg > Se > As > Pb > Sb. The spatial analysis revealed that samples with high concentrations of Fe, Al, and Mn were predominantly distributed in the main stream, Xichong River, and Yangchang River. Positive matrix factorization (PMF) identified four sources of PTEs in the surface water. Hg, As, and Se originated from wastewater discharged by coal preparation plants and coal mines. Mo, Li, and B originated from the dissolution of clay minerals in coal seams. Elevated concentrations of Cu, Fe, Al, Mn, Co, and Ni were attributed to the dissolution of kaolinite, illite, chalcopyrite, pyrite, and minerals associated with Co and Ni in coal seams. Cd, Zn, and Pb were derived from coal melting and traffic release. The deterministic health risks assessment showed that 94.12% of the surface water samples presented non-carcinogenic risks below the health limit of 1. Meanwhile, 73.56% of the surface water samples with elevated As posed level III carcinogenic risk to the local populations. Special attention to drinking water safety for children is warranted due to their lower metabolic capacity for detoxifying PTEs. This study provides insight for PTE management in sustainable water environments. Full article
(This article belongs to the Section Metals and Radioactive Substances)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of the Yunnan–Guizhou Plateau in China, (<b>b</b>) location of the study area in the Yunnan–Guizhou Plateau, and (<b>c</b>) location of surface water, groundwater, and mine water sampling sites in the study area (sample size = 34).</p>
Full article ">Figure 2
<p>Box plots with the standard of potentially toxic elements for drinking surface water (sample size = 34).</p>
Full article ">Figure 3
<p>Spatial distribution map of concentrations of potentially toxic elements: (<b>a</b>) Fe, (<b>b</b>) Mn, (<b>c</b>) Cu, (<b>d</b>) Zn, (<b>e</b>) Al, (<b>f</b>) Hg, (<b>g</b>) As, (<b>h</b>) Se, (<b>i</b>) Cd, (<b>j</b>) Pb, (<b>k</b>) Li, (<b>l</b>) B, (<b>m</b>) Ba, (<b>n</b>) Sb, (<b>o</b>) Ni, (<b>p</b>) Co, and (<b>q</b>) Mo (sample size = 34).</p>
Full article ">Figure 4
<p>Source contributions of PTEs based on the PMF model: (<b>a</b>) relative contributions of PTEs to PMF factors and (<b>b</b>) average contributions of PMF factors (sample size = 34).</p>
Full article ">Figure 5
<p>Non-carcinogenic health risks of surface water to children, men, and women: (<b>a</b>) Fe, (<b>b</b>) Mn, (<b>c</b>) Cu, (<b>d</b>) Zn, (<b>e</b>) Al, (<b>f</b>) As, (<b>g</b>) Se, (<b>h</b>) Cd, (<b>i</b>) Li, (<b>j</b>) B, (<b>k</b>) Ba, (<b>l</b>) Sb, (<b>m</b>) Ni, (<b>n</b>) Co, and (<b>o</b>) Mo (sample size = 34).</p>
Full article ">Figure 6
<p>Sensitive non-carcinogenic PTE ranking for HI.</p>
Full article ">Figure 7
<p>Non-carcinogenic and carcinogenic health risks of PTEs in surface water to children, men, and women: (<b>a</b>) hazard index (HI) and (<b>b</b>) cancer risk (CR) (sample size = 34).</p>
Full article ">Figure 8
<p>Spatial distribution map of hazard index (HI) and cancer risk (CR) of PTEs in surface water. HI to (<b>a</b>) children, (<b>b</b>) men, and (<b>c</b>) women and CR to (<b>d</b>) children, (<b>e</b>) men, and (<b>f</b>) women (sample size = 34).</p>
Full article ">
17 pages, 951 KiB  
Article
Multiphase Partitioning of Estrogens in a River Impacted by Feedlot Wastewater Discharge
by Kuo-Hui Yang, Hao-Shen Hung, Wei-Hsiang Huang, Chi-Ying Hsieh and Ting-Chien Chen
Toxics 2024, 12(9), 671; https://doi.org/10.3390/toxics12090671 (registering DOI) - 14 Sep 2024
Viewed by 204
Abstract
Estrogens in river systems can significantly impact aquatic ecosystems. This study aimed to investigate the multiphase partitioning of estrogens in Wulo Creek, Taiwan, which receives animal feedlot wastewater, to understand their distribution and potential environmental implications. Water samples were separated into suspended particulate [...] Read more.
Estrogens in river systems can significantly impact aquatic ecosystems. This study aimed to investigate the multiphase partitioning of estrogens in Wulo Creek, Taiwan, which receives animal feedlot wastewater, to understand their distribution and potential environmental implications. Water samples were separated into suspended particulate matter (SPM), colloidal, and soluble phases using centrifugation and cross-flow ultrafiltration. Concentrations of estrone (E1), 17β-estradiol (E2), and estriol (E3) in each phase were analyzed using LC/MS/MS. Partition coefficients were calculated to assess estrogen distribution among phases. Estrogens were predominantly found in the soluble phase (85.8–87.3%). The risk assessment of estrogen equivalent (EEQ) values suggests that estrogen concentration in water poses a higher risk compared to SPM, with a majority of the samples indicating a high risk to aquatic organisms. The colloidal phase contained 12.7–14.2% of estrogens. The log KCOC values (4.72–4.77 L/kg-C) were significantly higher than the log KOC and log KPOC values (2.02–3.40 L/kg-C) for all estrogens. Colloids play a critical role in estrogen distribution in river systems, potentially influencing their fate, transport, and biotoxicity. This finding highlights the importance of considering colloidal interactions in assessing estrogen behavior in aquatic environments. Full article
(This article belongs to the Special Issue Environmental Transport and Transformation of Pollutants)
Show Figures

Figure 1

Figure 1
<p>Geographical map of sampling site locations (adapted from Hung et al. [<a href="#B37-toxics-12-00671" class="html-bibr">37</a>], and Ministry of Agriculture, Republic of China (Taiwan)).</p>
Full article ">Figure 2
<p>The mass percentages of organic carbon (OC) distributed on colloidal and soluble phases.</p>
Full article ">Figure 3
<p>The mass percentages of estrogens distributed on colloidal and soluble phases: (<b>a</b>) (E1), (<b>b</b>) (E2), and (<b>c</b>) (E3).</p>
Full article ">
22 pages, 3621 KiB  
Article
Estimating Non-Stationary Extreme-Value Probability Distribution Shifts and Their Parameters Under Climate Change Using L-Moments and L-Moment Ratio Diagrams: A Case Study of Hydrologic Drought in the Goat River Near Creston, British Columbia
by Isaac Dekker, Kristian L. Dubrawski, Pearce Jones and Ryan MacDonald
Hydrology 2024, 11(9), 154; https://doi.org/10.3390/hydrology11090154 (registering DOI) - 14 Sep 2024
Viewed by 168
Abstract
Here, we investigate the use of rolling-windowed L-moments (RWLMs) and L-moment ratio diagrams (LMRDs) combined with a Multiple Linear Regression (MLR) machine learning algorithm to model non-stationary low-flow hydrological extremes with the potential to simultaneously understand time-variant shape, scale, location, and probability distribution [...] Read more.
Here, we investigate the use of rolling-windowed L-moments (RWLMs) and L-moment ratio diagrams (LMRDs) combined with a Multiple Linear Regression (MLR) machine learning algorithm to model non-stationary low-flow hydrological extremes with the potential to simultaneously understand time-variant shape, scale, location, and probability distribution (PD) shifts under climate change. By employing LMRDs, we analyse changes in PDs and their parameters over time, identifying key environmental predictors such as lagged precipitation for September 5-day low-flows. Our findings indicate a significant relationship between total August precipitation L-moment ratios (LMRs) and September 5-day low-flow LMRs (τ2-Precipitation and τ2-Discharge: R2 = 0.675, p-values < 0.001; τ3-Precipitation and τ3-Discharge: R2 = 0.925, p-value for slope < 0.001, intercept not significant with p = 0.451, assuming α = 0.05 and a 31-year RWLM), which we later refine and use for prediction within our MLR algorithm. The methodology, applied to the Goat River near Creston, British Columbia, aids in understanding the implications of climate change on water resources, particularly for the yaqan nuʔkiy First Nation. We find that future low-flows under climate change will be outside the Natural Range of Variability (NROV) simulated from historical records (assuming a constant PD). This study provides insights that may help in adaptive water management strategies necessary to help preserve Indigenous cultural rights and practices and to help sustain fish and fish habitat into the future. Full article
(This article belongs to the Section Hydrological and Hydrodynamic Processes and Modelling)
Show Figures

