[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

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (1,399)

Search Parameters:
Keywords = inland waters

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
23 pages, 9177 KiB  
Article
Shallow Water Depth Estimation of Inland Wetlands Using Landsat 8 Satellite Images
by Collins Owusu, Nicholas M. Masto, Alfred J. Kalyanapu, Justin N. Murdock and Bradley S. Cohen
Remote Sens. 2024, 16(16), 2986; https://doi.org/10.3390/rs16162986 - 14 Aug 2024
Abstract
Water depth affects many aspects of wetland ecology, hydrology, and biogeochemistry. However, acquiring water depth data is often difficult due to inadequate monitoring or insufficient funds. Satellite-derived bathymetry (SBD) data provides cost-effective and rapid estimates of the water depth across large areas. However, [...] Read more.
Water depth affects many aspects of wetland ecology, hydrology, and biogeochemistry. However, acquiring water depth data is often difficult due to inadequate monitoring or insufficient funds. Satellite-derived bathymetry (SBD) data provides cost-effective and rapid estimates of the water depth across large areas. However, the applicability and performance of these techniques for inland wetlands have not been thoroughly evaluated. Here, a time series of bathymetry data for inland wetlands in West Kentucky and Tennessee were derived from Landsat 8 images using two widely used empirical models, Stumpf and a modified Lyzenga model and three machine learning models, Random Forest, Support Vector regression, and k-Nearest Neighbor. We processed satellite images using Google Earth Engine and compared the performance of water depth estimation among the different models. The performance assessment at validation sites resulted in an RMSE in the range of 0.18–0.47 m and R2 in the range of 0.71–0.83 across all models for depths < 3.5 m, while in depths > 3.5 m, an RMSE = 1.43–1.78 m and R2 = 0.57–0.65 was obtained. Overall, the empirical models marginally outperformed the machine learning models, although statistical tests indicated the results from all the models were not significantly different. Testing of the models beyond the domain of the training and validation data suggested the potential for model transferability to other regions with similar hydrologic and environmental characteristics. Full article
29 pages, 20003 KiB  
Article
Chromaticity-Based Discrimination of Algal Bloom from Inland and Coastal Waters Using In Situ Hyperspectral Remote Sensing Reflectance
by Dongzhi Zhao, Qinshun Luo and Zhongfeng Qiu
Water 2024, 16(16), 2276; https://doi.org/10.3390/w16162276 - 13 Aug 2024
Viewed by 347
Abstract
The rapid growth of phytoplankton and microalgae has presented considerable environmental and societal challenges to the sustainable development of human society. Given the inherent limitations of satellite-based algal bloom detection techniques that rely on chlorophyll and fluorescence methods, this study proposes a method [...] Read more.
The rapid growth of phytoplankton and microalgae has presented considerable environmental and societal challenges to the sustainable development of human society. Given the inherent limitations of satellite-based algal bloom detection techniques that rely on chlorophyll and fluorescence methods, this study proposes a method that employs hyperspectral data to calculate water chromatic indices (WCIs), including hue, saturation (S), dominant wavelength (λd), and integrated apparent visual wavelength (IAVW), to identify algal blooms. A global in situ hyperspectral dataset was constructed, comprising 13,110 entries, of which 9595 were for normal waters and 3515 for algal bloom waters. The findings of our investigation indicate statistically significant discrepancies in chromaticity parameters between normal and algal bloom waters, with a p-value of 0.05. It has been demonstrated that different algal blooms exhibit distinct chromatic characteristics. For algae of the same type, the chromaticity parameters increase exponentially with chlorophyll concentration for hue and λd, while S shows low correlation and IAVW displays a good linear relationship with chlorophyll concentration. The application of this method to the Bohai Sea (coastal) and Taihu Lake (inland water) for the extraction of algal blooms revealed a clear separation in chromaticity parameters between normal and algal bloom waters. Moreover, the method can be applied to satellite data, offering an alternative approach for the detection of algal blooms based on satellite data. The indices can serve as ground truth values for colorimetric indices and provide a benchmark for the validation of satellite chromatic products. Full article
(This article belongs to the Section Biodiversity and Functionality of Aquatic Ecosystems)
Show Figures

