[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (165)

Search Parameters:
Keywords = runoff simulation uncertainty

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
29 pages, 11100 KiB  
Article
Assessing the Impact of Rainfall Inputs on Short-Term Flood Simulation with Cell2Flood: A Case Study of the Waryong Reservoir Basin
by Hyunjun Kim, Dae-Sik Kim, Won-Ho Nam and Min-Won Jang
Hydrology 2024, 11(10), 162; https://doi.org/10.3390/hydrology11100162 - 2 Oct 2024
Viewed by 516
Abstract
This study explored the impacts of various rainfall input types on short-term runoff simulations using the Cell2Flood model in the Waryong Reservoir Basin, South Korea. Six types of rainfall data were assessed: on-site gauge measurements, spatially interpolated data from 39 Automated Synoptic Observing [...] Read more.
This study explored the impacts of various rainfall input types on short-term runoff simulations using the Cell2Flood model in the Waryong Reservoir Basin, South Korea. Six types of rainfall data were assessed: on-site gauge measurements, spatially interpolated data from 39 Automated Synoptic Observing System (ASOS) and 117 Automatic Weather System (AWS) stations using inverse distance weighting (IDW), and Hybrid Surface Rainfall (HSR) data from the Korea Meteorological Administration. The choice of rainfall input significantly affected model accuracy across the three rainfall events. The point-gauged ASOS (P-ASOS) data demonstrated the highest reliability in capturing the observed rainfall patterns, with Pearson’s r values of up to 0.84, whereas the radar-derived HSR data had the lowest correlations (Pearson’s r below 0.2), highlighting substantial discrepancies. For runoff simulation, the P-ASOS and ASOS-AWS combined interpolated dataset (R-AWS) achieved relatively accurate predictions, with P-ASOS and R-AWS exhibiting Normalized Peak Error (NPE) values of approximately 0.03 and Peak Time Error (PTE) within 20 min. In contrast, the HSR data produced large errors, with NPE up to 4.66 and PTE deviations exceeding 200 min, indicating poor temporal accuracy. Although input-specific calibration improved performance, significant errors persisted because of the inherent uncertainty of rainfall data. These findings underscore the importance of selecting and calibrating appropriate rainfall inputs to enhance the reliability of short-term flood modeling, particularly in ungauged and data-sparse basins. Full article
Show Figures

Figure 1

Figure 1
<p>Study site located in southern Gyeongsangnam-do, with observation spots with one radar flow level gauge and one AWS.</p>
Full article ">Figure 2
<p>Screen interface of Cell2Flood model showing its graphical user interface and auto-calibration module.</p>
Full article ">Figure 3
<p>Spatial input data for Cell2Flood, which were converted into an ASCII file with 500 m spatial resolution as the input data for Cell2Flood.</p>
Full article ">Figure 4
<p>Geographical distribution of rain gauge (ASOS and AWS) stations around the study site.</p>
Full article ">Figure 5
<p>Comparison of accumulated rainfall for each rainfall input by events.</p>
Full article ">Figure 6
<p>Change in accumulated rainfall in 1 h intervals of each event by rainfall inputs.</p>
Full article ">Figure 7
<p>Comparison of 1 h maximum rainfall intensity for each rainfall input by events.</p>
Full article ">Figure 8
<p>Comparison of the spatial distribution of 10-min accumulated rainfall for each rainfall input across different events. Event (<b>I</b>) 2023-06-27 22:30, (<b>II</b>) 2023-08-10 07:10, and (<b>III</b>) 2024-06-29 19:00, respectively, and (<b>a</b>–<b>c</b>) corresponding to R-ASOS, R-AWS, R-RADAR.</p>
Full article ">Figure 9
<p>Results of rainfall–runoff simulations for Event I by rainfall inputs.</p>
Full article ">Figure 9 Cont.
<p>Results of rainfall–runoff simulations for Event I by rainfall inputs.</p>
Full article ">Figure 10
<p>Results of rainfall–runoff simulations for Event II by rainfall inputs.</p>
Full article ">Figure 10 Cont.
<p>Results of rainfall–runoff simulations for Event II by rainfall inputs.</p>
Full article ">Figure 11
<p>Results of rainfall–runoff simulations for Event III by rainfall inputs.</p>
Full article ">Figure 11 Cont.
<p>Results of rainfall–runoff simulations for Event III by rainfall inputs.</p>
Full article ">Figure 12
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for P-ASOS.</p>
Full article ">Figure 13
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for P-AWS.</p>
Full article ">Figure 14
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-ASOS.</p>
Full article ">Figure 15
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-AWS.</p>
Full article ">Figure 16
<p>Results of rainfall–runoff simulations with the input-specific calibrated parameters for R-RADAR.</p>
Full article ">
23 pages, 16666 KiB  
Review
Requirements for the Development and Operation of a Freeze-Up Ice-Jam Flood Forecasting System
by Karl-Erich Lindenschmidt, Robert Briggs, Amir Ali Khan and Thomas Puestow
Water 2024, 16(18), 2648; https://doi.org/10.3390/w16182648 - 18 Sep 2024
Viewed by 450
Abstract
This article provides a comprehensive overview of ice-jam flood forecasting methodologies applicable to rivers during freezing. It emphasizes the importance of understanding river ice processes and fluvial geomorphology for developing a freeze-up ice-jam flood forecasting system. The article showcases a stochastic modelling approach, [...] Read more.
This article provides a comprehensive overview of ice-jam flood forecasting methodologies applicable to rivers during freezing. It emphasizes the importance of understanding river ice processes and fluvial geomorphology for developing a freeze-up ice-jam flood forecasting system. The article showcases a stochastic modelling approach, which involves simulating a deterministic river ice model multiple times with varying parameters and boundary conditions. This approach has been applied to the Exploits River at Badger in Newfoundland, Canada, a river that has experienced several freeze-up ice-jam floods. The forecasting involves two approaches: predicting the extent of the ice cover during river freezing and using an ensemble method to determine backwater flood level elevations. Other examples of current ice-jam flood forecasting systems for the Kokemäenjoki River (Pori, Finland), Saint John River (Edmundston, NB, Canada), and Churchill River (Mud Lake, NL, Canada) that are operational are also presented. The text provides a detailed explanation of the processes involved in river freeze-up and ice-jam formation, as well as the methodologies used for freeze-up ice-jam flood forecasting. Ice-jam flood forecasting systems used for freeze-up were compared to those employed for spring breakup. Spring breakup and freeze-up ice-jam flood forecasting systems differ in their driving factors and methodologies. Spring breakup, driven by snowmelt runoff, typically relies on deterministic and probabilistic approaches to predict peak flows. Freeze-up, driven by cold temperatures, focuses on the complex interactions between atmospheric conditions, river flow, and ice dynamics. Both systems require air temperature forecasts, but snowpack data are more crucial for spring breakup forecasting. To account for uncertainty, both approaches may employ ensemble forecasting techniques, generating multiple forecasts using slightly different initial conditions or model parameters. The objective of this review is to provide an overview of the current state-of-the-art in ice-jam flood forecasting systems and to identify gaps and areas for improvement in existing ice-jam flood forecasting approaches, with a focus on enhancing their accuracy, reliability, and decision-making potential. In conclusion, an effective freeze-up ice-jam flood forecasting system requires real-time data collection and analysis, historical data analysis, ice jam modeling, user interface design, alert systems, and integration with other relevant systems. This combination allows operators to better understand ice jam behavior and make informed decisions about potential risks or mitigation measures to protect people and property along rivers. The key findings of this review are as follows: (i) Ice-jam flood forecasting systems are often based on simple, empirical models that rely heavily on historical data and limited real-time monitoring information. (ii) There is a need for more sophisticated modeling techniques that can better capture the complex interactions between ice cover, water levels, and channel geometry. (iii) Combining data from multiple sources such as satellite imagery, ground-based sensors, numerical models, and machine learning algorithms can significantly improve the accuracy and reliability of ice-jam flood forecasts. (iv) Effective decision-support tools are crucial for integrating ice-jam flood forecasts into emergency response and mitigation strategies. Full article
Show Figures