Figure 1

Figure 1
<p>L-moment ratio diagrams (LMRDs) for: (<b>a</b>) August total precipitation (mm) and (<b>b</b>) September 5-day low-flow (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>). Each panel includes the L-Coefficient of Variation (L-CV;<math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>) versus L-skewness (<math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>) and L-kurtosis versus L-skewness ratios.</p>
Full article ">Figure 2
<p>Relationship between L-moment ratios (LMRs) of August total precipitation (mm) and September 5-day low-flow (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>): (<b>a</b>) <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>. <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>-Precipitation and <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>2</mn> </msub> </semantics></math>-Discharge: <span class="html-italic">p</span>-values &lt; 0.001; <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>-Precipitation and <math display="inline"><semantics> <msub> <mi>τ</mi> <mn>3</mn> </msub> </semantics></math>-Discharge: <span class="html-italic">p</span>-value for slope &lt; 0.001, intercept not significant with <span class="html-italic">p</span> = 0.451, assuming <math display="inline"><semantics> <mi>α</mi> </semantics></math> = 0.05 and a 31-year rolling-windowed L-moments (RDLMs).</p>
Full article ">Figure 3
<p>Comparison of predicted and observed L-moments (LMs; testing during training) using a 31-year rolling-window. The plots display the predicted (green) and observed (blue) values for the first (<b>a</b>), second (<b>b</b>), third (<b>c</b>), and fourth (<b>d</b>) LMs. Each subplot includes the Overall Mean Squared Error (MSE) between the predicted and observed values computed by summing and averaging the best-fit model Squared Error (SE) for each step in the forward chaining process. The equations plotted alongside the model are derived from the final (best-fit) iteration (index 38), which demonstrated the lowest SE.</p>
Full article ">Figure 4
<p>Location, scale, and shape parameters estimated using [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>]’s method of L-moments (LMs) for the Generalized Extreme Value (GEV) probability distribution (PD) for the observed (blue) and predicted (testing during training; dashed red) LMs under a 31-year rolling-window.</p>
Full article ">Figure 5
<p>LMRDs using 31-year windows under two Representative Concentration Pathways (RCP) scenarios. Panels (<b>a</b>–<b>c</b>) correspond to the RCP 4.5 scenario, while panels (<b>d</b>–<b>f</b>) correspond to the RCP 8.5 scenario. Diagrams show: (<b>a</b>,<b>d</b>) L-CV/L-skewness, (<b>b</b>,<b>e</b>) L-kurtosis/L-skewness, with theoretical PDs described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] (distributions include Generalized Logistic (GLO), Generalized Extreme Value (GEV), Generalized Pareto (GPA), Generalized Normal (GNO), Pearson Type III (PE3), Wakeby Lower Bound (WAK_LB), and All Distribution Lower Bound (ALL_LB)). Plots (<b>c</b>,<b>f</b>) show the distribution count for each window. The observed LMRs for the 5-day September low-flow at the Water Survey of Canada (WSC) Goat River Near Erikson Hydrometric Gauge Station (<a href="https://wateroffice.ec.gc.ca/report/data_availability_e.html?type=historical&amp;station=08NH004&amp;parameter_type=Flow&amp;wbdisable=true" target="_blank">08NH004</a>) are plotted alongside simulated future data derived from Multiple Linear Regression (MLR) driven with total August precipitation LMs. Future data are generated using a splice of six Coupled Model Intercomparison Project Phase 5 (CMIP5) series downscaled climate models (median of “ACCESS1-0”, “CanESM2”, “CCSM4”, “CNRM-CM5”, “HadGEM2-ES”, and “MPI-ESM-LR” from 2018 to 2100) downloaded using the single cell extraction tool from the Pacific Climate Impacts Consortium (<a href="https://pacificclimate.org/data/gridded-hydrologic-model-output" target="_blank">PCIC</a>). Historical climate data are downloaded from Historical Climate Data Online (HCDO) repository for the Creston station (Climate ID <a href="https://climate.weather.gc.ca/climate_data/daily_data_e.html?hlyRange=%7C&amp;dlyRange=1912-06-01%7C2017-12-31&amp;mlyRange=1912-01-01%7C2007-02-01&amp;StationID=1111&amp;Prov=BC&amp;urlExtension=_e.html&amp;searchType=stnName&amp;optLimit=yearRange&amp;StartYear=1840&amp;EndYear=2024&amp;selRowPerPage=25&amp;Line=0&amp;searchMethod=contains&amp;Month=12&amp;Day=2&amp;txtStationName=Creston&amp;timeframe=2&amp;Year=2017" target="_blank">1142160</a>; available from 1996 to 2017).</p>
Full article ">Figure 6
<p>Results of the L-moments (LMs) derived from Multiple Linear Regression (MLR) fit to a Generalized Extreme Value (GEV) probability distribution (PD) to produce shape, scale, and location parameters: (<b>a</b>) GEV parameters (shape, scale, and location) over 144 rolling windowed time units under Representative Concentration Pathway (RCP) 4.5 and (<b>b</b>) RCP 8.5.</p>
Full article ">Figure 7
<p>LMs derived from MLR fit to a GEV PD to produce shape, scale, and location parameters to derive median flows (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) and confidence intervals (CIs) estimated from percentiles: (<b>a</b>) simulated September 5-day low-flows driven by MLR regression using August total precipitation LMs over a 31-year rolling time window under (<b>a</b>) RCP 4.5 and (<b>b</b>) RCP 8.5. The dashed red line denotes a flow of &lt;1 m<sup>3</sup>/s. Note: each simulation is based on <span class="html-italic">n</span> = 1000 iterations for both panels.</p>
Full article ">Figure 8
<p>Results of the LMs derived from MLR fit to the best-fit probability distribution (PD) (distributions described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>]) to produce shape, scale, and location parameters to derive median flows (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) and confidence intervals (CIs) estimated numerically from percentiles: (<b>a</b>) simulated September 5-day low-flows driven by MLR regression using August total precipitation LMs over a 31-year moving window under RCP 4.5 and (<b>b</b>) RCP 8.5. The dashed red line denotes a flow of &lt;1 m<sup>3</sup>/s. Note: each simulation is based on <span class="html-italic">n</span> = 1000 iterations for both panels.</p>
Full article ">Figure 9
<p>Overall standardized Mean Square Error (MSE) across different window sizes during model training.</p>
Full article ">Figure 10
<p>Sensitivity of window size on location, scale, and shape parameters for September 5-day low-flow estimated using the method of LMs described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] derived from MLR driven by total August precipitation LMs for six common distributions described in [<a href="#B45-hydrology-11-00154" class="html-bibr">45</a>] (Generalized Extreme Value (GEV; (<b>a</b>–<b>c</b>)), Generalized Logistic (GLO; (<b>d</b>–<b>f</b>)), Generalized Normal (GNO; (<b>g</b>–<b>i</b>)), Pearson Type III (PE3; (<b>j</b>–<b>l</b>)), and Generalized Pareto (GPA; (<b>m</b>–<b>o</b>)). The solid line displays data under the Representative Concentration Pathway (RCP) 4.5 emission scenario, while the dashed line displays the RCP 8.5 emissions scenario.</p>
Full article ">Figure 11
<p>Low-flow exceedance and cumulative exceedance probability for the Goat River near Erikson Gauge Station, showing values less than 2.7 cubic meters per second (<math display="inline"><semantics> <mrow> <msup> <mi mathvariant="normal">m</mi> <mn>3</mn> </msup> <mo>/</mo> <mi mathvariant="normal">s</mi> </mrow> </semantics></math>) assuming <span class="html-italic">n</span> = 1000 simulations and a Generalized Extreme Value (GEV) probability distribution (PD). Future data assume a Representative Concentration Pathway (RCP) 4.5 emissions scenario.</p>
Full article ">
24 pages, 5700 KiB  
Article
Temporal Scales of Mass Wasting Sedimentation across the Mississippi River Delta Front Delineated by 210Pb/137Cs Geochronology
by Jeffrey Duxbury, Samuel J. Bentley, Kehui Xu and Navid H. Jafari
J. Mar. Sci. Eng. 2024, 12(9), 1644; https://doi.org/10.3390/jmse12091644 - 13 Sep 2024
Viewed by 322
Abstract
The Mississippi River Delta Front (MRDF) is a subaqueous apron of rapidly deposited and weakly consolidated sediment extending from the subaerial portions of the Birdsfoot Delta of the Mississippi River, long characterized by mass-wasting sediment transport. Four (4) depositional environments dominate regionally (an [...] Read more.
The Mississippi River Delta Front (MRDF) is a subaqueous apron of rapidly deposited and weakly consolidated sediment extending from the subaerial portions of the Birdsfoot Delta of the Mississippi River, long characterized by mass-wasting sediment transport. Four (4) depositional environments dominate regionally (an undisturbed topset apron, mudflow gully, mudflow lobe, and prodelta), centering around mudflow distribution initiated by a variety of factors (hurricanes, storms, and fluid pressure). To better understand the spatiotemporal scales of the events as well as the controlling processes, eight cores (5.8–8.0 m long) taken offshore from the South Pass (SP) and the Southwest Pass (SWP) were analyzed for gamma density, grain size, sediment fabric (X-radiography), and geochronology (210Pb/137Cs radionuclides). Previous work has focused on the deposition of individual passes and has been restricted to <3 m core penetration, limiting its geochronologic completeness. Building on other recent studies, within the mudflow gully and lobe cores, the homogeneous stepped profiles of 210Pb activities and the corresponding decreased gamma density indicate the presence of gravity-driven mass failures. 210Pb/137Cs indicates that gully sedimentary sediment accumulation since 1953 is greater than 580 cm (sediment accumulation rate [SAR] of 12.8 cm/y) in the southwest pass site, and a lower SAR of the South Pass gully sites (2.6 cm/y). This study shows that (1) recent dated mudflow deposits are identifiable in both the SWP and SP; (2) SWP mudflows have return periods of 10.7 y, six times more frequent than at the SP (66.7 y); (3) 210Pb inventories display higher levels in the SWP area, with the highest focusing factors in proximal/gully sedimentation, and (4) submarine landslides in both study areas remain important for sediment transport despite the differences in sediment delivery and discharge source proximity. Full article
Show Figures