Figure 1

Figure 1
<p>Schematic diagram of chromatic coordinate.</p>
Full article ">Figure 2
<p>The global distribution map of spectral libraries for normal water and algal bloom water bodies. (The red dots are algal blooms water, and the blue dots are normal water). ((<b>a</b>) The global scale, (<b>b</b>) the Bohai scale).</p>
Full article ">Figure 3
<p>The library of spectral data for algal blooms and normal water bodies: ((<b>a</b>) normal water bodies (<b>b</b>) algae water bodies).</p>
Full article ">Figure 4
<p>The scatter plot of the XYZ<sub>390–830 nm</sub>, XYZ<sub>400–830 nm</sub>, and XYZ<sub>360–830 nm</sub>. ((<b>a</b>–<b>c</b>) is the adjustment scatter plot between XYZ<sub>390–830 nm</sub> and XYZ<sub>360–830 nm</sub>, (<b>d</b>–<b>f</b>) is the adjustment scatter plot between XYZ<sub>400–830 nm</sub> and XYZ<sub>360–830 nm</sub>).</p>
Full article ">Figure 4 Cont.
<p>The scatter plot of the XYZ<sub>390–830 nm</sub>, XYZ<sub>400–830 nm</sub>, and XYZ<sub>360–830 nm</sub>. ((<b>a</b>–<b>c</b>) is the adjustment scatter plot between XYZ<sub>390–830 nm</sub> and XYZ<sub>360–830 nm</sub>, (<b>d</b>–<b>f</b>) is the adjustment scatter plot between XYZ<sub>400–830 nm</sub> and XYZ<sub>360–830 nm</sub>).</p>
Full article ">Figure 5
<p>The adjustment of the IAVW 390–830 nm and IAVW400–830 nm. ((<b>a</b>) The IAVW with the wavelength from 390 to 830 nm, (<b>b</b>) the IAVW with the wavelength from 400 to 830 nm).</p>
Full article ">Figure 6
<p>The chromatic diagram of the normal water bodies and algae bloom water bodies. ((<b>a</b>) normal waters (<b>b</b>) algae bloom waters).</p>
Full article ">Figure 7
<p>Normal water bodies and algae bloom water bodies WCIs (Water Chromatic Indices) histogram. ((<b>a</b>) Hue angle, (<b>b</b>) saturation, (<b>c</b>) λd, (<b>d</b>) IAVW. Blue bars are normal water and red bars are algal bloom waters).</p>
Full article ">Figure 7 Cont.
<p>Normal water bodies and algae bloom water bodies WCIs (Water Chromatic Indices) histogram. ((<b>a</b>) Hue angle, (<b>b</b>) saturation, (<b>c</b>) λd, (<b>d</b>) IAVW. Blue bars are normal water and red bars are algal bloom waters).</p>
Full article ">Figure 8
<p>Scatter plots of each other among the WCIs (Water Chromatic Indices). (Left column: normal water, scatter plot in blue: (<b>a</b>) saturation (S) vs. hue angle, (<b>b</b>) hue angle vs. IAVW, (<b>c</b>) saturation (S) vs. λd, (<b>d</b>) λd vs. IAVW, (<b>e</b>) saturation (S) vs. IAVW; right column: algal bloom water, scatter plot in red: (<b>f</b>) saturation (S) vs. hue angle, (<b>g</b>) hue angle vs. IAVW, (<b>h</b>) saturation (S) vs. λd, (<b>i</b>) λd vs. IAVW, (<b>j</b>) saturation (S) vs. IAVW.</p>
Full article ">Figure 8 Cont.
<p>Scatter plots of each other among the WCIs (Water Chromatic Indices). (Left column: normal water, scatter plot in blue: (<b>a</b>) saturation (S) vs. hue angle, (<b>b</b>) hue angle vs. IAVW, (<b>c</b>) saturation (S) vs. λd, (<b>d</b>) λd vs. IAVW, (<b>e</b>) saturation (S) vs. IAVW; right column: algal bloom water, scatter plot in red: (<b>f</b>) saturation (S) vs. hue angle, (<b>g</b>) hue angle vs. IAVW, (<b>h</b>) saturation (S) vs. λd, (<b>i</b>) λd vs. IAVW, (<b>j</b>) saturation (S) vs. IAVW.</p>
Full article ">Figure 9
<p>The chromatic diagram of different algal species in CIE 1931 chromatic coordinate.</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>d</b>) are the scatter plots of different algal species. ((<b>a</b>) The verses of the IAVW and the hue angle, (<b>b</b>) the verses of the IAVW and the λd, (<b>c</b>) the verses of the S and the hue angle, (<b>d</b>) the verses of the S and the λd). (<b>e</b>–<b>h</b>) are the radar charts for different algal species. ((<b>e</b>) IAVW, (<b>f</b>) hue angle, (<b>g</b>) saturation, (<b>h</b>) λd).</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>–<b>d</b>) are the scatter plots of different algal species. ((<b>a</b>) The verses of the IAVW and the hue angle, (<b>b</b>) the verses of the IAVW and the λd, (<b>c</b>) the verses of the S and the hue angle, (<b>d</b>) the verses of the S and the λd). (<b>e</b>–<b>h</b>) are the radar charts for different algal species. ((<b>e</b>) IAVW, (<b>f</b>) hue angle, (<b>g</b>) saturation, (<b>h</b>) λd).</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>–<b>d</b>) are the scatter plots of different algal species. ((<b>a</b>) The verses of the IAVW and the hue angle, (<b>b</b>) the verses of the IAVW and the λd, (<b>c</b>) the verses of the S and the hue angle, (<b>d</b>) the verses of the S and the λd). (<b>e</b>–<b>h</b>) are the radar charts for different algal species. ((<b>e</b>) IAVW, (<b>f</b>) hue angle, (<b>g</b>) saturation, (<b>h</b>) λd).</p>
Full article ">Figure 11
<p>(<b>a</b>) The 3-D (three-dimensional) plot of the <span class="html-italic">Dinoflagellates</span>, chlorophyll <span class="html-italic">a</span>, and the dominant wavelength (λd), IAVW. (<b>b</b>) The spectra of the <span class="html-italic">Dinoflagellates</span>. The colors of the spectral lines stand for the colors of the objects.</p>
Full article ">Figure 12
<p>The scatter plot of the Dinoflagellates. (The chlorophyll a vs. the (<b>a</b>) IAVW, (<b>b</b>) hue angle, (<b>c</b>) λd, (<b>d</b>) saturation).The red curve is the fitted curve and the blue dotted line is the numerical identification line.</p>
Full article ">Figure 13
<p>(<b>a</b>–<b>c</b>) are the scatter plots between the AVW400–700 nm and IAVW360–830 nm in normal waters, algae bloom waters, and both of the two types of waters. (<b>d</b>–<b>f</b>) are the histogram of the value of IAVW360–830 nm minus AVW400–700 nm. ((<b>a</b>,<b>d</b>) The normal waters, (<b>b</b>,<b>e</b>) the algae bloom waters, (<b>c</b>,<b>f</b>) both of the two types of the waters).</p>
Full article ">Figure 14
<p>The color discrimination between normal water and algal bloom water in the Bohai Sea based on spectral wavelength range from 360–830 nm. ((<b>a</b>) The chromatic point of the normal and algae bloom waters in Bohai Sea. (<b>b</b>) The spectra of the normal waters in Bohai Sea. (<b>c</b>) The algae bloom waters in the Bohai Sea. (The color of the curve was standard for the value of λd). The colors of the spectral lines stand for the colors of the objects).</p>
Full article ">Figure 15
<p>(<b>a</b>–<b>d</b>) are the scatter plot of each other among WCIs of the Bohai Sea. (The red dots are algal bloom water, and the blue dots are normal water). (<b>e</b>–<b>h</b>) are the box plot of chromatic indices calculated by spectral data of the Bohai Sea with different wavelength ranges from 360−830 nm. ((<b>e</b>) the box plot of the IAVW, (<b>f</b>) the box plot of the hue angle, (<b>g</b>) the box plot of the saturation, (<b>h</b>) the box plot of the λd.</p>
Full article ">Figure 15 Cont.
<p>(<b>a</b>–<b>d</b>) are the scatter plot of each other among WCIs of the Bohai Sea. (The red dots are algal bloom water, and the blue dots are normal water). (<b>e</b>–<b>h</b>) are the box plot of chromatic indices calculated by spectral data of the Bohai Sea with different wavelength ranges from 360−830 nm. ((<b>e</b>) the box plot of the IAVW, (<b>f</b>) the box plot of the hue angle, (<b>g</b>) the box plot of the saturation, (<b>h</b>) the box plot of the λd.</p>
Full article ">Figure 16
<p>The color discrimination between normal water and algal bloom water in Taihu Lake based on spectral wavelength range from 360 to 830 nm. ((<b>a</b>) The chromatic point of the normal and algae bloom waters in Bohai Sea. (<b>b</b>) The spectra of the normal waters in Bohai Sea. (<b>c</b>) The algae bloom waters in the Bohai Sea. (The color of the curve was standard for the value of λd). The colors of the spectral lines stand for the colors of the objects).</p>
Full article ">Figure 17
<p>(<b>a</b>–<b>d</b>) are the scatter plots of each other among WCIs of Taihu Lake. (The red dots are algal bloom water, and the blue dots are normal water). (<b>e</b>–<b>h</b>) are the box plot of chromatic indices calculated by spectral data of the Bohai Sea with different wavelength ranges from 360 to 830 nm. ((<b>e</b>) The box plot of the IAVW, (<b>f</b>) the box plot of the hue angle, (<b>g</b>) the box plot of the saturation, (<b>h</b>) the box plot of the λd.</p>
Full article ">Figure 17 Cont.
<p>(<b>a</b>–<b>d</b>) are the scatter plots of each other among WCIs of Taihu Lake. (The red dots are algal bloom water, and the blue dots are normal water). (<b>e</b>–<b>h</b>) are the box plot of chromatic indices calculated by spectral data of the Bohai Sea with different wavelength ranges from 360 to 830 nm. ((<b>e</b>) The box plot of the IAVW, (<b>f</b>) the box plot of the hue angle, (<b>g</b>) the box plot of the saturation, (<b>h</b>) the box plot of the λd.</p>
Full article ">
21 pages, 9141 KiB  
Article
Heavy Metal Groundwater Transport Mitigation from an Ore Enrichment Plant Tailing at Kazakhstan’s Balkhash Lake
by Dauren Muratkhanov, Vladimir Mirlas, Yaakov Anker, Oxana Miroshnichenko, Vladimir Smolyar, Timur Rakhimov, Yevgeniy Sotnikov and Valentina Rakhimova
Sustainability 2024, 16(16), 6816; https://doi.org/10.3390/su16166816 - 8 Aug 2024
Viewed by 442
Abstract
Sustainable potable groundwater supply is crucial for human development and the preservation of natural habitats. The largest endorheic inland lake in Kazakhstan, Balkhash Lake, is the main water resource for the arid southeastern part of the country. Several ore enrichment plants that are [...] Read more.
Sustainable potable groundwater supply is crucial for human development and the preservation of natural habitats. The largest endorheic inland lake in Kazakhstan, Balkhash Lake, is the main water resource for the arid southeastern part of the country. Several ore enrichment plants that are located along its shore have heavy metal pollution potential. The study area is located around a plant that has an evident anthropogenic impact on the Balkhash Lake aquatic ecological system, with ten known heavy metal toxic hotspots endangering fragile habitats, including some indigenous human communities. This study assessed the risk of heavy metal contamination from tailing dump operations, storage ponds, and related facilities and suggested management practices for preventing this risk. The coastal zone risk assessment analysis used an innovative integrated groundwater numerical flow and transport model that predicted the spread of groundwater contamination from tailing dump operations under several mitigation strategies. Heavy metal pollution prevention models included a no-action scenario, a filtration barrier construction scenario, and two scenarios involving the drilling of drainage wells between the pollution sources and the lake. The scenario assessment indicates that drilling ten drainage wells down to the bedrock between the existing drainage channel and the lake is the optimal engineering solution for confining pollution. Under these conditions, pollution from tailings will not reach Lake Balkhash during the forecast period. The methods and tools used in this study to enable mining activity without environmental implications for the region can be applied to sites with similar anthropogenic influences worldwide. Full article
Show Figures