Figure 1

Figure 1
<p>Longitudinal profiles of water temperature, ice cover width and extent and water levels along the Kokemäenjoki River in Finland (from [<a href="#B19-water-16-02648" class="html-bibr">19</a>]; used with permission).</p>
Full article ">Figure 2
<p>An ensemble of backwater level profiles from the stochastic model framework [<a href="#B21-water-16-02648" class="html-bibr">21</a>].</p>
Full article ">Figure 3
<p>Ice affected water level hydrograph forecasted on 5 December 2023, an example for the location at gauge 03PC001 (Churchill River at English Point). The 1:20 AEP, 1:100 AEP, 2012 highwater mark, and 2017 highwater mark are all ice-affected flood reference levels. Left bank, right bank, and low bank points are also included for reference. The observed water level elevations fluctuate within the 10th and 90th percentile bounds of the forecasted ice-jam flood forecasted elevations.</p>
Full article ">Figure 4
<p>Processes of ice cover formation during river freezing (adapted from [<a href="#B25-water-16-02648" class="html-bibr">25</a>]; used with permission under ASCE license 1120756-2).</p>
Full article ">Figure 5
<p>Monte Carlo analysis framework setup and calibration.</p>
Full article ">Figure 6
<p>Monte Carlo analysis framework for ice-jam flood forecasting.</p>
Full article ">Figure 7
<p>Upper and lower subbasins of the Exploits River catchment with the reach of the lower subbasin.</p>
Full article ">Figure 8
<p>Average daily mean (within a band of ±standard deviations) flows from the Millertown Dam and at Badger (from 2010 to 2019).</p>
Full article ">Figure 9
<p>Longitudinal profile of the Exploits River thalweg between the Millertown and Goodyear dams for an open-water calibration (adapted from [<a href="#B7-water-16-02648" class="html-bibr">7</a>]; solid-line box—inset bounds for Figure 11; dashed-line box—inset bounds for Figures 13 and 14.</p>
Full article ">Figure 10
<p>Badger Rough Waters along the Exploits River in July 2023 (photo by Dwanda Newman).</p>
Full article ">Figure 11
<p>Longitudinal profile of the consolidated ice cover simulated upstream from Charlie Edwards Point to Badger.</p>
Full article ">Figure 12
<p>(<b>Top panel</b>): Sentinel-1 image acquired 2 February 2019 (data from European Space Agency); (<b>bottom panel</b>): C-Core’s ice classification of the top image (from <a href="https://www.exploitsriver.app/" target="_blank">https://www.exploitsriver.app/</a>; accessed on 10 September 2024).</p>
Full article ">Figure 13
<p>(<b>Top panel</b>): distribution of the boundary conditions of discharge <span class="html-italic">Q</span>, volume of ice <span class="html-italic">V<sub>ice</sub></span>, downstream water level elevations <span class="html-italic">W,</span> and ice-jam toe location <span class="html-italic">x</span>; (<b>bottom panel</b>): ensemble of longitudinal profiles of backwater level elevations, whose distribution is summarized using percentile profiles corresponding to 1:100, 1:200, and 1:500 AEP.</p>
Full article ">Figure 14
<p>(<b>Top panel</b>): constrained distributions of flow <span class="html-italic">Q</span>, volume of ice <span class="html-italic">V<sub>ice</sub></span>, and downstream water level elevations <span class="html-italic">W</span>; (<b>bottom panel</b>): ensemble of longitudinal profiles of backwater level elevations summarized using 25th, 50th, and 75th percentile profiles.</p>
Full article ">Figure 15
<p>Structure of the framework to forecast ice cover extent and backwater levels.</p>
Full article ">Figure 16
<p>Water level elevations attained at Badger as a function of the volume of ice produced upstream of Badger during freezing (data from Fenco [<a href="#B26-water-16-02648" class="html-bibr">26</a>,<a href="#B27-water-16-02648" class="html-bibr">27</a>]).</p>
Full article ">
19 pages, 2714 KiB  
Article
Evaluating the Effects of Parameter Uncertainty on River Water Quality Predictions
by André Fonseca, Cidália Botelho, Rui A. R. Boaventura and Vítor J. P. Vilar
Resources 2024, 13(8), 106; https://doi.org/10.3390/resources13080106 - 26 Jul 2024
Viewed by 882
Abstract
Due to the high uncertainty of model predictions, it is often challenging to draw definitive conclusions when evaluating river water quality in the context of management options. The major aim of this study is to present a statistical evaluation of the Hydrologic Simulation [...] Read more.
Due to the high uncertainty of model predictions, it is often challenging to draw definitive conclusions when evaluating river water quality in the context of management options. The major aim of this study is to present a statistical evaluation of the Hydrologic Simulation Program FORTRAN (HSPF), which is a water quality modeling system, and how this modeling system can be used as a valuable tool to enhance monitoring planning and reduce uncertainty in water quality predictions. The authors’ findings regarding the sensitivity analysis of the HSPF model in relation to water quality predictions are presented. The application of the computer model was focused on the Ave River watershed in Portugal. Calibration of the hydrology was performed at two stations over five years, starting from January 1990 and ending in December 1994. Following the calibration, the hydrology model was then validated for another five-year period, from January 1995 to December 1999. A comprehensive evaluation framework is proposed, which includes a two-step statistical evaluation based on commonly used hydrology criteria for model calibration and validation. To thoroughly assess model uncertainty and parameter sensitivity, a Monte Carlo method uncertainty evaluation approach is integrated, along with multi-parametric sensitivity analyses. The Monte Carlo simulation considers the probability distributions of fourteen HSPF water quality parameters, which are used as input factors. The parameters that had the greatest impact on the simulated in-stream fecal coliform concentrations were those that represented the first-order decay rate and the surface runoff mechanism, which effectively removed 90 percent of the fecal coliform from the pervious land surface. These parameters had a more significant influence compared to the accumulation and maximum storage rates. When it comes to the oxygen governing process, the parameters that showed the highest sensitivity were benthal oxygen demand and nitrification/denitrification rate. The insights that can be derived from this study play a critical role in the development of robust water management strategies, and their significance lies in their potential to contribute to the advancement of predictive models in the field of water resources. Full article
Show Figures

Figure 1

Figure 1
<p>Watershed segmentation of Ave River Basin. Location of hydrometric stations (black dots) and meteorological stations (green dots).</p>
Full article ">Figure 2
<p>Uncertainty band (95% confidence interval) for water quality constituents at Ave River station; (<b>a</b>) fecal coliforms; (<b>b</b>) NO<sub>3</sub>; (<b>c</b>) PO<sub>4</sub>; (<b>d</b>) BOD<sub>5</sub>; and (<b>e</b>) DO.</p>
Full article ">Figure 3
<p>Uncertainty band (95% confidence interval) for water quality constituents at Este River station; (<b>a</b>) fecal coliforms; (<b>b</b>) NO<sub>3</sub>; (<b>c</b>) PO<sub>4</sub>; (<b>d</b>) BOD<sub>5</sub>, and (<b>e</b>) DO.</p>
Full article ">Figure 4
<p>Results of the multi-parametric sensitivity analysis (MPSA) for fecal coliform concentration; behavioral (black line) non-behavioral (dashed line).</p>
Full article ">Figure 5
<p>Results of multi-parametric sensitivity analysis (MPSA) for DO, NO<sub>3</sub>, PO<sub>4</sub>, and BOD<sub>5</sub> concentration: (<b>a</b>) Este River station; (<b>b</b>) Ave River station; behavioral (black line); non-behavioral (dashed line).</p>
Full article ">
23 pages, 8137 KiB  
Article
SWAT-Driven Exploration of Runoff Dynamics in Hyper-Arid Region, Saudi Arabia: Implications for Hydrological Understanding
by Sajjad Hussain, Burhan Niyazi, Amro Mohamed Elfeki, Milad Masoud, Xiuquan Wang and Muhammad Awais
Water 2024, 16(14), 2043; https://doi.org/10.3390/w16142043 - 19 Jul 2024
Viewed by 795
Abstract
Hydrological modeling plays a vital role in water-resource management and climate-change studies in hyper-arid regions. In the present investigation, surface runoff was estimated by a Soil and Water Assessment Tool (SWAT) model for Wadi Al-Aqul, Saudi Arabia. The Sequential Uncertainty Fitting version 2 [...] Read more.
Hydrological modeling plays a vital role in water-resource management and climate-change studies in hyper-arid regions. In the present investigation, surface runoff was estimated by a Soil and Water Assessment Tool (SWAT) model for Wadi Al-Aqul, Saudi Arabia. The Sequential Uncertainty Fitting version 2 (SUFI-2) technique in SWAT-CUP was adopted for the sensitivity analysis, calibration, and validation of the SWAT model’s components. The observational runoff data were scarce and only available from 1979 to 1984; such data scarcity is a common problem in hyper-arid regions. The results show good agreement with the observed daily runoff, as indicated by a Pearson Correlation Coefficient (r) of 0.86, a regression (R2) of 0.76, and a Nash–Sutcliffe coefficient (NSE) of 0.61. Error metrics, including the Mean Absolute Error (MAE) and Root Mean Square Error (RMSE), were notably low at 0.05 and 0.58, respectively. In the daily validation, the model continued to perform well, with a correlation of 0.76 and regression of 0.58. As a new approach, fitted parameters of daily calibration were incorporated into the monthly simulation, and they demonstrated an even better performance. The correlation coefficient (regression) and Nash–Sutcliffe were found to be extremely high during the calibration period of the monthly simulation, reaching 0.97 (0.95) and 0.73, respectively; meanwhile, they reached 0.99 (0.98) and 0.63 in the validation period, respectively. The sensitivity analysis using the SUFI-2 algorithm highlighted that, in the streamflow estimation, the Curve Number (CN) was found to be the most responsive parameter, followed by Soil Bulk Density (SOL_BD). Notably, the monthly results showed a higher performance than the daily results, indicating the inherent capability of the model in regard to data aggregation and reducing the impact of random fluctuations. These findings highlight the applicability of the SWAT model in predicting runoff and its implication for climate-change studies in hyper-arid regions. Full article
Show Figures