Figure 1

Figure 1
<p>Map outlining the Mississippi River Delta Front (MRDF) study site in vicinity of the “birds’ foot”, (<b>a</b>) displaying the subaerial and subaqueous bathymetry with study sites outlined in red. (<b>b</b>) The Southwest Pass and (<b>c</b>) the South Pass display piston core locations, with black dots and dotted lines outlining a select gully–lobe complex within each. Bathymetry is from Baldwin et al. (2018) [<a href="#B3-jmse-12-01644" class="html-bibr">3</a>], imagery is open source “world imagery” from ESRI.</p>
Full article ">Figure 2
<p>Delta front seafloor diagram (adapted from Coleman et al. 1980 [<a href="#B1-jmse-12-01644" class="html-bibr">1</a>]) outlining major morphological features of the study sites. Upper, intermediate, and lower zones of the environment range from 20 to 300 m in depth and feature incising gullies coalescing into mudflow lobes downslope overlying earlier, Holocene-aged deposits [<a href="#B14-jmse-12-01644" class="html-bibr">14</a>].</p>
Full article ">Figure 3
<p>Downcore physical property profiles for piston cores. Gamma density (solid), porosity (dashed), and mean grain size (phi units in black dots with error bars showing standard deviation) are laid out for the Southwest Pass (<b>top</b>) and South Pass (<b>bottom</b>), ordered by depositional environment.</p>
Full article ">Figure 4
<p>Diagnostic X-radiography for each of the depositional environments showing common fabrics present within cores. Red lines indicate possible unconformity locations. By core, (<b>a</b>) PS17-03 undisturbed topset apron with laminated bedding present throughout; (<b>b</b>) PS17-06, a mudflow gully core with large amounts of biogenic gas expansion exacerbated by desiccation with no visible bedding beside an unconformity separating two homogenous layers; (<b>c</b>) mudflow lobe core PS17-07, showing biogenic gas voids below an unconformity; (<b>d</b>) prodelta core PS17-09, with cm-scale sandy layers and abundant burrowing throughout; (<b>e</b>) PS17-24, a mudflow gully core from the South Pass showing a possible unconformity with angled bedding below and homogenous above.</p>
Full article ">Figure 5
<p>Depositional mechanism interpretation displayed over stratigraphic profile, gamma density, and <sup>210</sup>Pb/<sup>137</sup>Cs profiles by pass.</p>
Full article ">Figure 6
<p>CHIRP seismic profiles parallel to shore, progressing distally (<b>A</b>–<b>C</b>) and perpendicular (<b>D</b>), outlined as tracts with corresponding A’–D’ in <a href="#jmse-12-01644-f001" class="html-fig">Figure 1</a>. Identified depositional environments are listed on each transect down to observable seismic basement.</p>
Full article ">Figure 7
<p>Depositional environment analysis. (<b>a</b>) Relative composition of cores by sedimentation mechanism, (<b>b</b>) accumulation rates by depositional environments, (<b>c</b>) calculated mudflow return period (years) by depositional environment.</p>
Full article ">Figure 8
<p>Site-wide radioisotope analysis with (<b>a</b>) <sup>210</sup>Pb radioisotope inventories (<b>top</b>) and (<b>b</b>) <sup>210</sup>Pb index analysis by depositional environment. Concentrations of <sup>210</sup>Pb show preferential deposition in the undisturbed and gully cores of the Southwest Pass.</p>
Full article ">Figure 9
<p>Timeseries of major forcing events (floods/hurricanes/dams) plotted along estimated mudflow occurrence dates by core. Major hurricane occurrences were referenced from the NOAA Historical Hurricane Tracker as category 3+ hurricanes with tracks within 70 miles of the Head of Passes. High-risk hurricanes are those described by Guidroz (2009) [<a href="#B12-jmse-12-01644" class="html-bibr">12</a>], and other focused river discharge (Talbert’s Landing) are referenced from the River Gauges Database (USACE). The first occurrence of <sup>137</sup>Cs (1953) forms a backstop for cores PS17-09, PS17-24, and PS17-30 and a forestop in PS17-03. The arrows indicate the South Pass depositional hiatus of hypopycnal deposition at much lower rates to the base of the core (listed in <a href="#jmse-12-01644-t001" class="html-table">Table 1</a>). Asterix (*) indicates calculation based off of <sup>137</sup>Cs due to the full penetration.</p>
Full article ">
18 pages, 10223 KiB  
Article
Flood Modeling of the June 2023 Flooding of Léogâne City by the Overflow of the Rouyonne River in Haiti
by Rotchild Louis, Yves Zech, Adermus Joseph, Nyankona Gonomy and Sandra Soares-Frazao
Water 2024, 16(18), 2594; https://doi.org/10.3390/w16182594 - 13 Sep 2024
Viewed by 516
Abstract
Evaluating flood risk though numerical simulations in areas where hydrometric and bathymetric data are scarcely available is a challenge. This is, however, of paramount importance, particularly in urban areas, where huge losses of human life and extensive damage can occur. This paper focuses [...] Read more.
Evaluating flood risk though numerical simulations in areas where hydrometric and bathymetric data are scarcely available is a challenge. This is, however, of paramount importance, particularly in urban areas, where huge losses of human life and extensive damage can occur. This paper focuses on the 2–3 June 2023 event at Léogâne in Haiti, where the Rouyonne River partly flooded the city. Water depths in the river have been recorded since April 2022, and a few discharges were measured manually, but these were not sufficient to produce a reliable rating curve. Using a uniform-flow assumption combined with the Bayesian rating curve (BaRatin) method, it was possible to extrapolate the existing data to higher discharges. From there, a rainfall–runoff relation was developed for the site using a distributed hydrological model, which allowed the discharge of the June 2023 event to be determined, which was estimated as twice the maximum conveying capacity of the river in the measurement section. Bathymetric data were obtained using drone-based photogrammetry, and two-dimensional simulations were carried out to represent the flooded area and the associated water depths. By comparing the water depths of 21 measured high-water marks with the simulation results, we obtained a Kling–Gupta Efficiency (KGE) and Nash–Sutcliffe Efficiency (NSE) values of 0.890 and 0.882, respectively. This allows us to conclude that even when only scarce official data are available, it is possible to use field data acquired by low-cost methodologies to build a model that is sufficiently accurate and that can be used by flood managers and decision makers to assess flood risk and vulnerability in Haiti. Full article
Show Figures