Figure 1

Figure 1
<p>The study site location map and the Balkhash Industrial Area aerial photo display industrial objects included in the model’s schematization where the orange line is an interface with water bodies, the purple line is the tailing storage interface, black lines are barriers, and green lines are drains. The figure was prepared by Corel Draw with a base experimental site image taken from Google Earth.</p>
Full article ">Figure 2
<p>Hydrogeological cross-section along lines A–B (<a href="#sustainability-16-06816-f001" class="html-fig">Figure 1</a>). 1—upper-middle Quaternary lacustrine aquifer; 2—Pliocene aquitard of Pavlodar formations; 3—Miocene aquitard of Argyn formations; 4—Meso-Cenozoic water-bearing formations; 5—Carboniferous aquifer; 6—Paleozoic zone of fractured intrusive rocks; 7—tectonic faults; 8—groundwater level; 9—upper Quaternary technogenic aquifer, bulk soil; 10—sands with gravel inclusions; 11—crushed stones; 12—loams; 13—clays; 14—granites; 15—syenite porphyries; 16—dacite porphyries; 17—fractured rock; 18—well. Numbers: on top—well number, bottom—well depth, m; on the left in the numerator—mineralization, g/L; in the denominator—temperature, °C; on the right: in the numerator—well flow rate, L/s; in the denominator—drawdown, m. Shading corresponds to the chemical composition of groundwater in the sampled interval with the predominance of chloride and sulfate anions.</p>
Full article ">Figure 3
<p>The conceptual working process applied for the Balkhash Lake contamination risk assessment (<b>a</b>) and model application steps (<b>b</b>).</p>
Full article ">Figure 4
<p>Model calibration results.</p>
Full article ">Figure 5
<p>(<b>a</b>) Path lines of particles released from the source of pollution by the area tracked with MODPATH and (<b>b</b>) heavy metal spatial distribution in groundwater for ten years after contaminant release without a change in hydrogeological conditions (legend in <a href="#sustainability-16-06816-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 6
<p>Spatial distribution of heavy metals in groundwater from 14 drainage wells drilled between the drainage channel and Lake Balkhash (<b>a</b>) and for the scenario of drilling ten drainage wells between the drainage channel and Lake Balkhash (<b>b</b>).</p>
Full article ">Figure A1
<p>Spatial distribution of heavy metals in groundwater for the scenario of boundary construction between the drainage channel and Lake Balkhash.</p>
Full article ">Figure A2
<p>Spatial distribution of heavy metals in groundwater for the scenario of drainage wells drilled between the tailings pond drainage channel and the lake.</p>
Full article ">Figure A3
<p>Results of heavy metal concentration in groundwater monitoring wells over time.</p>
Full article ">Figure A4
<p>Sampling points on a map of heavy metal halo distribution in groundwater at the time of sampling in 2020.</p>
Full article ">Figure A5
<p>Calibration graph of the observed and calculated heavy metal concentrations at the sampling points.</p>
Full article ">Figure A6
<p>Relative sensitivity coefficients concerning different input parameters for a ±50% change in each parameter.</p>
Full article ">
22 pages, 3742 KiB  
Article
LAQUA: a LAndsat water QUality retrieval tool for east African lakes
by Aidan Byrne, Davide Lomeo, Winnie Owoko, Christopher Mulanda Aura, Kobingi Nyakeya, Cyprian Odoli, James Mugo, Conland Barongo, Julius Kiplagat, Naftaly Mwirigi, Sean Avery, Michael A. Chadwick, Ken Norris, Emma J. Tebbs and on behalf of the NSF-IRES Lake Victoria Research Consortium
Remote Sens. 2024, 16(16), 2903; https://doi.org/10.3390/rs16162903 - 8 Aug 2024
Viewed by 677
Abstract
East African lakes support the food and water security of millions of people. Yet, a lack of continuous long-term water quality data for these waterbodies impedes their sustainable management. While satellite-based water quality retrieval methods have been developed for lakes globally, African lakes [...] Read more.
East African lakes support the food and water security of millions of people. Yet, a lack of continuous long-term water quality data for these waterbodies impedes their sustainable management. While satellite-based water quality retrieval methods have been developed for lakes globally, African lakes are typically underrepresented in training data, limiting the applicability of existing methods to the region. Hence, this study aimed to (1) assess the accuracy of existing and newly developed water quality band algorithms for East African lakes and (2) make satellite-derived water quality information easily accessible through a Google Earth Engine application (app), named LAndsat water QUality retrieval tool for east African lakes (LAQUA). We collated a dataset of existing and newly collected in situ surface water quality samples from seven lakes to develop and test Landsat water quality retrieval models. Twenty-one published algorithms were evaluated and compared with newly developed linear and quadratic regression models, to determine the most suitable Landsat band algorithms for chlorophyll-a, total suspended solids (TSS), and Secchi disk depth (SDD) for East African lakes. The three-band algorithm, parameterised using data for East African lakes, proved the most suitable for chlorophyll-a retrieval (R2 = 0.717, p < 0.001, RMSE = 22.917 μg/L), a novel index developed in this study, the Modified Suspended Matter Index (MSMI), was the most accurate for TSS retrieval (R2 = 0.822, p < 0.001, RMSE = 9.006 mg/L), and an existing global model was the most accurate for SDD estimation (R2 = 0.933, p < 0.001, RMSE = 0.073 m). The LAQUA app we developed provides easy access to the best performing retrieval models, facilitating the use of water quality information for management and evidence-informed policy making for East African lakes. Full article
(This article belongs to the Special Issue Remote Sensing Band Ratios for the Assessment of Water Quality)
Show Figures