Figure 1

Figure 1
<p>Study area map of Wadi Al-Aqul, located in Al-Madinah Al-Munawarah Province.</p>
Full article ">Figure 2
<p>Rainfall hyetograph and runoff hydrograph of observed events during study period.</p>
Full article ">Figure 3
<p>Digital Elevation Model with stream network of Wadi Al-Aqul.</p>
Full article ">Figure 4
<p>LULC classification of the Wadi Al-Aqul.</p>
Full article ">Figure 5
<p>FAO soil map of Wadi Al-Aqul.</p>
Full article ">Figure 6
<p>Slope classification of Wadi Al-Aqul.</p>
Full article ">Figure 7
<p>Methodological framework of SWAT modeling.</p>
Full article ">Figure 8
<p>Regression plot between simulated and observed runoff for daily calibration (<b>left</b>) and validation (<b>right</b>).</p>
Full article ">Figure 9
<p>Regression plot between simulated and observed runoff for monthly calibration (<b>left</b>) and validation (<b>right</b>).</p>
Full article ">Figure 10
<p>(<b>a</b>) Observed (dotted black line) and simulated (dotted red line) hydrograph for calibration and validation period of Wadi Al-Aqul on the daily time scale. (<b>b</b>) Monthly calibration/validation in lower panel, indicating observed (dotted black line) and simulated (dotted red line) hydrograph for Wadi Al-Aqul.</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>) Observed (dotted black line) and simulated (dotted red line) hydrograph for calibration and validation period of Wadi Al-Aqul on the daily time scale. (<b>b</b>) Monthly calibration/validation in lower panel, indicating observed (dotted black line) and simulated (dotted red line) hydrograph for Wadi Al-Aqul.</p>
Full article ">Figure 11
<p>Radar diagram of sensitivity analysis; <span class="html-italic">p</span>-test on left and <span class="html-italic">t</span>-test (absolute values) on right.</p>
Full article ">Figure 12
<p>Sequential Uncertainty Fitting version 2 (SUFI-2) results for the value of Nash–Sutcliffe vs. parameters value.</p>
Full article ">Figure 12 Cont.
<p>Sequential Uncertainty Fitting version 2 (SUFI-2) results for the value of Nash–Sutcliffe vs. parameters value.</p>
Full article ">Figure 12 Cont.
<p>Sequential Uncertainty Fitting version 2 (SUFI-2) results for the value of Nash–Sutcliffe vs. parameters value.</p>
Full article ">
22 pages, 11223 KiB  
Article
Potential Legacy of SWOT Mission for the Estimation of Flow–Duration Curves
by Alessio Domeneghetti, Serena Ceola, Alessio Pugliese, Simone Persiano, Irene Palazzoli, Attilio Castellarin, Alberto Marinelli and Armando Brath
Remote Sens. 2024, 16(14), 2607; https://doi.org/10.3390/rs16142607 - 17 Jul 2024
Viewed by 544
Abstract
Flow–duration curves (FDCs) provide a compact view of the historical variability of river flows, reflecting climate conditions and the main hydrologic features of river basins. The Surface Water and Ocean Topography (SWOT) satellite mission will enable the estimation of river flows globally, by [...] Read more.
Flow–duration curves (FDCs) provide a compact view of the historical variability of river flows, reflecting climate conditions and the main hydrologic features of river basins. The Surface Water and Ocean Topography (SWOT) satellite mission will enable the estimation of river flows globally, by sensing rivers wider than 100 m with a sampling recurrence from 3 to 21 days. This study investigated the lifetime mission potential for FDC estimation through the comparison between remotely-sensed and empirical FDCs. We employed the Global Runoff Data Center dataset and derived SWOT-like river flows by selecting gauging stations of rivers wider than 100 m with more than 10-year long daily river flow time series. Overall, 1200 gauged river cross-sections were examined. For each site, we created a set of 24 SWOT-simulated FDCs (i.e., based on different sampling recurrences, mean biases, and random errors) to be compared against their empirical counterparts through the Nash–Sutcliffe efficiency and the mean relative error. Our results show that climate and the sampling recurrence play a key role on the performance of SWOT-based FDCs. Tropical and temperate climates performed the best, whereas arid climates mostly revealed higher uncertainties, especially for high- and low-flows. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Methodological scheme for the estimation of flow–duration curves (FDC) from SWOT-like river flow data. (<b>a</b>–<b>d</b>) SWOT-like river flow data sampled from observed GRDC time series every <span class="html-italic">k</span> days within a 3-year moving time frame (red areas). (<b>e</b>,<b>f</b>) Computation of FDCs, showing the comparison between the observed and SWOT-based FDCs within a single (<b>e</b>) or multiple time frames (<b>f</b>).</p>
Full article ">Figure 2
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. (<b>a</b>) Values for the NSE computed from SWOT-like data derived from GRDC river flow observations without any bias or random error and sampled every 3 days (i.e., Q<sub>SWOT,0,0,3</sub>). (<b>b</b>) Difference between NSE values computed from SWOT-like data derived from GRDC river flow observations without any bias or random error and sampled every 3 and 10 days (i.e., Q<sub>SWOT,0,0,3</sub> and Q<sub>SWOT,0,0,10</sub>, respectively). Additional maps showing NSE values for the remaining SWOT-like river flow datasets are available in <a href="#app1-remotesensing-16-02607" class="html-app">Appendix A</a>.</p>
Full article ">Figure 3
<p>Average site-specific NSE values, computed from SWOT-like data derived from GRDC river flow observations without any bias or random error (i.e., Q<sub>SWOT,0,0,<span class="html-italic">k</span></sub>) and grouped by macro-climate and sampling scenario. Additional boxplots showing NSE values for the remaining SWOT-like river flow datasets are available in <a href="#app1-remotesensing-16-02607" class="html-app">Appendix A</a>.</p>
Full article ">Figure 4
<p>Average site-specific mean relative error (MRE) values, computed from SWOT-like data derived from GRDC river flow observations (<b>a</b>) without any bias or random error (i.e., Q<sub>SWOT,0,0,<span class="html-italic">k</span></sub>), (<b>b</b>) assuming a random error only (i.e., Q<sub>SWOT,0,20,<span class="html-italic">k</span></sub>), and grouped by macro-climate and sampling scenario. Additional boxplots showing MRE values for the remaining SWOT-like river flow datasets are available in <a href="#app1-remotesensing-16-02607" class="html-app">Appendix A</a>.</p>
Full article ">Figure A1
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations without any bias or random error (i.e., Q<sub>swot,0,0,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A2
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations considering a 20% random error (i.e., Q<sub>swot,0,20,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A3
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations considering a 15% negative bias and a 20% random error (i.e., Q<sub>swot,−15,20,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A4
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations considering a 15% positive bias and a 20% random error (i.e., Q<sub>swot,15,20,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A5
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations considering a 30% negative bias and a 20% random error (i.e., Q<sub>swot,−30,20,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A6
<p>Assessment of SWOT-like river flow data performance in reproducing long-term FDCs, expressed in terms of the Nash–Sutcliffe efficiency (NSE). The geographic location and the climatic characterization of the considered 1200 GRDC river gauge stations are also shown. Values for NSE computed from SWOT-like data derived from GRDC river flow observations considering a 30% positive bias and a 20% random error (i.e., Q<sub>swot,30,20,k</sub>) and sampled every (<b>a</b>) 3 days, (<b>b</b>) 5 days, (<b>c</b>) 7 days, and (<b>d</b>) 10 days.</p>
Full article ">Figure A7