Figure 1

Figure 1
<p>Study area: (<b>a</b>) Study site location in Haiti; (<b>b</b>) Rouyonne river channel and its upper watershed; (<b>c</b>) Altitude distribution in the upper watershed.</p>
Full article ">Figure 1 Cont.
<p>Study area: (<b>a</b>) Study site location in Haiti; (<b>b</b>) Rouyonne river channel and its upper watershed; (<b>c</b>) Altitude distribution in the upper watershed.</p>
Full article ">Figure 2
<p>Illustration of image acquisition: (<b>a</b>) Aerial image of the river during the dry season; (<b>b</b>) DJI drone equipped with a GoPro camera (sensor type: 1/2.3” CMOS; camera type: sport/action camera; equivalent focal length: 16.41 mm; lens type: wide angle; aperture: f/2.8).</p>
Full article ">Figure 3
<p>Illustration of the morphological changes in the Rouyonne river channel: (<b>a</b>) Cross-section 54—54 of the Rouyonne River; (<b>b</b>) Bathymetric data comparison between UAV photogrammetry DTM (2022) and the manual survey (2022); (<b>c</b>) Evolution of morphological changes between 2014 and 2022 in cross-section 54—54.</p>
Full article ">Figure 4
<p>Illustration of the damage caused by the 3 June 2023 event: (<b>a</b>) Buildings destroyed by the flood in the town of Léogâne; (<b>b</b>) Pressure sensor broken by flood at the measuring section; (<b>c</b>) High-water mark measurement; (<b>d</b>) Spatial distribution of high-water marks measured for the 2–3 June 2023 event.</p>
Full article ">Figure 4 Cont.
<p>Illustration of the damage caused by the 3 June 2023 event: (<b>a</b>) Buildings destroyed by the flood in the town of Léogâne; (<b>b</b>) Pressure sensor broken by flood at the measuring section; (<b>c</b>) High-water mark measurement; (<b>d</b>) Spatial distribution of high-water marks measured for the 2–3 June 2023 event.</p>
Full article ">Figure 5
<p>Illustration of the unstructured mesh of the study area.</p>
Full article ">Figure 6
<p>Cross-section at the limnimetric station with the equivalent rectangle (discontinuous black line) used in the BaRatin method.</p>
Full article ">Figure 7
<p>Relationships between water depth and discharge at the measurement section: comparison of the Bayesian rating curve with uncertainties and the uniform-flow assumption.</p>
Full article ">Figure 8
<p>Hydrological modeling: (<b>a</b>) Calibration (August 2022); (<b>b</b>) Validation (September 2022).</p>
Full article ">Figure 9
<p>Hydrological modeling applied to the event of 2–3 June 2023.</p>
Full article ">Figure 10
<p>Illustration of the 2–3 June 2023 event simulation.</p>
Full article ">Figure 11
<p>Model evaluation: Comparison between the observed and modeled water depths.</p>
Full article ">Figure 12
<p>Identification of the overflow points on the Rouyonne river: (<b>a</b>) Right bank overtopping to downtown Léogâne; (<b>b</b>) Left bank overtopping where the probe was installed.</p>
Full article ">
28 pages, 8636 KiB  
Article
Karst Hydrological Connections of Lakes and Neoproterozoic Hydrogeological System between the Years 1985–2020, Lagoa Santa—Minas Gerais, Brazil
by Wallace Pacheco Neto, Rodrigo de Paula and Paulo Galvão
Water 2024, 16(18), 2591; https://doi.org/10.3390/w16182591 - 12 Sep 2024
Viewed by 357
Abstract
This study focuses on a complex Brazilian Neoproterozoic karst (hydro)geological and geomorphological area, consisting of metapelitic–carbonate sedimentary rocks of ~740–590 Ma, forming the largest carbonate sequence in the country. At the center of the area lies the Lagoa Santa Karst Environmental Protection Area [...] Read more.
This study focuses on a complex Brazilian Neoproterozoic karst (hydro)geological and geomorphological area, consisting of metapelitic–carbonate sedimentary rocks of ~740–590 Ma, forming the largest carbonate sequence in the country. At the center of the area lies the Lagoa Santa Karst Environmental Protection Area (LSKEPA), located near the Minas Gerais’ state capital, Belo Horizonte, and presents a series of lakes associated with the large fluvial system of the Velhas river under the influence, locally, of carbonate rocks. The hydrodynamics of carbonate lakes remain enigmatic, and various factors can influence the behavior of these water bodies. This work analyzed the hydrological behavior of 129 lakes within the LSKEPA to understand potential connections with the main karst aquifer, karst-fissure aquifer, and porous aquifer, as well as their evolution patterns in the physical environment. Pluviometric surveys and satellite image analysis were conducted from 1984 to 2020 to observe how the lakes’ shorelines behaved in response to meteorological variations. The temporal assessment for understanding landscape evolution proves to be an effective tool and provides important information about the interaction between groundwater and surface water. The 129 lakes were grouped into eight classes representing the hydrological connection patterns with the aquifers in the region, with classes defined for perennial lakes: (1) constantly connected, (2) seasonally disconnected, and (3) disconnected; for intermittent lakes: (4) disconnected during the analyzed time interval, (5) seasonally connected, (6) disconnected, (7) extremely disconnected, and (8) intermittent lakes that connected and stopped drying up. The patterns observed in the variation of lakes’ shorelines under the influence of different pluviometric moments showed a positive correlation, especially in dry periods, where these water bodies may be functioning as recharge or discharge zones of the karst aquifer. These inputs and outputs are conditioned to the well-developed karst tertiary porosity, where water flow in the epikarst moves according to the direction of enlarged karstified fractures, rock foliation planes, and lithological contacts. Other factors may condition the hydrological behavior of the lakes, such as rates of evapotranspiration, intensity of rainfall during rainy periods, and excessive exploitation of water. Full article
(This article belongs to the Special Issue Recent Advances in Karstic Hydrogeology, 2nd Edition)
Show Figures

Figure 1

Figure 1
<p>Geological and location map of the study area highlighting the Lagoa Santa Karst Environmental Protection Area, Minas Gerais, Brazil, and the lakes analyzed in this work. Geology modified from “Projeto Vida” [<a href="#B21-water-16-02591" class="html-bibr">21</a>] and profile modified from [<a href="#B22-water-16-02591" class="html-bibr">22</a>].</p>
Full article ">Figure 2
<p>Example of Landsat satellite images used in this study, representing the rainy season of 1999 and the dry season of the same year.</p>
Full article ">Figure 3
<p>Graphical representation of perimeter variation over the years for the studied lakes (example Lake 28).</p>
Full article ">Figure 4
<p>Flowchart summarizing the steps taken to identify the expansion or contraction behavior of each lake over the analyzed years.</p>
Full article ">Figure 5
<p>Bar chart representing the annual average precipitation between the hydrological years 1984–1985 and 2019–2020. The orange bars indicate precipitation during the dry months (April to September), while the blue bars show precipitation during the wet months (October to March). The red line marks the average precipitation for the 36 years analyzed in this study, with confidence intervals represented by the dashed blue line (positive confidence interval) and the dashed yellow line (negative confidence interval).</p>
Full article ">Figure 6
<p>Bar chart representing rainfall and drought cycles. Values above the historical average (black dashed line) during dry precipitation cycles were defined as atypical dry hydrological years, while values below the historical average (black dashed line) during wet cycles were defined as atypical wet hydrological years.</p>
Full article ">Figure 7
<p>Graph of the precipitation cycles, showing their averages, and trend lines representing the precipitation variation within each cycle. Below is a table summarizing the averages and equations of the trend lines for each cycle, with angular coefficients in blue (positive) and red (negative).</p>
Full article ">Figure 8
<p>Identification and distribution of perennial lakes (in blue) and intermittent lakes (in orange) in the study area. (<b>A</b>) and (<b>B</b>): Highlights of some lakes.</p>
Full article ">Figure 9
<p>Graphical representation of the perimeter variation of perennial lake 23 over the time interval used. The precipitation cycles and the trend lines of perimeter variation in each cycle can be observed. The table below provides the trend line equations within each cycle, along with their positive and negative angular coefficients.</p>
Full article ">Figure 10
<p>Graphical representation with examples of the behavior of perennial lakes that are constantly connected (<b>a</b>), seasonally disconnected perennial lakes (<b>b</b>), and disconnected perennial lakes (<b>c</b>).</p>
Full article ">Figure 11
<p>Illustration of the proposed classes of hydrological connection for the analyzed perennial lakes.</p>
Full article ">Figure 12
<p>Graphical representation, with examples of the behavior of intermittent lakes that disconnected from the aquifer at some point (<b>a</b>), intermittently connected lakes (<b>b</b>), disconnected intermittent lakes (<b>c</b>), extremely disconnected intermittent lakes (<b>d</b>), and fully connected intermittent lakes (<b>e</b>).</p>
Full article ">
15 pages, 710 KiB  
Article
Trends and Drivers of Flood Occurrence in Germany: A Time Series Analysis of Temperature, Precipitation, and River Discharge
by Mohannad Alobid, Fatih Chellai and István Szűcs
Water 2024, 16(18), 2589; https://doi.org/10.3390/w16182589 - 12 Sep 2024
Viewed by 504
Abstract
Floods in Germany have become increasingly frequent and severe over recent decades, with notable events in 2002, 2013, and 2021. This study examines the trends and drivers of flood occurrences in Germany from 1990 to 2024, focusing on the influence of climate-change-related variables, [...] Read more.
Floods in Germany have become increasingly frequent and severe over recent decades, with notable events in 2002, 2013, and 2021. This study examines the trends and drivers of flood occurrences in Germany from 1990 to 2024, focusing on the influence of climate-change-related variables, such as temperature, precipitation, and river discharge. Using a comprehensive time series analysis, including Auto-Regressive Integrated Moving Average (ARIMA) and Artificial Neural Network (ANN) models and correlation and regression analyses, we identify significant correlations between these climatic variables and flood events. Our findings indicate that rising temperatures (with a mean of 8.46 °C and a maximum of 9 °C) and increased precipitation (averaging 862.26 mm annually)are strongly associated with higher river discharge (mean 214.6 m3/s) and more frequent floods (mean 197.94 events per year). The ANN model outperformed the ARIMA model in flood forecasting, showing lower error metrics (e.g., RMSE of 10.86 vs. 18.83). The analysis underscores the critical impact of climate change on flood risks, highlighting the necessity of adaptive flood-management strategies that incorporate the latest climatic and socio-economic data. This research contributes to the understanding of flood dynamics in Germany and provides valuable insights into future flood risks. Combining flood management with groundwater recharge could effectively lower flood risks and enhance water resources’ mitigation and management. Full article
Show Figures