Figure 1

Figure 1
<p>Overview of the methodology used for data collection, model assessment, and application development. Detailed app development steps are provided in <a href="#sec2dot6-remotesensing-16-02903" class="html-sec">Section 2.6</a>. TOA means top-of-atmosphere.</p>
Full article ">Figure 2
<p>(<b>A</b>) The seven study lakes with <span class="html-italic">in situ</span> water quality data used for model development: 1 is Ziway, Ethiopia; 2 is Chamo, Ethiopia; 3 is Turkana, Kenya/Ethiopia; 4 is Baringo, Kenya; 5 is Bogoria, Kenya; 6 is Oloidien, Kenya; 7 is Victoria, Kenya/Uganda/Tanzania. (<b>B</b>) The data collection transects for Lake Baringo, Kenya, in September 2023. Diamonds indicate data collected on 18 September and triangles indicate data collected on 19 September. (<b>C</b>) The region in which <span class="html-italic">in situ</span> data were collected for Lake Victoria in this study. (<b>D</b>) The data collection transects in Winam Gulf, in the Kenyan region of Lake Victoria, on 13 September 2023. Diamonds indicate individual sampling points along each transect.</p>
Full article ">Figure 3
<p>Flow diagram of the Google Earth Engine app development steps, image preprocessing methods, and model application.</p>
Full article ">Figure 4
<p>(<b>A</b>) Models for the best predictive band algorithms for chlorophyll-<span class="html-italic">a</span> and total suspended solids (TSS). There is no plot for Secchi disk depth (SDD) as the best performing model utilises the existing Song <span class="html-italic">et al.</span> (2022) equation. Grey bars indicate 95% confidence intervals. Data points from each study lake are distinguished by colour and marker shape and are summarized in the lake key. (<b>B</b>) Predicted vs. observed values for the best performing models for chlorophyll-<span class="html-italic">a</span>, TSS, and SDD. The black lines represent the linear relationship, and the grey bars are the 95% confidence intervals. Dashed lines indicate a perfect match with a slope of 1 and intercept of 0.</p>
Full article ">Figure 5
<p>The best performing retrieval models applied to a Landsat 9 image from September 2023 for (<b>A</b>) Lake Baringo—a turbid freshwater lake in Kenya, and (<b>B</b>) Lake Bogoria—a highly productive alkaline–saline lake approximately 20 km south of Baringo.</p>
Full article ">
22 pages, 5322 KiB  
Review
Trends and Innovations in Surface Water Monitoring via Satellite Altimetry: A 34-Year Bibliometric Review
by Zhengkai Huang, Rumiao Sun, Haihong Wang and Xin Wu
Remote Sens. 2024, 16(16), 2886; https://doi.org/10.3390/rs16162886 - 7 Aug 2024
Viewed by 367
Abstract
The development of satellite altimetry has significantly advanced the application of satellite Earth observation technologies in surface water monitoring, resulting in a substantial body of research. Although numerous reviews have summarized progress in this field, their analyses are often limited in scope and [...] Read more.
The development of satellite altimetry has significantly advanced the application of satellite Earth observation technologies in surface water monitoring, resulting in a substantial body of research. Although numerous reviews have summarized progress in this field, their analyses are often limited in scope and fail to provide a systematic, quantitative assessment of the current research prospects and trends. To address this gap, we utilize CiteSpace and VOSviewer bibliometric software to analyze 13,500 publications from the WOS database, spanning the years from 1988 to 2022. Our analysis focused on publication volume, authorship, collaboration networks, and content. We also compare data from Google Scholar and Scopus to validate the reliability of our dataset. Our findings indicate a steadily growing research potential in this field, as evidenced by trends in publication volume, authorship, journal influence, and disciplinary focus. Notably, the leading journals are primarily in the realm of remote sensing, while key disciplines include geology, remote sensing science, and oceanography. Keyword analysis revealed current research hotspots such as sea-level rise, snow depth, and machine learning applications. Among various water body types, research on glaciers ranks second only to ocean studies. Furthermore, research focus areas are shifting from large oceanic regions like the Pacific and Atlantic Oceans to significant inland water bodies, notably the Tibetan Plateau and the Amazon basin. This study combines qualitative and quantitative methods to analyze vast amounts of information in the field of surface water monitoring by satellite altimetry. The resulting visualizations provide researchers with clear insights into the development trends and patterns within this domain, offering valuable support for identifying future research priorities and directions. Full article
Show Figures