<p>Average site-specific NSE values, computed from SWOT-like data derived from GRDC river flow observations and grouped by macro-climate and sampling scenario for (<b>a</b>) 20% random error (Q<sub>swot,0,20,k</sub>), (<b>b</b>) 15% negative bias and 20% random error (Q<sub>swot,−15,20,k</sub>), (<b>c</b>) 15% positive bias and 20% random error (Q<sub>swot,15,20,k</sub>), (<b>d</b>) 30% negative bias and 20% random error (Q<sub>swot,−30,20,k</sub>), and (<b>e</b>) 30% positive bias and 20% random error (Q<sub>swot,30,20,k</sub>).</p>
Full article ">Figure A8
<p>Average site-specific mean relative error (MRE) values, computed from SWOT-like data derived from GRDC river flow observations and grouped by macro-climate and sampling scenario for (<b>a</b>) 15% negative bias and 20% random error (Q<sub>swot,−15,20,k</sub>), (<b>b</b>) 15% positive bias and 20% random error (Q<sub>swot,15,20,k</sub>), (<b>c</b>) 30% negative bias and 20% random error (Q<sub>swot,−30,20,k</sub>), and (<b>d</b>) 30% positive bias and 20% random error (Q<sub>swot,30,20,k</sub>).</p>
Full article ">
20 pages, 6250 KiB  
Article
A Modified Xinanjiang Model for Quantifying Streamflow Components in a Typical Watershed in Eastern China
by Kaibin Wu, Minpeng Hu, Yu Zhang, Jia Zhou and Dingjiang Chen
Hydrology 2024, 11(7), 90; https://doi.org/10.3390/hydrology11070090 - 25 Jun 2024
Viewed by 931
Abstract
An accurate quantification of flow components and an understanding of water source dynamics are essential for effective water resource and quality management. However, the complexity of hydrological processes and the interference of intensive human activities pose significant challenges in precisely separating water discharge [...] Read more.
An accurate quantification of flow components and an understanding of water source dynamics are essential for effective water resource and quality management. However, the complexity of hydrological processes and the interference of intensive human activities pose significant challenges in precisely separating water discharge into distinct components such as surface runoff, interflow, and groundwater. The Xinanjiang (XAJ) model, a conceptual watershed hydrological model, has been developed and successfully implemented for rainfall–runoff simulations and hydrograph separations across various Chinese watersheds. While the model framework is robust, it fails to account for agricultural irrigation water withdrawals and the variations in in-stream water travel times across different hydrological regimes, introducing considerable uncertainty in simulating low-flow conditions. This study introduced modifications to the XAJ model by allowing parameter adjustments across different flow regimes and incorporating irrigation withdrawals into the runoff routing process. Utilizing a decade of hydrometeorological data (2013–2022) from the Yongan River watershed in eastern China, the modified model demonstrated improved efficiency metrics in low- and medium-flow regimes compared to the original model, with a Nash–Sutcliffe coefficient improvement from −4.43~−0.49 to 0.40~0.46, R2 from 0.21~0.36 to 0.53~0.63, and BIAS reduction from 7.60~89.08% to 2.06~12.71%. Furthermore, the modified XAJ model provided a more accurate estimation of the spatial and temporal distribution of streamflow components across sub-watersheds. The original model tended to overestimate groundwater contributions (13%) and underestimate interflow (14%), particularly in low-flow conditions. The enhanced XAJ model, thus, offers a more effective tool for identifying streamflow components, providing essential insights into hydrological processes for better management decisions. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the Yongan River watershed in China and the Zhejiang Province, including hydrological stations (BZA), the Xia’an Reservoir station, and weather stations.</p>
Full article ">Figure 2
<p>Historical trends for irrigation water consumption and effective irrigated agricultural area in the Yongan River watershed over the 2013–2022 period.</p>
Full article ">Figure 3
<p>The Xinanjiang model structure and modified framework. The red dashed box denotes a modification to the original XAJ model. FSW, MSW, and CSW represent sub-watersheds that are located far from, a middle distance from, and close to the outlet of the watershed. <span class="html-italic">L</span> represents the lag time in the stream channel.</p>
Full article ">Figure 4
<p>The parameter regionalization (four categories) and the number of each sub-watershed for the Yongan River watershed.</p>
Full article ">Figure 5
<p>Model results for daily discharge of the original (<b>a</b>–<b>h</b>) and modified (<b>i</b>–<b>p</b>) XAJ model showing observed versus modeled values under different flow regimes within calibration (<b>a</b>–<b>d</b>,<b>i</b>–<b>l</b>) and validation (<b>e</b>–<b>h</b>,<b>m</b>–<b>p</b>) periods.</p>
Full article ">Figure 6
<p>The modeled daily river discharge for the original (red dot) and modified (red line) XAJ model versus observed values (black line) during the typical flood process and the typical low-flow process.</p>
Full article ">Figure 7
<p>The (<b>a</b>) average discharge and (<b>b</b>) proportions of surface runoff, interflow, and groundwater at the outlet of the watershed (2013–2022), estimated using the modified XAJ model in the Yongan watershed. The (<b>c</b>) proportions of streamflow components in three flow regimes, and the (<b>d</b>) contribution of three flow regimes to streamflow components at the outlet of the watershed.</p>
Full article ">Figure 8
<p>Spatial distribution maps of the 10-year average proportion of (<b>a</b>) surface runoff, (<b>b</b>) interflow, and (<b>c</b>) groundwater estimated using the modified XAJ model in the Yongan River watershed. The white area was upstream of the reservoir.</p>
Full article ">Figure 9
<p>Correlation between (<b>a</b>) surface runoff proportion and impervious area, (<b>b</b>) surface runoff proportion and forest land, and (<b>c</b>) interflow proportion and slope gradient in the sub-watershed. (<b>d</b>) The proportion of agricultural land in the four categories of sub-watershed. ** denote significant correlations (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">Figure 10
<p>The proportion of surface runoff, interflow, and groundwater estimated using the original and modified XAJ model in low-, medium-, and high-flow regimes. Capital letters above bars denote significant differences (<span class="html-italic">p</span> &lt; 0.01).</p>
Full article ">
21 pages, 8536 KiB  
Article
Multi-Model Comparison in the Attribution of Runoff Variation across a Humid Region of Southern China
by Qiang Wang, Fang Yang, Xiaozhang Hu, Peng Hou, Yin Zhang, Pengjun Li and Kairong Lin
Water 2024, 16(12), 1729; https://doi.org/10.3390/w16121729 - 18 Jun 2024
Viewed by 551
Abstract
The natural hydrological cycle of basins has been significantly altered by climate change and human activities, leading to considerable uncertainties in attributing runoff. In this study, the impact of climate change and human activities on runoff of the Ganjiang River Basin was analyzed, [...] Read more.
The natural hydrological cycle of basins has been significantly altered by climate change and human activities, leading to considerable uncertainties in attributing runoff. In this study, the impact of climate change and human activities on runoff of the Ganjiang River Basin was analyzed, and a variety of models with different spatio-temporal scales and complexities were used to evaluate the influence of model choice on runoff attribution and to reduce the uncertainties. The results show the following: (1) The potential evapotranspiration in the Ganjiang River Basin showed a significant downward trend, precipitation showed a significant upward trend, runoff showed a nonsignificant upward trend, and an abrupt change was detected in 1968; (2) The three hydrological models used with different temporal scales and complexity, GR1A, ABCD, DTVGM, can simulate the natural distribution of water resources in the Ganjiang River Basin; and (3) The impact of climate change on runoff change ranges from 60.07% to 82.88%, while human activities account for approximately 17.12% to 39.93%. The results show that climate change is the main driving factor leading to runoff variation in the Ganjiang River Basin. Full article
(This article belongs to the Special Issue China Water Forum 2024)
Show Figures