Figure 1

Figure 1
<p>Time series plot of flood events, rainfall, river discharge, and temperature in Germany from 1990 to 2024.</p>
Full article ">Figure 2
<p>Forecast from NNAR (3,2) of flood events from2025 to 2034.</p>
Full article ">Figure 3
<p>Forecast from ARIMA (0,1,1) of flood events (m<sup>3</sup>/s) from 2025 to 2034.</p>
Full article ">
16 pages, 3006 KiB  
Article
Biomonitoring of Waters and Tambacu (Colossoma macropomum × Piaractus mesopotamicus) from the Amazônia Legal, Brazil
by Karuane Saturnino da Silva Araújo, Thiago Machado da Silva Acioly, Ivaneide Oliveira Nascimento, Francisca Neide Costa, Fabiano Corrêa, Ana Maria Gagneten and Diego Carvalho Viana
Water 2024, 16(18), 2588; https://doi.org/10.3390/w16182588 - 12 Sep 2024
Viewed by 355
Abstract
Fish farming is increasingly important globally and nationally, playing a crucial role in fish production for human consumption. Monitoring microbiological and chemical contaminants from water discharge is essential to mitigate the risk of contaminating water and fish for human consumption. This study analyzes [...] Read more.
Fish farming is increasingly important globally and nationally, playing a crucial role in fish production for human consumption. Monitoring microbiological and chemical contaminants from water discharge is essential to mitigate the risk of contaminating water and fish for human consumption. This study analyzes the physicochemical and E. coli parameters of water and tambacu fish muscles (Colossoma macropomum × Piaractus mesopotamicus) in Western Maranhão, Brazil. It also includes a qualitative characterization of zooplankton in the ponds. Samples were collected from tambacu ponds in a dam system fed by natural watercourses from the Tocantins River tributaries, located at the connection of the Brazilian savanna and Amazon biomes. The physicochemical and E. coli parameters of water did not meet national standards. The zooplankton community included Rotifera, Cladocera, Copepoda, and Protozoa representatives, with no prior studies on zooplankton in the region, making these findings unprecedented. The biological quality of freshwater is crucial in fish farming, as poor quality can lead to decreased productivity and fish mortality, raising significant food safety concerns. The water quality studied is related to the potential influence of untreated wastewater as a source of contamination, leaving the studied region still far from safe water reuse practices. The findings on chemical and E. coli contamination of fish farming waters concern human health and emphasize the need for appropriate regulations. Full article
Show Figures