Figure 1

Figure 1
<p>Flowchart of the technological roadmap.</p>
Full article ">Figure 2
<p>Graph of publication volume and citation frequency of the literature.</p>
Full article ">Figure 3
<p>Information on the 20 authors with the most publications: (<b>a</b>) Research Output, (<b>b</b>) Centrality Index, (<b>c</b>) H-index.</p>
Full article ">Figure 4
<p>Author partnership analysis: (<b>a</b>) CiteSpace partnership analysis, (<b>b</b>) VOSviewer–based partnership network.</p>
Full article ">Figure 5
<p>Journal and discipline analysis: (<b>a</b>) journal titles; (<b>b</b>) disciplines.</p>
Full article ">Figure 6
<p>Literature co-citation network diagram.</p>
Full article ">Figure 7
<p>Statistics of hotspot areas: (<b>a</b>) is the percentage of each water body in the total number of documents; (<b>b</b>) represents the top 5 hotspot research areas in the water body.</p>
Full article ">Figure 8
<p>Keyword co-occurrence network diagram.</p>
Full article ">Figure 9
<p>Keyword emergence analysis chart, visualizing 52 keywords. The solid circle in the figure marks the peak research year for the keyword during this timeframe, with its size indicative of the volume of literature published on the keyword in that year.</p>
Full article ">Figure 10
<p>Comparison of citation counts between Web of Science, Google Scholar, and Scopus.</p>
Full article ">
16 pages, 3712 KiB  
Article
Plants Restoration Drives the Gobi Soil Microbial Diversity for Improving Soil Quality
by Lizhi Wang, Junyong Ma, Qifeng Wu, Yongchao Hu and Jinxiao Feng
Plants 2024, 13(15), 2159; https://doi.org/10.3390/plants13152159 - 5 Aug 2024
Viewed by 402
Abstract
Desertification and salt stress are major causes of terrestrial ecosystem loss worldwide, and the Gobi, representing a salt-stressed area in inland China, has a major impact on the ecosystems and biodiversity of its surrounding environment. The restoration of the Gobi Desert is an [...] Read more.
Desertification and salt stress are major causes of terrestrial ecosystem loss worldwide, and the Gobi, representing a salt-stressed area in inland China, has a major impact on the ecosystems and biodiversity of its surrounding environment. The restoration of the Gobi Desert is an important way to control its expansion, but there are few studies on the evaluation of restoration. In this study, soils under different restoration scenarios, namely, soils in restored areas (R1, R2), semi-restored areas (SR1, SR2), and unrestored control areas (C1, C2), were used to investigate differences in microbial diversity and physicochemical properties. The results showed that the soil was mainly dominated by particles of 4–63 μm (26.45–37.94%) and >63 μm (57.95–72.87%). Across the different restoration levels, the soil pH (7.96–8.43) remained basically unchanged, salinity decreased from 9.23–2.26 to 0.24–0.25, and water content remained constant (10.98–12.27%) except for one restored sample in which it was higher (22.32%). The effective Al, Cu, and Zn in the soil increased, but only slightly. Total organic matter (TOM) decreased from 3.86–5.20% to 1.31–1.47%, and total organic nitrogen (TON) decreased from 0.03–0.06% to 0.01–0.02%, but the difference in total organic carbon (TOC) was not significant. High-throughput testing revealed that the bacterial population of the restored area was dominated by A4b (6.33–9.18%), MND1 (4.94–7.39%), and Vicinamibacteraceae (7.04–7.39%). Regarding archaea, samples from the restored areas were dominated by Marine Group II (76.17–81.49%) and Candidatus Nitrososphaera (6.07–9.75%). PCoA showed that the different restoration levels were the main cause of the differences between the samples. Additionally, salinity was the dominant factor that induced this difference, but it was inhibited by the restoration and targeted enrichment of some of these functional genera. Desert restoration should therefore focus on conserving water rather than adding nutrients. Planting salt- and drought-tolerant vegetation will contribute to the initial restoration of the desert and the restoration of the microbiological content of the soil as it migrates over time, creating a cycle of elements. Restoration stimulates and enhances the microbial diversity of the soil via beneficial microorganisms. Full article
Show Figures