Figure 1

Figure 1
<p>The location of the Ganjiang River Basin (GRB) and corresponding rain gauges, meteorological stations and streamflow gauge.</p>
Full article ">Figure 2
<p>Model structure of the ABCD model.</p>
Full article ">Figure 3
<p>The time series of hydro-climatic variables in Ganjiang River Basin ((<b>a</b>) potential evapotranspiration; (<b>b</b>) precipitation; (<b>c</b>) runoff). The black square lines are observed values, and the best-fit lines estimated by the OLS (ordinary least square) method are the blue dotted lines.</p>
Full article ">Figure 4
<p>The mutation point detection of runoff series at Waizhou station.</p>
Full article ">Figure 5
<p>Comparison between simulated runoff and observed runoff at Waizhou station during 1955–1968 by using GR1A model, for (<b>a</b>) runoff processes and (<b>b</b>) scatter curve chart.</p>
Full article ">Figure 6
<p>Comparison between simulated runoff and observed runoff and curve chart at Waizhou station during 1955–1968 by using ABCD model.</p>
Full article ">Figure 7
<p>Comparison between simulated runoff and observed runoff and curve chart at Waizhou station during 1955–1968 by using the DTVGM model.</p>
Full article ">Figure 8
<p>Relative contribution rate of runoff to climate change and human activities estimated by climate elasticity coefficient analysis.</p>
Full article ">Figure 9
<p>Comparison of annual streamflow with simulated streamflow obtained by different hydrological models during the changed period.</p>
Full article ">Figure 10
<p>The average relative contribution rates of runoff to climate change and human activities are estimated by hydrological simulation analysis.</p>
Full article ">Figure 11
<p>The contribution rates of climate change and human activities are estimated by different hydrological models.</p>
Full article ">
18 pages, 5867 KiB  
Article
Estimating Stage-Frequency Curves for Engineering Design in Small Ungauged Arctic Watersheds
by Chandler Engel, Anna Wagner, Jeremy Giovando, David Ho, Blaine Morriss and Elias Deeb
Water 2024, 16(10), 1321; https://doi.org/10.3390/w16101321 - 7 May 2024
Viewed by 797
Abstract
The design of hydraulic structures in the Arctic is complicated by shallow relief, which cause unique runoff processes that promote snow-damming and refreeze of runoff. We discuss the challenges encountered in modeling snowmelt runoff into two coastal freshwater lagoons in Utqiaġvik, Alaska. Stage-frequency [...] Read more.
The design of hydraulic structures in the Arctic is complicated by shallow relief, which cause unique runoff processes that promote snow-damming and refreeze of runoff. We discuss the challenges encountered in modeling snowmelt runoff into two coastal freshwater lagoons in Utqiaġvik, Alaska. Stage-frequency curves with quantified uncertainty were required to design two new discharge gates that would allow snowmelt runoff flows through a proposed coastal revetment. To estimate runoff hydrographs arriving at the lagoons, we modeled snowpack accumulation and ablation using SnowModel which in turn was used to force a physically-based hydraulic runoff model (HEC-RAS). Our results demonstrate the successful development of stage-frequency curves by incorporating a Monte Carlo simulation approach that quantifies the variability in runoff timing and volume. Our process highlights the complexities of Arctic hydrology by incorporating significant delays in runoff onset due to localized snow accumulation and melting processes. This methodology not only addresses the uncertainty in snow-damming and refreeze processes which affect the arrival time of snowmelt inflow peaks, but is also adaptable for application in other challenging environments where secondary runoff processes are predominant. Full article
(This article belongs to the Special Issue Cold Region Hydrology and Hydraulics)
Show Figures

Figure 1

Figure 1
<p>Utqiaġvik (latitude: 71.290556, longitude: −156.788611) location map. Middle Salt Lagoon and Tasigarook Lagoon watersheds are labeled with the names of the lagoons at their downstream ends, adjacent to the ocean.</p>
Full article ">Figure 2
<p>Schematic of method to develop stage-frequency curves for lagoons. Yellow box is input data; red boxes represent processes; blue boxes represent outputs/inputs to the processes.</p>
Full article ">Figure 3
<p>Comparison of cumulative precipitation from the NWS weather station at the Utqiaġvik Wiley Post-Will Rogers Airport (STN 700260) and ERA5-Land hourly product with snowmelt runoff at the NC gage for the period 1982-2004. Cumulative precipitation is from the start of the water year ending when runoff begins at the NC gage.</p>
Full article ">Figure 4
<p>Example iteration of MC simulation combination of bootstrapped fit volume-frequency curve combined with a randomly selected stage-volume curve transfer function to obtain a single updated stage-volume curve.</p>
Full article ">Figure 5
<p>SnowModel SWE (blue line) at the CALM location. CALM site measurements are in orange diamonds while the CRREL field campaign data are represented by red triangles where minimum and maximum SWE are illustrated with red error bars.</p>
Full article ">Figure 6
<p>Runoff hydrographs (in cubic meter per second, cms) modeled results (solid blue line) and observed (dashed black line) at NC for the period WY 1982–2004.</p>
Full article ">Figure 7
<p>Scaled snowmelt runoff discharge curve shapes for the (<b>a</b>) Middle Salt Lagoon and (<b>b</b>) Tasigarook Lagoon watersheds. Each colored line represents a different total volume of runoff.</p>
Full article ">Figure 8
<p>Minimum and maximum air temperatures (blue and red line, respectively), and precipitation (light blue bars) driving the simulated runoff hydrographs for the WY 1983, 1985, 1989, and 2014.</p>
Full article ">Figure 9
<p>Stage-volume curves for (<b>a</b>) Middle Salt Lagoon and (<b>b</b>) Tasigarook Lagoon for different inflow hydrograph shapes (WY 1983, 1985, 1989, 2014, and outlet closed for Middle Salt Lagoon). Each color (black, red, green, blue) represents a different hydrograph shape.</p>
Full article ">Figure 10
<p>Example stage-frequency curve for the (<b>a</b>) Middle Salt, and (<b>b</b>) Tasigarook Lagoons including uncertainty.</p>
Full article ">
28 pages, 8281 KiB  
Article
Development of an Integrated Urban Flood Model and Its Application in a Concave-Down Overpass Area
by Yuna Yan, Han Zhang, Na Zhang and Chuhan Feng
Remote Sens. 2024, 16(10), 1650; https://doi.org/10.3390/rs16101650 - 7 May 2024
Viewed by 1382
Abstract
Urban floods caused by extreme rainstorm events have increased in recent decades, particularly in concave-down bridge zones. To simulate urban flooding processes accurately, an integrated urban flood model (IUFM) was constructed by coupling a distributed urban surface runoff model based on the cellular [...] Read more.
Urban floods caused by extreme rainstorm events have increased in recent decades, particularly in concave-down bridge zones. To simulate urban flooding processes accurately, an integrated urban flood model (IUFM) was constructed by coupling a distributed urban surface runoff model based on the cellular automata framework (CA-DUSRM), a widely used pipe convergence module in the storm water management model (SWMM), with an inundation module that describes the overflow expansion process associated with terrain and land-cover. The IUFM was used in a case study of the Anhua Bridge (a typical concave-down overpass) study area in Beijing, China. The spatial-temporal variations in flood depth modeled by the IUFM were verified to be reliable by comparison with actual measurements and other simulations. The validated IUFM was used to obtain temporal variations in flood range, depth, and volume under four rainstorm scenarios (return periods of 3-year, 10-year, 50-year, and 100-year). The results showed that the surface runoff process, overflow from drainage networks, and overflow expansion process could affect the flooding status by changing the composition and spatial configuration of pervious or impervious patches, drainage capacity, and underlying surface characteristics (such as terrain and land-cover). Overall, although the simulation results from the IUFM contain uncertainties from the model structures and inputs, the IUFM is an effective tool that can provide accurate and timely information to prevent and control urban flood disasters and provide decision-making support for long-term storm water management and sponge city construction. Full article
Show Figures