Figure 1

Figure 1
<p>Location and key characteristics of the study area, Maranhão, Brazil.</p>
Full article ">Figure 2
<p>Specimen of tambacu (<span class="html-italic">Colossoma macropomum × Piaractus mesopotamicus</span>) from the Amazônia legal, Brazil.</p>
Full article ">Figure 3
<p>Zooplankton species sampled in the water of tambacu fish farming tanks.</p>
Full article ">
20 pages, 10622 KiB  
Article
Machine Learning Model for River Discharge Forecast: A Case Study of the Ottawa River in Canada
by M. Almetwally Ahmed and S. Samuel Li
Hydrology 2024, 11(9), 151; https://doi.org/10.3390/hydrology11090151 - 12 Sep 2024
Viewed by 202
Abstract
River discharge is an essential input to hydrosystem projects. This paper aimed to modify the group method of data handling (GMDH) to create a new artificial intelligent forecast model (abbreviated as MGMDH) for predicting discharges at river cross-sections (CSs). The basic idea was [...] Read more.
River discharge is an essential input to hydrosystem projects. This paper aimed to modify the group method of data handling (GMDH) to create a new artificial intelligent forecast model (abbreviated as MGMDH) for predicting discharges at river cross-sections (CSs). The basic idea was to optimise the weights for selected hydrometric and meteorological predictors. One novelty of this study was that MGMDH could take the discharge observed from a neighbouring CS as a predictor when observations from the CS of interest had ceased. Another novelty was that MGMDH could include meteorological parameters as extra predictors. The model was validated using data from natural rivers. For given lead times, MGMDH automatically determined the best forecast equations, consistent with physical river hydraulics laws. This automation minimised computing time while improving accuracy. The model gave reliable forecasts, with a coefficient of determination greater than 0.978. For lead times close to the advection time from upstream to the CS of interest, the forecast had the highest reliability. MGMDH results compared well with some other machine learning models, like neural networks and the adaptive structure of the group method of data handling. It has potential applications for efficiently forecasting discharge and offers a tool to support flood management. Full article
(This article belongs to the Section Water Resources and Risk Management)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Close-up view of the Ottawa River between hydrometric stations 02KF009 (CS upstream or CSU) and 02KF005 (CS downstream or CSD); (<b>b</b>) broad view of the stream network, watershed boundaries, and outlet points, illustrating how these watersheds flow into and connect with the Ottawa River.</p>
Full article ">Figure 2
<p>Schematic time series of continuous observations of <span class="html-italic">q</span> (solid black curve), discontinued observations of <span class="html-italic">q</span> (solid red curve), and predicted future values of <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>q</mi> </mrow> <mo stretchy="false">^</mo> </mover> </mrow> </semantics></math> (dashed curves).</p>
Full article ">Figure 3
<p>Definition diagram of river flow: (<b>a</b>) top view of the river channel; (<b>b</b>) CS at upstream (CSU) with discharge <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>U</mi> </mrow> </msub> </mrow> </semantics></math> and water level <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>U</mi> </mrow> </msub> </mrow> </semantics></math> (above a certain reference datum); (<b>c</b>) CS at downstream (CSD) with discharge <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> and water level <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 4
<p>Flowchart of the methods for river discharge forecast.</p>
Full article ">Figure 5
<p>Time series of hourly averaged variable: (<b>a</b>) discharge <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>; (<b>b</b>) water level <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math>, observed from the CS of interest (02KF005); (<b>c</b>) discharge <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>U</mi> </mrow> </msub> </mrow> </semantics></math>; (<b>d</b>) water level <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>η</mi> </mrow> <mrow> <mi>U</mi> </mrow> </msub> </mrow> </semantics></math>, observed from 02KF009, covering a period of 180 days (1 January–30 June 2023). The dotted lines divide the time series into two parts: one for model training, and the other for model testing (the same in subsequent figures).</p>
Full article ">Figure 6
<p>Time series of observed hourly averaged variable: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>T</mi> </mrow> </semantics></math>; (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>θ</mi> </mrow> </semantics></math>; (<b>c</b>) <math display="inline"><semantics> <mrow> <mi>ϕ</mi> </mrow> </semantics></math>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>P</mi> </mrow> <mrow> <mi>a</mi> <mi>t</mi> <mi>m</mi> </mrow> </msub> </mrow> </semantics></math>; (<b>e</b>) <math display="inline"><semantics> <mrow> <mi>P</mi> </mrow> </semantics></math> for the period of 1 January–30 June 2023 at WMO station (ID: 71063) located at 45°23′00″ N, 75°43′00″ W.</p>
Full article ">Figure 7
<p>Values of <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>q</mi> </mrow> <mo stretchy="false">^</mo> </mover> </mrow> </semantics></math> predicted from Equation (14) for lead times: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 2, (<b>b</b>) 4, (<b>c</b>) 6, (<b>d</b>) 8, (<b>e</b>) 10, (<b>f</b>) 12, (<b>g</b>) 16, and (<b>h</b>) 18 h, in comparison with observed <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>q</mi> </mrow> <mrow> <mi>D</mi> </mrow> </msub> </mrow> </semantics></math> (test data points in <a href="#hydrology-11-00151-f005" class="html-fig">Figure 5</a>a).</p>
Full article ">Figure 8
<p>Time series of hourly-averaged discharges, observed at CSD (black curve), and forecasted using Equation (14) for the training period (blue curve) and for the testing period (red curve). The forecast is for lead times: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 2; (<b>b</b>) 4; (<b>c</b>) 6; (<b>d</b>) 8; (<b>e</b>) 10; (<b>f</b>) 12; (<b>g</b>) 16; and (<b>h</b>) 18 h.</p>
Full article ">Figure 9
<p>Performance of the forecast model: (<b>a</b>) AIC <span class="html-italic">c</span>; (<b>b</b>) normalised RMSE <math display="inline"><semantics> <mrow> <mover accent="true"> <mrow> <mi>ε</mi> </mrow> <mo stretchy="false">~</mo> </mover> </mrow> </semantics></math>; (<b>c</b>) the coefficient of determination <math display="inline"><semantics> <mrow> <msup> <mrow> <mi>R</mi> </mrow> <mrow> <mn>2</mn> </mrow> </msup> </mrow> </semantics></math>; (<b>d</b>) mean absolute relative error <math display="inline"><semantics> <mrow> <mfenced open="|" close="|" separators="|"> <mrow> <msub> <mrow> <mi>ε</mi> </mrow> <mrow> <mi>a</mi> </mrow> </msub> </mrow> </mfenced> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>Reliability (Equation (7)) of the best model functions for: (<b>a</b>) the case of discontinued discharge observation; and (<b>b</b>) the case of continuous discharge observation.</p>
Full article ">Figure 11
<p>Time series of 15-min-averaged discharges, observed at CSD of the Boise River (black curve), and forecasted for the training period (blue curve) and for the testing period (red curve). The forecast is for lead times: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 2; (<b>b</b>) 4; (<b>c</b>) 6; (<b>d</b>) 8; (<b>e</b>) 10; (<b>f</b>) 12; (<b>g</b>) 18; and (<b>h</b>) 24 h.</p>
Full article ">Figure 12
<p>Comparison of performance between MGMDH and other MLMs. The lead time is: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 12 h; and (<b>b</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 2 h.</p>
Full article ">Figure 13
<p>Time series of hourly averaged discharges, observed at CSD of the Missouri River (black curve), and forecasted for the training period (blue curve) and for the testing period (red curve). The forecast is for lead times: (<b>a</b>) <math display="inline"><semantics> <mrow> <mi>δ</mi> <mi>t</mi> </mrow> </semantics></math> = 2; (<b>b</b>) 4; (<b>c</b>) 6; (<b>d</b>) 8; (<b>e</b>) 10; (<b>f</b>) 12; (<b>g</b>) 18; and (<b>h</b>) 24 h.</p>
Full article ">
28 pages, 4703 KiB  
Article
Modeling the Impact of Urban and Industrial Pollution on the Quality of Surface Water in Intermittent Rivers in a Semi-Arid Mediterranean Climate
by Abdelillah Bouriqi, Naaila Ouazzani and Jean-François Deliege
Hydrology 2024, 11(9), 150; https://doi.org/10.3390/hydrology11090150 - 11 Sep 2024
Viewed by 376
Abstract
Ensuring the protection of the aquatic environment and addressing the water scarcity and degradation of water quality in the Mediterranean region pose significant challenges. This study specifically aims to assess the impact of urban and industrial pollution on the ZAT River water quality. [...] Read more.
Ensuring the protection of the aquatic environment and addressing the water scarcity and degradation of water quality in the Mediterranean region pose significant challenges. This study specifically aims to assess the impact of urban and industrial pollution on the ZAT River water quality. The study exploits a combination of field measurements and mathematical simulations using the PEGASE model. The objective is to evaluate how water quality changes throughout the different seasons and to determine whether olive oil factories discharge industrial wastewater into the river. The study reveals that the river water quality remains relatively stable along its course, up to km 64 in winter and km 71.77 in summer, where poor water quality is recorded. This degradation can be attributed to multiple factors. One of these factors is the discharge of industrial wastewater, which accounts for 47% of the COD pollution load. This industrial wastewater is released into the river without treatment during the production period (January–February) and inactivity period (March–May). The combined impact of urban and industrial wastewater is also associated with the decrease in water flow resulting from water withdrawals due to irrigation canals and groundwater recharge, which both contribute to the observed changes in river water quality. Importantly, field measurements combined with results obtained from the calibrated model provide compelling evidence of unauthorized wastewater discharges from the olive oil factories into the river. These results emphasize the need for stricter regulation, such as developing water quality monitoring strategies based on the use of modeling methodologies. They also emphasize the importance of improving wastewater management practices, such as setting up treatment plants for different sources of pollution or developing a co-treatment plant to mitigate the adverse impact of industrial pollution on river water quality. Full article
Show Figures