Figure 1

Figure 1
<p>High-throughput bacterial assays of samples from the study areas. R1, R2: restored areas; SR1, SR2: semi-restored areas; C1, C2: unrestored control areas in the Xinjiang Gobi. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi. In the figure, the samples are UPGMA-clustered according to the Euclidean distances between the species. Species are clustered via UPGMA clustering. The data in the figure are shown on a logarithmic scale.</p>
Full article ">Figure 2
<p>High-throughput assays of archaea in the study area. R1, R2: restored areas; SR1, SR2: semi-restored areas; C1, C2: unrestored control areas in the Xinjiang Gobi. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi. In the figure, the samples are UPGMA-clustered according to the Euclidean distances between the species. Species are clustered via UPGMA clustering. The data in the figure are shown on a logarithmic scale.</p>
Full article ">Figure 3
<p>Principal coordinate analysis of bacteria. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi.</p>
Full article ">Figure 4
<p>Principal coordinate analysis of archaea. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi.</p>
Full article ">Figure 5
<p>A Venn analysis based on different restoration statuses revealed the number of bacterial OTUs in each. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi.</p>
Full article ">Figure 6
<p>A Venn analysis based on different restoration statuses revealed the number of archaeal OTUs in each. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi.</p>
Full article ">Figure 7
<p>CCAs based on bacterial communities in samples. R: restored area; SR: semi-restored area; C: unrestored control area in the Xinjiang Gobi.</p>
Full article ">Figure 8
<p>CCAs based on archaeal communities in samples. R: restored area; SR: semi-restored area; C: unrestored control areas in the Xinjiang Gobi.</p>
Full article ">Figure 9
<p>The study area: the saline and alkaline cultivated land at Xinjiang Production and Construction Corps. The black dots denote the stations in unrestored control areas (C1, C2); the brownish dots denote the stations in the semi-restored areas (SR1, SR2); the brownish dots denote the stations in the fully restored areas (R1, R2). The bule was represent the Xinjiang Uighur Autonomous Region.</p>
Full article ">
19 pages, 6977 KiB  
Article
Population Biology of the Non-Indigenous Rayed Pearl Oyster (Pinctada radiata) in the South Evoikos Gulf, Greece
by Dimitris Pafras, Alexandros Theocharis, Gerasimos Kondylatos, Alexis Conides and Dimitris Klaoudatos
Diversity 2024, 16(8), 460; https://doi.org/10.3390/d16080460 - 1 Aug 2024
Viewed by 472
Abstract
The Atlantic pearl oyster Pinctada radiata (Leach, 1814), the first documented Lessepsian bivalve species to enter the Mediterranean basin, is present in various coastal areas in Greece, and constitutes, almost exclusively, a domestic commercial bivalve resource. The present study aimed to contribute to [...] Read more.
The Atlantic pearl oyster Pinctada radiata (Leach, 1814), the first documented Lessepsian bivalve species to enter the Mediterranean basin, is present in various coastal areas in Greece, and constitutes, almost exclusively, a domestic commercial bivalve resource. The present study aimed to contribute to the limited information available on P. radiata population structure and dynamics in Hellenic waters, especially following the recent enforcement of legislation for regulation of its fishery. A total of 703 individuals were collected using scuba diving from the South Evoikos Gulf. The male-to-female ratio (1:1.70) significantly departed from 1:1. A higher probability for female prevalence was exhibited for shell heights over 50.77 mm. Significant differences were exhibited in the shell height–total weight relationship between the sexes. The fourth-year class was the dominant cohort, comprising 50.09% of the population, out of the seven age classes identified. Asymptotic length was estimated at 109.1 mm and growth index at 3.35, respectively. Longevity was estimated at 15.7 years, with natural mortality (M) at 0.39 and total mortality (Z) at 0.76. The probability of capture (LC50) was estimated at 50.72 mm at 2.8 years. Biological reference points FMSY and EMSY were higher than the fishing mortality and current exploitation rate, respectively, indicating the potential for further population exploitation. Full article
Show Figures