Figure 1

Figure 1
<p>Anhua Bridge study area in Beijing and six subcatchments (sub1, sub2, sub3, sub4, sub5, and sub6).</p>
Full article ">Figure 2
<p>Digital elevation model (DEM) data (<b>a</b>) and land-cover types (<b>b</b>) in the Anhua Bridge study area, Beijing.</p>
Full article ">Figure 3
<p>Developed integrated urban flood model (IUFM).</p>
Full article ">Figure 4
<p>Conceptual model of a distributed urban surface runoff model based on the cellular automata framework (CA-DUSRM) [<a href="#B33-remotesensing-16-01650" class="html-bibr">33</a>].</p>
Full article ">Figure 5
<p>Conceptual model of overflow inundation simulation in the integrated urban flood model (IUFM). <span class="html-italic">R<sub>c</sub></span> is the expanding range in the horizontal direction from the central cell (node) outward; <span class="html-italic">c</span> is the number of cells on a side; <span class="html-italic">n</span> is the total number of calculations in the vertical direction, one calculation for one given water level during the cut-and-trial process; <span class="html-italic">H<sub>n</sub></span> is the <span class="html-italic">n</span>th flood elevation at the node (m); and Δ<span class="html-italic">h</span> is the given increment in water level for one calculation in vertical direction (m).</p>
Full article ">Figure 6
<p>Landscape pattern metrics for impervious patches in the six subcatchments (sub1, sub2, sub3, sub4, sub5, and sub6) of the Anhua Bridge study area in Beijing. PLAND, percentage of landscape; LPI, largest patch index; PLADJ, percentage of like adjacencies; COHESION, patch cohesion index.</p>
Full article ">Figure 7
<p>Rainfall and modeled flooding processes during the four actual rainfall events on 1 August 2007 (<b>a</b>), 6 August 2007 (<b>b</b>), 21 July 2012 (<b>c</b>), and 20 July 2016 (<b>d</b>) (see <a href="#remotesensing-16-01650-t002" class="html-table">Table 2</a>) over the entire Anhua Bridge study area in Beijing.</p>
Full article ">Figure 8
<p>Spatial distribution of the modeled maximum flood depth during the four actual rainfall events on 1 August 2007 (<b>a</b>), 6 August 2007 (<b>b</b>), 21 July 2012 (<b>c</b>), and 20 July 2016 (<b>d</b>) in the Anhua Bridge study area of Beijing.</p>
Full article ">Figure 9
<p>Modeled temporal variations in flood depths at three sites (A, B, and C) during the three actual rainfall events on 1 August 2007 (<b>a</b>), 6 August 2007 (<b>b</b>), and 21 July 2012 (<b>c</b>) in the Anhua Bridge study area of Beijing. The spatial locations of the three sites are shown in <a href="#remotesensing-16-01650-f008" class="html-fig">Figure 8</a>. The numbers of 1, 2, 3, …, 7 correspond to the numbers of flood depth data from measurements or other simulations in <a href="#remotesensing-16-01650-t004" class="html-table">Table 4</a>.</p>
Full article ">Figure 10
<p>Modeled rainfall (<b>a</b>), mean flood depth (<b>b</b>), total flood volume (<b>c</b>) and flood range (<b>d</b>) during the four rainstorms with different return periods over the entire Anhua Bridge study area in Beijing.</p>
Full article ">Figure 11
<p>Surface runoff coefficients (<b>a</b>) and overflow ratios (<b>b</b>) for the six subcatchments (sub1, sub2, sub3, sub4, sub5, and sub6) in the Anhua Bridge study area of Beijing. The five horizontal lines from top to bottom in the box plot represent the maximum, upper quartile, median, lower quartile, and minimum values, respectively, and “×” denotes the mean value. For each subcatchment, the samples came from the four surface runoff coefficients and four overflow ratios under the four rainstorm scenarios (return periods of 3-year, 10-year, 50-year, and 100-year).</p>
Full article ">Figure 12
<p>Comparisons in flood range and depth from the modeled overflow expansion process when not considering (<b>a</b>) and considering (<b>b</b>) the differences in land-cover types under the assumption of an input total overflow of 5 × 10<sup>4</sup> m<sup>3</sup>. The red polygons represent the subcatchments; the yellow arrows point to the enlarged image in the black circle area.</p>
Full article ">
18 pages, 2707 KiB  
Article
Operational Robustness Assessment of the Hydro-Based Hybrid Generation System under Deep Uncertainties
by Jianhua Jiang, Bo Ming, Qiang Huang and Qingjun Bai
Energies 2024, 17(8), 1974; https://doi.org/10.3390/en17081974 - 22 Apr 2024
Viewed by 673
Abstract
The renewable-dominant hybrid generation systems (HGSs) are increasingly important to the electric power system worldwide. However, influenced by uncertain meteorological factors, the operational robustness of HGSs must be evaluated to inform the associated decision-making. Additionally, the main factors affecting the HGS’s robustness should [...] Read more.
The renewable-dominant hybrid generation systems (HGSs) are increasingly important to the electric power system worldwide. However, influenced by uncertain meteorological factors, the operational robustness of HGSs must be evaluated to inform the associated decision-making. Additionally, the main factors affecting the HGS’s robustness should be urgently identified under deep uncertainties, as this provides valuable guidance for HGS capacity configuration. In this paper, a multivariate stochastic simulation method is developed and used to generate uncertain resource scenarios of runoff, photovoltaic power, and wind power. Subsequently, a long-term stochastic optimization model of the HGS is employed to derive the optimal operating rules. Finally, these operating rules are used to simulate the long-term operation of an HGS, and the results are used to evaluate the HGS’s robustness and identify its main sensitivities. A clean energy base located in the Upper Yellow River Basin, China, is selected as a case study. The results show that the HGS achieves greater operational robustness than an individual hydropower system, and the robustness becomes weaker as the total capacity of photovoltaic and wind power increases. Additionally, the operational robustness of the HGS is found to be more sensitive to the total capacity than to the capacity ratio between photovoltaic and wind power. Full article
Show Figures

Figure 1

Figure 1
<p>Sketch map of hypothetical uncertain resource scenarios.</p>
Full article ">Figure 2
<p>Locations of the clean energy bases in the Upper Yellow River Basin.</p>
Full article ">Figure 3
<p>Simulated inflow series: (<b>a</b>) Cihaxia; (<b>b</b>) Banduo; (<b>c</b>) Yangqu (−10% change scenario).</p>
Full article ">Figure 4
<p>Simulated series of PV and wind power output (−10% change scenario with a −10% inflow condition).</p>
Full article ">Figure 5
<p>Technical and economic evaluation indicators under different operation schemes in the historical baseline state.</p>
Full article ">Figure 6
<p>The HGS’s robustness under different operation schemes.</p>
Full article ">Figure 7
<p>Cumulative distribution curves of the technical and economic indicators under uncertain resource scenarios: (<b>a</b>) hydroelectricity production; (<b>b</b>) total energy production; (<b>c</b>) guaranteed rate; (<b>d</b>) spilled water; (<b>e</b>) electricity curtailment rate.</p>
Full article ">Figure 8
<p>Comparison of the robustness of the HGS under uncertain scenarios: (<b>a</b>) <span class="html-italic">S</span><sub>1<span class="html-italic">j</span></sub> under PV and inflow uncertainties; (<b>b</b>) <span class="html-italic">S</span><sub>1<span class="html-italic">j</span></sub> under wind power and inflow uncertainties; (<b>c</b>) <span class="html-italic">S</span><sub>1<span class="html-italic">j</span></sub> under PV and wind power uncertainties; (<b>d</b>) <span class="html-italic">S</span><sub>2<span class="html-italic">j</span></sub> under PV and inflow uncertainties; (<b>e</b>) <span class="html-italic">S</span><sub>2<span class="html-italic">j</span></sub> under wind and inflow uncertainties; (<b>f</b>) <span class="html-italic">S</span><sub>2<span class="html-italic">j</span></sub> under PV and wind power uncertainties.</p>
Full article ">Figure 9
<p>Robustness indicators under all installed capacity scenarios of PV and wind power plants.</p>
Full article ">Figure 10
<p>Robustness indicators under all capacity ratio scenarios of PV and wind power plants.</p>
Full article ">
15 pages, 2435 KiB  
Article
Optimizing Parameters in the Common Land Model by Using Gravity Recovery and Climate Experiment Satellite Observations
by Yuan Su and Shupeng Zhang
Land 2024, 13(4), 508; https://doi.org/10.3390/land13040508 - 12 Apr 2024
Viewed by 970
Abstract
Terrestrial water storage (TWS) is pivotal in understanding environmental dynamics, climate change, and human impacts. Despite the utility of land surface models, uncertainties persist in their parameterization schemes. This study employs GRACE (Gravity Recovery and Climate Experiment) satellite data to optimize the runoff [...] Read more.
Terrestrial water storage (TWS) is pivotal in understanding environmental dynamics, climate change, and human impacts. Despite the utility of land surface models, uncertainties persist in their parameterization schemes. This study employs GRACE (Gravity Recovery and Climate Experiment) satellite data to optimize the runoff parameterization scheme within the Common Land Model by a data assimilation and parameter optimization method. The optimization algorithm sets an adjustment factor that varies with time and space for runoff simulation and updates it along with the running of the land surface model. The evaluation reveals that there are improved correlation coefficients and reduced root mean square errors compared to GRACE observations. Independent assessments by using in situ river discharge observations demonstrate enhanced model performance, particularly in mountainous regions such as western North America. This study underscores the efficacy of integrating GRACE data to improve land surface model parameterization, offering more accurate predictions of TWS changes. Full article
(This article belongs to the Section Land Systems and Global Change)
Show Figures