Figure 1

Figure 1
<p>Geographic location of the studied stations (S1 to S9) in ZAT River, Morocco.</p>
Full article ">Figure 2
<p>A workflow scheme of the methodology used in this study.</p>
Full article ">Figure 3
<p>Preprocessing of geographical data with PEGASE.</p>
Full article ">Figure 4
<p>Longitudinal evolution of physicochemical parameters before calibration (blue line) and after calibration (green line) compared with the measured value (black dots) along the ZAT River on 15 June 2021.</p>
Full article ">Figure 5
<p>Longitudinal evolution of physicochemical parameters along the ZAT River on 19 January 2021, with simulated value in blue line and measured value in black dots.</p>
Full article ">Figure 6
<p>Longitudinal evolution of physicochemical parameters along the ZAT River on 16 March 2021, simulated value (blue line), measured value (black dots).</p>
Full article ">Figure 7
<p>Longitudinal evolution of physicochemical parameters along the ZAT River on 1 June 2021, simulated value (blue line), measured value (black dots).</p>
Full article ">Figure 8
<p>Temporal evolution of physicochemical parameters along the ZAT River in point 9, at 71.77 km, simulated value (blue line), measured value (black dots).</p>
Full article ">Figure 9
<p>Temporal evolution of chemical oxygen demand parameters along the ZAT River at point 9, at 71.77 km, simulated value (blue line), measured value (black dots).</p>
Full article ">Figure 10
<p>First scenario (the oil mills discharged 100% of their daily wastewater into the river over two months in January and February).</p>
Full article ">Figure 11
<p>Final scenario (The oil mills discharged 10% of their daily wastewater in January and February, 50% in March, and 20% between 1 April and 15 May, with 0% discharge of wastewater from the oil mills between 16 April and 31 June).</p>
Full article ">
16 pages, 18130 KiB  
Article
Two-Way Coupling of the National Water Model (NWM) and Semi-Implicit Cross-Scale Hydroscience Integrated System Model (SCHISM) for Enhanced Coastal Discharge Predictions
by Hongyuan Zhang, Dongliang Shen, Shaowu Bao and Pietrafesa Len
Hydrology 2024, 11(9), 145; https://doi.org/10.3390/hydrology11090145 - 10 Sep 2024
Viewed by 289
Abstract
This study addresses the limitations of and the common challenges faced by one-dimensional river-routing methods in hydrological models, including the National Water Model (NWM), in accurately representing coastal regions. We developed a two-way coupling between the NWM and the Semi-implicit Cross-scale Hydroscience Integrated [...] Read more.
This study addresses the limitations of and the common challenges faced by one-dimensional river-routing methods in hydrological models, including the National Water Model (NWM), in accurately representing coastal regions. We developed a two-way coupling between the NWM and the Semi-implicit Cross-scale Hydroscience Integrated System Model (SCHISM). The approach demonstrated improvements in modeling coastal river dynamics, particularly during extreme events like Hurricane Matthew. The coupled model successfully captured tidal influences, storm surge effects, and complex river–river interactions that the standalone NWM missed. The approach revealed more accurate representations of peak discharge timing and magnitude as well as water storage and release in coastal floodplains. However, we also identified challenges in reconciling variable representations between hydrological and hydraulic models. This work not only enhances the understanding of coastal–riverine interactions but also provides valuable insights for the development of next-generation hydrological models. The improved modeling capabilities have implications for flood forecasting, coastal management, and climate change adaptation in vulnerable coastal areas. Full article
(This article belongs to the Section Hydrological and Hydrodynamic Processes and Modelling)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of the two-way coupling between the NWM and the SCHISM. The left side illustrates the NWM to SCHISM coupling at the upstream boundary of the SCHISM domain (blue arrows) where river discharge (<math display="inline"><semantics> <mi>Q</mi> </semantics></math>) was converted to velocity (<math display="inline"><semantics> <mi>V</mi> </semantics></math>) using the cross-sectional area (<math display="inline"><semantics> <mi>A</mi> </semantics></math>). The right side shows the SCHISM to NWM coupling at the red cross-sections along the river channel (red lines) where velocities (<math display="inline"><semantics> <mrow> <msub> <mi>V</mi> <mi>i</mi> </msub> </mrow> </semantics></math>) were integrated across a cross-section to calculate discharge (<math display="inline"><semantics> <mi>Q</mi> </semantics></math>) using water depths (<math display="inline"><semantics> <mrow> <msub> <mi>h</mi> <mi>i</mi> </msub> </mrow> </semantics></math>) and segment widths (<math display="inline"><semantics> <mrow> <msub> <mi>w</mi> <mi>i</mi> </msub> </mrow> </semantics></math>). The bottom part depicts the interface between the NWM’s 1D channel and the SCHISM’s 2D coastal domain.</p>
Full article ">Figure 2
<p>Study area and domains for different model components in this coupled model. The northeastern coast of South Carolina was selected as the domain for the SCHISM. The color in SCHISM doamin represents elevation, with negative values indicating bathymetry. The entire Pee Dee Basin was the domain for the NWM. The bold blue lines delineate the river channel. USGS stations 02135200 and 02110704 were used to provide upstream discharge data. USGS stations 02110802, 021108125, and 02110815 were used to provide water-level data for the model validation.</p>
Full article ">Figure 3
<p>(<b>a</b>) Water levels near an estuary (from the ROMS); (<b>b</b>) averaged rainfall intensity for the whole Pee Dee Basin (from NLDAS); (<b>c</b>) wind vectors at latitude −79.16 and longitude 33.19, near the estuary (from CFS).</p>
Full article ">Figure 4
<p>Validation of coupled modeling results. The left panel shows the comparison of water levels between USGS observations and the SCHISM results. ed lines are the result of the SCHISM driven by NWM-produced discharge; blue lines are the result of the SCHISM driven by USGS-observed discharge from two (Waccamaw and Pee Dee) river inlets; black lines are the water-level observations from USGS gauge stations. The color in SCHISM doamin represents elevation, with negative values indicating bathymetry.</p>
Full article ">Figure 5
<p>Coupling model result. In total, 15 points along the river channel were selected as sampling points (the numbers were reach numbers defined in the NWM domain) for the comparison of discharge from “coupling” (red lines on right) and “without coupling” (green lines on right) results. Three points were selected as examples to show the river discharge changes under the influence of importing downstream water-level-change data. The color represents elevation, with negative values indicating bathymetry.</p>
Full article ">Figure 6
<p>River–river interactions observed in the SCHISM results (EXP1). Three times were selected to show the flow direction change caused by river–river interactions: (<b>a</b>) 10 October, 12:00 UTC; (<b>b</b>) 11 October, 02:00 UTC; and (<b>c</b>) 11 October, 12:00 UTC. The colors on the arrows indicate water speed from 0–1.6 m/s. The color on background represents elevation, with negative values indicating bathymetry.</p>
Full article ">Figure 7
<p>Comparisons of water levels from the SCHISM, water levels (head + elevation) from the NWM, and water-level observations from three USGS stations. (<b>a</b>) UGSG 02110802, (<b>b</b>) USGS 021108125, and (<b>c</b>) USGS 02110815. There were significant differences in water-level results between the SCHISM and the NWM.</p>
Full article ">
15 pages, 12106 KiB  
Article
Evaluating the Non-Stationarity, Seasonality and Temporal Risk to Water Resources in the Wei River Basin
by Xin Yuan and Fiachra O’Loughlin
Water 2024, 16(17), 2513; https://doi.org/10.3390/w16172513 - 4 Sep 2024
Viewed by 423
Abstract
Due to the changing climate and human activity, more and more researchers started to focus on non-stationarity in hydrology. In the Wei River Basin, which is the largest tributary of the Yellow River, there is a significant reduction in the total amount of [...] Read more.
Due to the changing climate and human activity, more and more researchers started to focus on non-stationarity in hydrology. In the Wei River Basin, which is the largest tributary of the Yellow River, there is a significant reduction in the total amount of water resources which has been found in past decades. Additionally, the distribution of water resources within the basin is unbalanced, with the lower reaches and southern regions having relatively abundant water resources and other regions lacking these resources. Within this situation, it is important to consider the spatial aspect of water resource management. Four non-stationarity detection methods have been applied to investigate variation in seasonal discharge series. Two meteorological factors have also been analyzed. Based on test results and Köppen Geiger Climate classification, the water resource management has been investigated spatially. As for results, the Baojixia Channel has significant impact on the abrupt change of discharge, while the precipitation and temperature may have an impact on the discharge trend change. In addition, there was no clear evidence to prove that the climate zones impact spatially on the non-stationarity of discharge. Full article
(This article belongs to the Section Hydrology)
Show Figures