Figure 1

Figure 1
<p>Map of the study area (red outline) and location of sampling area (black outline) (color variation indicates depth).</p>
Full article ">Figure 2
<p>Morphological measurements of the shell of the rayed pearl oyster of <span class="html-italic">Pinctada radiata</span> collected from the coastal waters of South Evoikos Gulf (Greece). L: left valve, R: right valve, SH: shell height, SWI: shell width, HL: hinge length, WNL: width of the nacreous part in left valve, and LNR: length of nacreous part in right valve.</p>
Full article ">Figure 3
<p>Frequency distributions with overlayed fitted normal distribution of biometric characters (mm and g) of <span class="html-italic">Pinctada radiata</span> collected from the coastal waters of South Evoikos Gulf (Greece). (<b>A</b>): Total weight; (<b>B</b>) Flesh weight; (<b>C</b>) Shell weight; (<b>D</b>) Shell height; (<b>E</b>) Hinge length; and (<b>F</b>) Shell width.</p>
Full article ">Figure 4
<p>Scatterplot matrix with fitted line plots (lower left triangle of the scatterplot matrix) and heat map with Pearson correlation and associated probability values (upper-right triangle of the scatterplot matrix) of <span class="html-italic">Pinctada radiata</span> biometric characters measured from the coastal waters of South Evoikos Gulf (Greece). The color of each circle represents correlation strength circle size represents correlation significance between each pair of variables (larger circle indicates a more significant relationship).</p>
Full article ">Figure 5
<p>Temporal comparison of densities plot of flesh weight of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece).</p>
Full article ">Figure 6
<p>Nominal logistic curve and threshold values (values above which there is an increasingly higher probability that sex is female) of the effect of size (shell height) on the sex ratio of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece).</p>
Full article ">Figure 7
<p>Shell height–total weight relationship of (<b>A</b>) both sexes and (<b>B</b>) each sex separately of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece).</p>
Full article ">Figure 8
<p>Characteristics of the identified age groups for all captured individuals of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece). Confidence intervals indicate the standard deviation.</p>
Full article ">Figure 9
<p>Seasonally oscillating growth curve fitted to monthly length-frequency data of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece).</p>
Full article ">Figure 10
<p>Probability of capture for different length classes (LC<sub>25</sub>, LC<sub>50</sub>, LC<sub>75</sub>) of <span class="html-italic">Pinctada radiata</span> individuals collected from the coastal waters of South Evoikos Gulf (Greece).</p>
Full article ">Figure 11
<p>Yield per recruit (Y/R) and biomass per recruit (B/R) of <span class="html-italic">Pinctada radiata</span> collected from the coastal waters of South Evoikos Gulf (Greece), for different fishing mortalities.</p>
Full article ">Figure 12
<p>Yield per recruit (Y/R) and biomass per recruit (B/R) of <span class="html-italic">Pinctada radiata</span> collected from the coastal waters of South Evoikos Gulf (Greece), for different fishing exploitation rates.</p>
Full article ">
22 pages, 7632 KiB  
Article
Exploring Spatial Aggregations and Temporal Windows for Water Quality Match-Up Analysis Using Sentinel-2 MSI and Sentinel-3 OLCI Data
by Tanja Schröder, Susanne I. Schmidt, Rebecca D. Kutzner, Hendrik Bernert, Kerstin Stelzer, Kurt Friese and Karsten Rinke
Remote Sens. 2024, 16(15), 2798; https://doi.org/10.3390/rs16152798 - 30 Jul 2024
Viewed by 407
Abstract
Effective monitoring and management of inland waterbodies depend on reliable assessments of water quality through remote sensing technologies. Match-up analysis plays a significant role in investigating the comparability between in situ and remote sensing data of physical and biogeochemical variables. By exploring different [...] Read more.
Effective monitoring and management of inland waterbodies depend on reliable assessments of water quality through remote sensing technologies. Match-up analysis plays a significant role in investigating the comparability between in situ and remote sensing data of physical and biogeochemical variables. By exploring different spatial aggregations and temporal windows, we aimed to identify which configurations are most effective and which are less effective for the assessment of remotely sensed water quality data within the context of governmental monitoring programs. Therefore, in this study, remote sensing data products, including the variables of Secchi depth, chlorophyll-a, and turbidity, derived from the Copernicus satellites Sentinel-2 and Sentinel-3, were compared with in situ laboratory data from >100 waterbodies (lakes and reservoirs) in Germany, covering a period of 5 years (2016–2020). Processing was carried out using two different processing schemes, CyanoAlert from Brockmann Consult GmbH and eoapp AQUA from EOMAP GmbH & Co. KG, in order to analyze the influence of different processors on the results. To investigate appropriate spatial aggregations and time windows for validation (the match-up approach), we performed a statistical comparison of different spatial aggregations (1 pixel; 3 × 3, 5 × 5, and 15 × 15 macropixels; and averaging over the whole waterbody) and time windows (same day, ±1 day, and ±5 days). The results show that waterbody-wide values achieved similar accuracies and biases compared with the macropixel variants, despite the large differences in spatial aggregation and spatial variability. An expansion of the temporal window to up to ±5 days did not impair the agreement between the in situ and remote sensing data for most target variables and sensor–processor combinations, while resulting in a marked rise in the number of matches. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Inland waterbodies investigated. (<b>b</b>) Exemplary illustration of the spatial aggregations extracted for MSI data.</p>
Full article ">Figure 2
<p>Overall performance of different spatial aggregations using the match-up dataset (same day) with MSI data processed by CyanoAlert (orange) and EOMAP-MIP (blue). Scatterplots are shown in log-log scale with selected error metrics and the number of observations (both processors combined) for the three target variables of chlorophyll-a (<b>top</b>), turbidity (<b>middle</b>), and Secchi depth (<b>bottom</b>). The grey dashed line refers to the 1:1 line.</p>
Full article ">Figure 3
<p>Overall performance of different spatial aggregations (left column, 1 pixel; middle column, 3 × 3 macropixels; right column, waterbody scale) using the match-up dataset (same day) with OLCI data processed by CyanoAlert (orange) and EOMAP-MIP (blue). Scatterplots are shown in log-log scale with selected error metrics and the number of observations (both processors combined) for the three target variables of chlorophyll-a (<b>top</b> row), turbidity (<b>middle</b> row), and Secchi depth (<b>bottom</b> row). The grey dashed line refers to the 1:1 line.</p>
Full article ">Figure 4
<p>Overall performance of different spatial aggregations using the match-up dataset (same day) with MSI and OLCI data processed by CyanoAlert (top) and EOMAP-MIP (bottom). The table shows selected error metrics, the number of observations (matches), and the number of waterbodies (N Lakes) for the three target variables of chlorophyll-a, turbidity, and Secchi depth. Dark grey shades indicate the poorer performance of the variant, while the lightest shade represents the best performance within each processing scheme.</p>
Full article ">Figure 5
<p>Overall performance of different temporal windows (left column, same day; middle column, ±1 day; right column, ±5 days) using the match-up dataset (3 × 3 macropixel) with MSI data processed by CyanoAlert (orange) and EOMAP-MIP (dark blue), and OLCI data processed by CyanoAlert (yellow) and EOMAP-MIP (light blue). Scatterplots are shown in log-log scale with selected error metrics and the number of observations for the three target variables of chlorophyll-a (<b>top</b> row), turbidity (<b>middle</b> row), and Secchi depth (<b>bottom</b> row). The grey dashed line refers to the 1:1 line.</p>
Full article ">Figure 6
<p>Overall performance of different temporal windows using the match-up dataset with MSI and OLCI data processed by CyanoAlert (top) and EOMAP-MIP (bottom). The table shows selected error metrics, the number of observations (N, “matches”), and the number of waterbodies (N Lakes) for the three target variables of chlorophyll-a, turbidity, and Secchi depth. Dark grey shades indicate poorer performance of the variant, while the lightest shades represent the best performance within each processing scheme.</p>
Full article ">Figure A1
<p>Scatterplots of chlorophyll-a (<b>left</b>), turbidity (<b>middle</b>), and Secchi depth (<b>right</b>) match-up dataset, log10 transformed (<b>bottom</b>) and untransformed (<b>top</b>).</p>
Full article ">Figure A2
<p>Overall performance of different spatial aggregations using the match-up dataset (same-day) with MSI and OLCI data separated, as in <a href="#remotesensing-16-02798-f004" class="html-fig">Figure 4</a>, but the two processors EOMAP-MIP and CynaoAlert combined. Table shows selected error metrics, number of observations (matches) and number of waterbodies (N Lakes) for the three target variables chlorophyll-a, turbidity and Secchi depth. Dark-grey shades indicate poorer performance of variant, while lightest shades represent the best performance.</p>
Full article ">Figure A3
<p>Overall performance of two different spatial aggregations (3 × 3 macropixels and waterbody-scale variants) using the match-up dataset (same day) with data from both sensors (S2-MSI, S3-OLCI) and processors (EOMAP-MIP, CyanoAlert) combined. The table shows selected error metrics, the number of observations (matches; N), and the number of waterbodies (N Lakes) for the three target variables of chlorophyll-a, turbidity, and Secchi depth.</p>
Full article ">Figure A4
<p>Overall performance of different temporal windows using the match-up dataset (3 × 3 macropixels) with data of both sensors (MSI, OLCI) and processors (EOMAP-MIP, CyanoAlert) combined. The table shows selected error metrics, the number of observations (matches), and the number of waterbodies (N Lakes) for the three target variables of chlorophyll-a, turbidity, and Secchi depth. Dark grey shades indicate poorer performance of the variant, while the lightest shades represent the best performance.</p>
Full article ">
30 pages, 12891 KiB  
Article
Evaluation of GPM IMERG Early, Late, and Final Run in Representing Extreme Rainfall Indices in Southwestern Iran
by Mohammad Sadegh Keikhosravi-Kiany and Robert C. Balling
Remote Sens. 2024, 16(15), 2779; https://doi.org/10.3390/rs16152779 - 30 Jul 2024
Viewed by 281
Abstract
The growing concerns about floods have highlighted the need for accurate and detailed precipitation data as extreme precipitation occurrences can lead to catastrophic floods, resulting in significant economic losses and casualties. Integrated Multi-satellitE Retrievals for the Global Precipitation Measurement (GPM IMERG) is a [...] Read more.
The growing concerns about floods have highlighted the need for accurate and detailed precipitation data as extreme precipitation occurrences can lead to catastrophic floods, resulting in significant economic losses and casualties. Integrated Multi-satellitE Retrievals for the Global Precipitation Measurement (GPM IMERG) is a commonly used high-resolution gridded precipitation dataset and is recognized as trustworthy alternative sources of precipitation data. The aim of this study is to comprehensively evaluate the performance of GPM IMERG Early (IMERG-E), Late (IMERG-L), and Final Run (IMERG-F) in precipitation estimation and their capability in detecting extreme rainfall indices over southwestern Iran during 2001–2020. The Asfezari gridded precipitation data, which are developed using a dense of ground-based observation, were utilized as the reference dataset. The findings indicate that IMERG-F performs reasonably well in capturing many extreme precipitation events (defined by various indices). All three products showed a better performance in capturing fixed and non-threshold precipitation indices across the study region. The findings also revealed that both IMERG-E and IMERG-L have problems in rainfall estimation over elevated areas showing values of overestimations. Examining the effect of land cover type on the accuracy of the precipitation products suggests that both IMERG-E and IMERG-L show large and highly unrealistic overestimations over inland water bodies and permanent wetlands. The results of the current study highlight the potential of IMERG-F as a valuable source of data for precipitation monitoring in the region. Full article
(This article belongs to the Special Issue Remote Sensing of Precipitation Extremes)
Show Figures