Figure 1

Figure 1
<p>Workflow of data assimilation algorithm.</p>
Full article ">Figure 2
<p>Time-averaged logarithm of adjustment factor to runoff in CoLM.</p>
Full article ">Figure 3
<p>Annual mean runoff simulated by CoLM without data assimilation.</p>
Full article ">Figure 4
<p>Difference in annual mean runoff between simulations with and without assimilating GRACE data into CoLM.</p>
Full article ">Figure 5
<p>Correlation coefficients between time series of terrestrial water storage changes simulated by CoLM and those observed by GRACE satellite.</p>
Full article ">Figure 6
<p>Difference between correlation coefficients by assimilating GRACE data into CoLM and those without data assimilation.</p>
Full article ">Figure 7
<p>Root mean square errors in time series of terrestrial water storage changes simulated by CoLM relative to those observed by GRACE satellite.</p>
Full article ">Figure 8
<p>Difference in root mean square errors by assimilating GRACE data into CoLM compared to those without data assimilation.</p>
Full article ">Figure 9
<p>A total of 13 regions selected for further evaluations.</p>
Full article ">Figure 10
<p>Parameter optimization results in western North America. (<b>a</b>) Time series of terrestrial water storage changes by CoLM without data assimilation, CoLM with data assimlation, and GRACE observations. (<b>b</b>) Time series of logarithm of scaling factor and change in water table depth. (<b>c</b>) Spatial distribution of time-averaged logarithm of scaling factor. (<b>d</b>) Difference in correlation coefficients between CoLM simulations with data assimilation and those without data assimilation. (<b>e</b>) Difference in root mean square errors between CoLM simulations with data assimilation and those without data assimilation.</p>
Full article ">Figure 11
<p>Difference in Nash–Sutcliffe efficiency coefficient between coupled model with data assimilation and that without data assimilation.</p>
Full article ">
27 pages, 4897 KiB  
Article
Calibrating Agro-Hydrological Model under Grazing Activities and Its Challenges and Implications
by Amanda M. Nelson, Mahesh L. Maskey, Brian K. Northup and Daniel N. Moriasi
Hydrology 2024, 11(4), 42; https://doi.org/10.3390/hydrology11040042 - 22 Mar 2024
Cited by 1 | Viewed by 1819
Abstract
Recently, the Agricultural Policy Extender (APEX) model was enhanced with a grazing module, and the modified grazing database, APEXgraze, recommends sustainable livestock farming practices. This study developed a combinatorial deterministic approach to calibrate runoff-related parameters, assuming a normal probability distribution for each parameter. [...] Read more.
Recently, the Agricultural Policy Extender (APEX) model was enhanced with a grazing module, and the modified grazing database, APEXgraze, recommends sustainable livestock farming practices. This study developed a combinatorial deterministic approach to calibrate runoff-related parameters, assuming a normal probability distribution for each parameter. Using the calibrated APEXgraze model, the impact of grazing operations on native prairie and cropland planted with winter wheat and oats in central Oklahoma was assessed. The existing performance criteria produced four solutions with very close values for calibrating runoff at the farm outlet, exhibiting equifinality. The calibrated results showed that runoff representations had coefficients of determination and Nash–Sutcliffe efficiencies >0.6 in both watersheds, irrespective of grazing operations. Because of non-unique solutions, the key parameter settings revealed different metrics yielding different response variables. Based on the least objective function value, the behavior of watersheds under different management and grazing intensities was compared. Model simulations indicated significantly reduced water yield, deep percolation, sediment yield, phosphorus and nitrogen loadings, and plant temperature stress after imposing grazing, particularly in native prairies, as compared to croplands. Differences in response variables were attributed to the intensity of tillage and grazing activities. As expected, grazing reduced forage yields in native prairies and increased crop grain yields in cropland. The use of a combinatorial deterministic approach to calibrating parameters offers several new research benefits when developing farm management models and quantifying sensitive parameters and uncertainties that recommend optimal farm management strategies under different climate and management conditions. Full article
(This article belongs to the Special Issue Hydrological Processes in Agricultural Watersheds)
Show Figures

Figure 1

Figure 1
<p>Location of study site within Water Resources and Erosion (WRE) watersheds in El Reno, OK. Circles are the outlets with installed H-flume for runoff measurement.</p>
Full article ">Figure 2
<p>Generalized research protocol with basic workflow diagram. While “Basic parameter files” are the APEX-related input files generated from the NTT, “Parameter files” refer to the updated files using APEXgraze Editor before simulation.</p>
Full article ">Figure 3
<p>Calibration protocol for hydrological models.</p>
Full article ">Figure 4
<p>Daily time series of best representations of surface runoff (January 1983 to September 2000) optimized at daily scale for native prairie, WRE1. (<b>a</b>) Without grazing (<b>top</b>), (<b>b</b>) with grazing (<b>bottom</b>).</p>
Full article ">Figure 5
<p>Daily time series of best representations of surface runoff (January 1982 to September 2000) optimized at daily scale for cropland, WRE8. (<b>a</b>) Without grazing (<b>top</b>), (<b>b</b>) with grazing (<b>bottom</b>).</p>
Full article ">Figure 6
<p>Water balance components at a monthly scale implied by the calibrated parameter sets for WRE1 (native prairie) over 1983–2000 and WRE8 over 1982–2000 (cropland) without (<b>top</b>) and with (<b>bottom</b>) grazing operations.</p>
Full article ">Figure 7
<p>Annual average biomass generated from WRE1 (grassland) and WRE8 (cropland) under without and with grazing operations.</p>
Full article ">Figure 8
<p>Annual yields generated by WRE1 (native prairie, <b>top</b>) and WRE8 (cropland, <b>bottom</b>) without and with grazing operations. Bar chart with stars refer to crop yield while farm is under grazing activities.</p>
Full article ">Figure 9
<p>Biomass and leaf area index with and without grazing activities for the year 1983. Top (<b>a</b>,<b>c</b>): cropland (WRE8); bottom (<b>b</b>,<b>d</b>): native prairie (WRE1).</p>
Full article ">
19 pages, 5206 KiB  
Article
Many-Objective Hierarchical Pre-Release Flood Operation Rule Considering Forecast Uncertainty
by Yongqi Liu, Guibing Hou, Baohua Wang, Yang Xu, Rui Tian, Tao Wang and Hui Qin
Water 2024, 16(5), 785; https://doi.org/10.3390/w16050785 - 6 Mar 2024
Cited by 1 | Viewed by 1016
Abstract
Flood control operation of cascade reservoirs is an important technology to reduce flood disasters and increase economic benefits. Flood forecast information can help reservoir managers make better use of flood resources and reduce flood risks. In this paper, a hierarchical pre-release flood operation [...] Read more.
Flood control operation of cascade reservoirs is an important technology to reduce flood disasters and increase economic benefits. Flood forecast information can help reservoir managers make better use of flood resources and reduce flood risks. In this paper, a hierarchical pre-release flood operation rule considering the flood forecast and its uncertainty information is proposed for real-time flood control. A many-objective optimization model considering the cascade reservoir’s power generation objective, flood control objective, and navigation objective is established. Then, a region search evolutionary algorithm is applied to optimize the many-objective optimization model in a real-world case study upstream of the Yangtze River basin. The optimization experimental results show that the region search evolutionary algorithm can balance convergence and diversity well, and the HV value is 40% higher than the MOEA/D algorithm. The simulation flood control results of cascade reservoirs upstream of the Yangtze River demonstrate that the optimized flood control rule can increase the average multi-year power generation of cascade reservoirs by a maximum of 27.72 × 108 kWh under the condition of flood control safety. The rules proposed in this paper utilize flood resources by identifying runoff forecast information, and pre-release to the flood limit level 145 m before the big flood occurs, so as to ensure the safety downstream and the dam’s own flood control and provide reliable decision support for reservoir managers. Full article
Show Figures