Figure 1

Figure 1
<p>Locations of discharge and meteorological stations overlaying the climate zones. Catchment outlines are shown in red.</p>
Full article ">Figure 2
<p>Discharge, precipitation, and temperature for the annual and seasonal average time-series at Xianyang station, where the vertical lines show the locations of the corresponding abrupt changes (only discharge and temperature have change points).</p>
Full article ">Figure 3
<p>A plot of the annual average discharge time-series of daily and monthly data for Linjiacun station.</p>
Full article ">
21 pages, 10998 KiB  
Article
Developing Sustainable Groundwater for Agriculture: Approach for a Numerical Groundwater Flow Model in Data-Scarce Sia Kouanza, Niger
by Alexandra Lutz, Yahaya Nazoumou, Adamou Hassane, Diafarou Moumouni Ali, Abdou Guero, Susan Rybarski and David Kreamer
Water 2024, 16(17), 2511; https://doi.org/10.3390/w16172511 - 4 Sep 2024
Viewed by 439
Abstract
The area of Sia Kouanza in the Sahel of southwestern Niger is a potential location for expanding agriculture through irrigation with groundwater. Agriculture is key to supporting smallholders and promoting food security. As plans proceed, questions include how much water is available, how [...] Read more.
The area of Sia Kouanza in the Sahel of southwestern Niger is a potential location for expanding agriculture through irrigation with groundwater. Agriculture is key to supporting smallholders and promoting food security. As plans proceed, questions include how much water is available, how is groundwater replenished, many hectares to develop, and where to locate the wells. While these questions can be addressed with a model, it is difficult to find detailed procedures, especially when data are scarce. How can we use existing information to develop a model of a natural system where groundwater development will take place? We describe an approach that can be employed in data-scarce areas where similar questions are being asked. The approach includes setting details; conceptual model development; water balance; numerical code MODFLOW; model construction, calibration, and statistics; and result interpretation. Conceptual model component estimates are derived from field data: recharge, evapotranspiration, wetlands discharge, existing extraction, and river stages. When field data are not available or scarce, we employ other sources and describe how they are validated with field data or analog sites. The calibrated steady-state model gives a water balance of 22 × 106 m3/yr with inflows (recharge 22 × 106 m3/yr) and outflows (extraction 7.2 × 105 m3/yr, wetlands 5.7 × 106 m3/yr, evapotranspiration 11.9 × 106 m3/yr). The model is a point of departure; approaches for transient and predictive models, which can be used to simulate changes in irrigation pumping volumes and drought, for example, will be described subsequently. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>Maps showing: (<b>a</b>) the location of Sia Kouanza in the Sahel of southwestern Niger, West Africa; and (<b>b</b>) the outline of the model domain is in black, and the area of primary interest for development—the terrasse—is outlined in pink.</p>
Full article ">Figure 2
<p>Diagram of the approach employed for the Sia Kouanza model, describing the setting, conceptual model, numerical model, calibration, and steady-state model.</p>
Full article ">Figure 3
<p>Maps showing: (<b>a</b>) the location of the Sia Kounza modeling domain in black, the terrasse area of primary interest outlined in pink, and hydrogeologic setting of the domain in yellow and green; and (<b>b</b>) delineation of kori watersheds outlined in black and denoted as BV1 through BV8, ephemeral surface water runoff as blue lines, and chloride mass balance (CMB) sample sites shown as triangles.</p>
Full article ">Figure 4
<p>Average monthly Niger River flow (between 1952 and 2021) and precipitation (between 1981 and 2021) in the Sia Kouanza area. Seasonal variability is observed for both precipitation and river flow.</p>
Full article ">Figure 5
<p>One-to-one plot of monthly CHIRPS precipitation and measured monthly precipitation between 2015 and 2021.</p>
Full article ">Figure 6
<p>Sketch of the chloride mass balance (CMB) approach. Chloride originates from precipitation (P) and is transported to groundwater by water infiltrating the subsurface (blue arrows) as recharge (R).</p>
Full article ">Figure 7
<p>Identification of vegetation from satellite imagery. (<b>a</b>) Smaller shrubs were identified from the higher resolution Lidar flight imagery collected over the terrasse area for use in this study, within a 250 m × 250 m square. (<b>b</b>) Trees were identified from Landsat imagery within a 1 km × 1 km square.</p>
Full article ">Figure 8
<p>Maps showing: (<b>a</b>) the delineation of riparian areas (green) along the southwest boundary, which is also the river and locations of wetlands (blue) in and near the terrasse (pink) believed to be fed by groundwater; and (<b>b</b>) locations and populations of existing communities shown in the terrasse (outlined in pink) and larger modeling domain (outlined in black).</p>
Full article ">Figure 9
<p>Maps showing: (<b>a</b>) location of aquifer pump test locations in and adjacent to the terrasse (pink) that were used for pilot points of hydraulic conductivity (K); and (<b>b</b>) Distribution of K values based on pilot points of the aquifer pump test locations for alluvium in Layer 1 of the groundwater flow model.</p>
Full article ">Figure 10
<p>Maps showing: (<b>a</b>) residual head, or observed groundwater level subtracted from the simulated groundwater level, in the steady-state model where larger circles show larger differences and smaller circles show smaller differences; and (<b>b</b>) contours of simulated head (groundwater level) shown as meters above sea level (masl) and generally following the topography.</p>
Full article ">Figure 11
<p>One-to-one plot of circles that represent observed heads and the corresponding simulated heads in the steady-state model. No extreme departures or biases (generally high or low) are observed.</p>
Full article ">
11 pages, 7567 KiB  
Proceeding Paper
Towards an Automatic Tool for Resilient Waterway Transport: The Case of the Italian Po River
by Maria Luisa Villani, Ebrahim Ehsanfar, Sohith Dhavaleswarapu, Alberto Agnetti, Luca Crose, Giancarlo Focherini and Sonia Giovinazzi
Eng. Proc. 2024, 68(1), 64; https://doi.org/10.3390/engproc2024068064 - 4 Sep 2024
Viewed by 134
Abstract
Improved navigability can enhance inland waterway transportation efficiency, contributing to synchro-modal logistics and promoting sustainable development in regions that can benefit from the presence of considerable waterways. Modern technological solutions, such as digital twins in corridor management systems, must integrate functions of navigability [...] Read more.
Improved navigability can enhance inland waterway transportation efficiency, contributing to synchro-modal logistics and promoting sustainable development in regions that can benefit from the presence of considerable waterways. Modern technological solutions, such as digital twins in corridor management systems, must integrate functions of navigability forecasts that provide timely and reliable information for safe trip planning. This information needs to account for the type of vessel and for the environmental and geomorphological characteristics of each navigation trait. This paper presents a case study, within the EU project CRISTAL, focusing on the Italian Po River, of which the navigability forecast requirements of a digital twin are illustrated. Preliminary results to deliver navigability risk information were obtained. In particular, the statistical correlation of water discharge and water depth, computed from historical data, suggested that efficient forecast models for navigability risk, given some water discharge forecasts, could be built. To this aim, the LSTM (long-short-term-memory) technique was used on the same data to provide models linking water discharge and water depth predictions. Future work involves further testing these models with updated real data and integrating outcomes with climatic and infrastructure management information to enhance the accuracy of the risk information. Full article
(This article belongs to the Proceedings of The 10th International Conference on Time Series and Forecasting)
Show Figures

Figure 1

Figure 1
<p>Flood and Drought Early Warning System of the Po River basin. The figure highlights the monitoring points and a type of chart provided by the tool by combining the source data.</p>
Full article ">Figure 2
<p>Monitored critical sections in the main Po River where different colors represent the different stretches.</p>
Full article ">Figure 3
<p>Draft of a navigability risk forecast matrix. A navigability risk level (red = high risk, yellow = medium risk, and green = low risk), based on the navigability forecast, will be provided for each stretch and for different vessel classes (according to the draught) for the next 10 days (gg = day).</p>
Full article ">Figure 4
<p>Surveyed daily water depth collected at each critical section and river discharge data recorded at each monitoring section of the river Po River derived from the monitoring network. Data source: <a href="http://www.agenziapo.it" target="_blank">www.agenziapo.it</a>.</p>
Full article ">Figure 5
<p>The correlation between daily water depth and river discharge in the Po River. The red trendline highlights the relationship with a computed Pearson correlation coefficient.</p>
Full article ">Figure 6
<p>Year-long trend of water discharge and water depth data at a station.</p>
Full article ">Figure 7
<p>Water depth time series decomposition. Trend and seasonal components revealed through additive seasonal decomposition offer insights into temporal dynamics.</p>
Full article ">Figure 8
<p>Monthly Box Plots (2021) illustrating the distribution and variability of river depth (<b>a</b>) and discharge (<b>b</b>) at the station.</p>
Full article ">Figure 9
<p>Discharge forecast from the probabilistic processing of DEWS/FEWS Po early warning systems data.</p>
Full article ">Figure 10
<p>Problem statement representation of the deep learning method.</p>
Full article ">Figure 11
<p>Development process of a prediction model.</p>
Full article ">Figure 12
<p>Water depth forecast vs. water depth observations for critical section 1 of stretch X. The blue line indicates the lowest water depth before the last 100 days.</p>
Full article ">Figure 13
<p>Water depth forecast vs. water depth observations for critical section 2 of stretch Y.</p>
Full article ">Figure 14
<p>Water discharge (blue) and water depth (brown) time series: (<b>a</b>) critical section 1 of stretch X; (<b>b</b>) critical section 2 of stretch Y.</p>
Full article ">
13 pages, 5514 KiB  
Article
Water Storage–Discharge Relationship with Water Quality Parameters of Carhuacocha and Vichecocha Lagoons in the Peruvian Puna Highlands
by Samuel Pizarro, Maria Custodio, Richard Solórzano-Acosta, Duglas Contreras and Patricia Verástegui-Martínez
Water 2024, 16(17), 2505; https://doi.org/10.3390/w16172505 - 4 Sep 2024
Viewed by 537
Abstract
Most Andean lakes and lagoons are used as reservoirs to manage hydropower generation and cropland irrigation, which, in turn, alters river flow patterns through processes of storage and discharge. The Carhuacocha and Vichecocha lagoons, fed by glaciers, are important aquatic ecosystems regulated by [...] Read more.
Most Andean lakes and lagoons are used as reservoirs to manage hydropower generation and cropland irrigation, which, in turn, alters river flow patterns through processes of storage and discharge. The Carhuacocha and Vichecocha lagoons, fed by glaciers, are important aquatic ecosystems regulated by dams. These dams increase the flow of the Mantaro River during the dry season, supporting both energy production and irrigation for croplands. Water quality in the Carhuacocha and Vichecocha lagoons was assessed between storage and discharge events by using the Canadian Council of Ministers of the Environment Water Quality Index (CCME WQI) and multivariate statistical methods. The quality of both lagoons is excellent during the storage period; however, it decreases when they are discharged during the dry season. The most sensitive parameters are pH, dissolved oxygen (DO), and biochemical oxygen demand (BOD). This paper details the changes in water quality in the Carhuacocha and Vichecocha lagoons during storage and discharge events. Full article
(This article belongs to the Section Water Quality and Contamination)
Show Figures

Figure 1

Figure 1
<p>Locations where water samples were taken from Carhuacocha and Vichecocha lagoons.</p>
Full article ">Figure 2
<p>Lake shorelines delineated from PlanetScope from period January 2023–February 2024 for Carhuacocha and Vichecocha lakes.</p>
Full article ">Figure 3
<p>Pearson’s correlation coefficients between physical–chemical parameters and metals in the water samples.</p>
Full article ">Figure 4
<p>Principal component analysis (PCA) for 27 water parameters for storage and discharge events in Carhuacocha and Vichecocha lagoons.</p>
Full article ">
Back to TopTop