Figure 1

Figure 1
<p>A flowchart of the many-objective hierarchical pre-release flood operation rule considering forecast uncertainty.</p>
Full article ">Figure 2
<p>The process of the two-stage reservoir dispatch model.</p>
Full article ">Figure 3
<p>The impact of forecast uncertainty on reservoir decision.</p>
Full article ">Figure 4
<p>The representation of the HPFOR.</p>
Full article ">Figure 5
<p>Location of Xiluodu, Xiangjiaba, and Three Gorges in the Yangtze River basin.</p>
Full article ">Figure 6
<p>Radar figures of the final solution set obtained by MOEA/D, NSGAIII, <span class="html-italic">θ</span>-DEA, and RSEA (F1: power generation objective, F2: reservoir safety objective, F3: flood control station security objective, F4: upstream cascade reservoirs flood control objective, F5: navigation objective). (<b>a</b>) MOEA/D, (<b>b</b>) NSGA-III, (<b>c</b>) <span class="html-italic">θ</span>-DEA, and (<b>d</b>) RSEA.</p>
Full article ">Figure 7
<p>Release processes and water level processes of Three Gorges in 1981: (<b>a</b>) scheme 1, (<b>b</b>) scheme 25, (<b>c</b>) scheme 57, and (<b>d</b>) scheme 85.</p>
Full article ">Figure 8
<p>Release processes and water level processes of Three Gorges in 1998: (<b>a</b>) scheme 1, (<b>b</b>) scheme 25, (<b>c</b>) scheme 57, and (<b>d</b>) scheme 85.</p>
Full article ">Figure 9
<p>Water level processes of Three Gorges in 1981 and 1998: (<b>a</b>) 1981 and (<b>b</b>) 1998.</p>
Full article ">
14 pages, 3532 KiB  
Article
Quantifying Uncertainty in Runoff Simulation According to Multiple Evaluation Metrics and Varying Calibration Data Length
by Ghaith Falah Ziarh, Jin Hyuck Kim, Jae Yeol Song and Eun-Sung Chung
Water 2024, 16(4), 517; https://doi.org/10.3390/w16040517 - 6 Feb 2024
Cited by 1 | Viewed by 1175
Abstract
In this study, the uncertainty in runoff simulations using hydrological models was quantified based on the selection of five evaluation metrics and calibration data length. The calibration data length was considered to vary from 1 to 11 years, and runoff analysis was performed [...] Read more.
In this study, the uncertainty in runoff simulations using hydrological models was quantified based on the selection of five evaluation metrics and calibration data length. The calibration data length was considered to vary from 1 to 11 years, and runoff analysis was performed using a soil and water assessment tool (SWAT). SWAT parameter optimization was then performed using R-SWAT. The results show that the uncertainty was lower when using a calibration data length of five to seven years, with seven years achieving the lowest uncertainty. Runoff simulations using a calibration data length of more than seven years yielded higher uncertainty overall but lower uncertainty for extreme runoff simulations compared to parameters with less than five years of calibration data. Different uncertainty evaluation metrics show different levels of uncertainty, which means it is necessary to consider multiple evaluation metrics rather than relying on any one single metric. Among the evaluation metrics, the Nash–Sutcliffe model efficiency coefficient (NSE) and normalized root-mean-squared error (NRMSE) had large uncertainties at short calibration data lengths, whereas the Kling–Gupta efficiency (KGE) and Percent Bias (Pbias) had large uncertainties at long calibration data lengths. Full article
(This article belongs to the Topic Hydrology and Water Resources Management)
Show Figures

Figure 1

Figure 1
<p>Study workflow.</p>
Full article ">Figure 2
<p>Study area.</p>
Full article ">Figure 3
<p>Description of the calibration data lengths and validation periods (The blue arrow means calibration data length).</p>
Full article ">Figure 4
<p>Variation in model performance (NSE) before (<b>a</b>) and after (<b>b</b>) parameter optimization. Numbers on the horizontal axis refer to the length of the calibration data used for parameter optimization.</p>
Full article ">Figure 5
<p>Box-plots of evaluation metrics for the validation period according to different calibration data lengths.</p>
Full article ">Figure 6
<p>Box-plot of uncertainty index for validation period.</p>
Full article ">Figure 7
<p>Difference between the simulated and the observed extreme runoff using JS-D.</p>
Full article ">Figure 8
<p>Matrix chart of overall uncertainty ranks.</p>
Full article ">Figure 9
<p>Overall uncertainty for long and short calibration data lengths together: (<b>a</b>)—NSE, (<b>b</b>)—KGE, (<b>c</b>)—Pbias, (<b>d</b>)—RMSE, (<b>e</b>)—uncertainty index, and (<b>f</b>)—JSD.</p>
Full article ">
30 pages, 49910 KiB  
Article
Climate Change Impacts on Nitrate Leaching and Groundwater Nitrate Dynamics Using a Holistic Approach and Med-CORDEX Climatic Models
by Aikaterini Lyra, Athanasios Loukas, Pantelis Sidiropoulos and Lampros Vasiliades
Water 2024, 16(3), 465; https://doi.org/10.3390/w16030465 - 31 Jan 2024
Cited by 1 | Viewed by 1431
Abstract
This study presents the projected future evolution of water resource balance and nitrate pollution under various climate change scenarios and climatic models using a holistic approach. The study area is Almyros Basin and its aquifer system, located in Central Greece, Thessaly, Greece. Almyros [...] Read more.
This study presents the projected future evolution of water resource balance and nitrate pollution under various climate change scenarios and climatic models using a holistic approach. The study area is Almyros Basin and its aquifer system, located in Central Greece, Thessaly, Greece. Almyros Basin is a coastal agricultural basin and faces the exacerbation of water deficit and groundwater nitrate pollution. Using an Integrated Modeling System (IMS), which consists of the surface hydrology model (UTHBAL), the nitrate leachate model (REPIC, an R-ArcGIS-based EPIC model), the groundwater hydrology model (MODFLOW), and the nitrates’ advection, dispersion, and transport model (MT3MDS), the projected values of the variables of water quantity and quality are simulated. Nineteen climatic models from the Med-CORDEX database were bias-corrected with the Quantile Empirical Mapping method and employed to capture the variability in the simulated surface and groundwater water balance and nitrate dynamics. The findings indicate that future precipitation, runoff, and groundwater recharge will decrease while temperature and potential evapotranspiration will increase. Climate change will lead to reduced nitrogen leaching, lower groundwater levels, and persistent nitrate pollution; however, it will be accompanied by high variability and uncertainty, as simulations of IMS under multiple climatic models indicate. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the Almyros Basin, its watersheds, and the Almyros aquifer system.</p>
Full article ">Figure 2
<p>(<b>a</b>) Map of the Almyros aquifer hydrogeological formations at different elevated horizontal sections. (<b>b</b>) Map of the averaged hydraulic heads as simulated using the Integrated Modeling System and the averaged observed nitrate concentrations at well locations for the observed period 1991–2018.</p>
Full article ">Figure 3
<p>Box–Whisker plots of Mean Annual Precipitation (mm) for the observed periods 1971–2000 and 2001–2018, and the Bias–corrected Med–CORDEX Models under RCP8.5 for the periods 1971–2000 and 2001–2018, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 4
<p>Box–Whisker plots of Mean Annual Precipitation (mm) for the observed periods 1971–2000 and 2001–2018, and the Bias–corrected Med–CORDEX Models under RCP4.5 for the periods 1971–2000 and 2001–2018, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 5
<p>Box–Whisker plots of Mean Annual Temperature in °C for the observed periods 1971–2000 and 2001–2018, and the Bias–corrected Med–CORDEX Models under RCP8.5 for the periods 1971–2000 and 2001–2018, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 6
<p>Box–Whisker plots of Mean Annual Temperature in °C for the observed periods 1971–2000 and 2001–2018, and the Bias–corrected Med–CORDEX Models under RCP4.5 for the periods 1971–2000 and 2001–2018, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 7
<p>Bar charts for the Simulated variables of Potential Evapotranspiration (PET), Actual Evapotranspiration (AET), Surface runoff (Q), and Groundwater Recharge (Rg) in mm for the Bias–corrected Med–CORDEX Models under RCP8.5 for the historical time periods of 1971–2000 and 2001–2018, and the future periods 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 8
<p>Bar charts for the simulated variables of Potential Evapotranspiration (PET), Actual Evapotranspiration (AET), Surface runoff (Q), and Groundwater Recharge (Rg) in mm for the Bias–corrected Med–CORDEX Models under RCP4.5, for the historical time periods 1971–2000 and 2001–2018, and the future periods 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 9
<p>Box–Whisker of Nitrogen Leaching (kg/ha/yr) as simulated with the REPIC model for the Bias–corrected Med–CORDEX Models under RCP8.5, for historical periods 1971–2000 and 2001–2018, and future periods, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 10
<p>Box–Whisker of Nitrogen Leaching (kg/ha/yr) as simulated with the REPIC model for the Bias–corrected Med–CORDEX Models under RCP4.5, for historical periods 1971–2000 and 2001–2018, and future periods, 2019–2050, 2051–2080 and 2081–2100.</p>
Full article ">Figure 11
<p>Section AA’ (<b>left</b>) and Section BB’ (<b>right</b>) for Nitrate Concentrations in mg/L for the Bias–corrected Med–CORDEX Models under RCP8.5 for the time periods of 1991–2018, 2019–2050, 2051–2080, and 2081–2100.</p>
Full article ">Figure 12
<p>Section AA’ (<b>left</b>) and Section BB’ (<b>right</b>) for Nitrate Concentrations in mg/L for the Bias–corrected Med–CORDEX Models under RCP4.5 for the time periods of 1991–2018, 2019–2050, 2051–2080, and 2081–2100.</p>
Full article ">
Back to TopTop