Figure 1

Figure 1
<p>General location of the study region in Northern Hemisphere (<b>a</b>) and topography of the region (<b>b</b>).</p>
Full article ">Figure 2
<p>Spatial values of (POD), (FAR), and (CSI) in the study area for IMERG-E (<b>a</b>,<b>d</b>,<b>g</b>), IMERG-L (<b>b</b>,<b>e</b>,<b>h</b>), and IMERG-F (<b>c</b>,<b>f</b>,<b>i</b>).</p>
Full article ">Figure 3
<p>Monthly amount of precipitation in the study area averaged over 2001–2020 derived from Asfezari, IMERG-E, IMERG-L, and IMERG-F.</p>
Full article ">Figure 4
<p>Scatterplots of precipitation for IMERG-E, IMERG-L, and IMERG-F compared to Asfezari for winter (<b>a</b>–<b>c</b>), spring (<b>d</b>–<b>f</b>), summer (<b>g</b>–<b>i</b>), and fall (<b>j</b>–<b>l</b>).</p>
Full article ">Figure 5
<p>Seasonal values of precipitation for winter (<b>a</b>–<b>d</b>), spring (<b>e</b>–<b>h</b>), summer (<b>i</b>–<b>l</b>), and fall (<b>m</b>–<b>p</b>) derived from Asfezari, IMERG-E, IMERG-L, and IMERG-F.</p>
Full article ">Figure 6
<p>Density-colored scatterplots of IMERG-E, IMERG-L, and IMERG-F against Asfezari, for winter (<b>a</b>–<b>c</b>), spring (<b>d</b>–<b>f</b>), summer (<b>g</b>–<b>i</b>), and fall (<b>j</b>–<b>l</b>) over the study region. The color represents the occurrence frequency.</p>
Full article ">Figure 7
<p>Annual values of precipitation derived from (<b>a</b>) IMERG-E, (<b>b</b>) IMERG-L, and (<b>c</b>) IMERG-F, and Asfezari (<b>d</b>) averaged over 2001–2020 and annual density-colored scatterplots of IMERG-E (<b>e</b>), IMERG-L (<b>f</b>), and IMERG-F (<b>g</b>) against Asfezari over the study region. The color represents the occurrence frequency.</p>
Full article ">Figure 8
<p>Land cover types of the study area derived from MCD12Q1.061 (<b>a</b>) along with focus on the permanent wet lands and inland water bodies (<b>b</b>) and Map of Google Earth image depicting earth surface features (<b>c</b>).</p>
Full article ">Figure 9
<p>The bias (%) of IMERG-E (<b>a</b>), IMERGE-L (<b>b</b>), and IMERG-F (<b>c</b>) against Asfezari for each of the elevation levels.</p>
Full article ">Figure 10
<p>Annual values of precipitation derived from (<b>a</b>) IMERG-E V07, (<b>b</b>) IMERG-L V07, and (<b>c</b>) IMERG-F V07, and Asfezari (<b>d</b>) averaged over 2001–2020 and annual density-colored scatterplots of IMERG-E V07 (<b>e</b>), IMERG-L V07 (<b>f</b>), and IMERG-F V07 (<b>g</b>) against Asfezari over the study region. The color represents the occurrence frequency.</p>
Full article ">Figure 11
<p>The bias (%) of IMERG-E V07 (<b>a</b>), IMERGE-L V07 (<b>b</b>), and IMERG-F V07 (<b>c</b>) against Asfezari for each of the elevation levels.</p>
Full article ">Figure 12
<p>Long-term means of fixed threshold extreme precipitation indices (R10, R20, CWD, CDD) for Asfezari (<b>a</b>,<b>e</b>,<b>i</b>,<b>m</b>), IMERG-E (<b>b</b>,<b>f</b>,<b>j</b>,<b>n</b>), IMERG-L(<b>c</b>,<b>g</b>,<b>k</b>,<b>o</b>), and IMERG-F (<b>d</b>,<b>h</b>,<b>l</b>,<b>p</b>).</p>
Full article ">Figure 13
<p>Density-colored scatterplots of extreme precipitation indices (R10, R20, CWD, CDD) for IMERG-E (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>), IMERG-L (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>), and IMERG-F (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) against the Asfezari dataset. The color represents the occurrence frequency.</p>
Full article ">Figure 14
<p>Long-term means of grid-related extreme precipitation indices (R95p, R99p, R95pTOT, and R99pTOT) for Asfezari (<b>a</b>,<b>e</b>,<b>i</b>,<b>m</b>), IMERG-E (<b>b</b>,<b>f</b>,<b>j</b>,<b>n</b>), IMERG-L(<b>c</b>,<b>g</b>,<b>k</b>,<b>o</b>), and IMERG-F (<b>d</b>,<b>h</b>,<b>l</b>,<b>p</b>).</p>
Full article ">Figure 15
<p>Density-colored scatterplots of extreme precipitation indices (R95p, R99p, R95pTOT, and R99pTOT) for IMERG-E (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>), IMERG-L (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>), and IMERG-F (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) against the Asfezari dataset. The color represents the occurrence frequency.</p>
Full article ">Figure 16
<p>Long-term means of non-threshold indices extreme precipitation indices (Rx1day (mm), SDII (mm), and PRCPTOT (mm)) for Asfezari (<b>a</b>,<b>e</b>,<b>i</b>), IMERG-E (<b>b</b>,<b>f</b>,<b>j</b>), IMERG-L(<b>c</b>,<b>g</b>,<b>k</b>), and IMERG-F (<b>d</b>,<b>h</b>,<b>l</b>).</p>
Full article ">Figure 17
<p>Density-colored scatterplots of extreme precipitation indices (Rx1day, SDII, PRCPTOT) for IMERG-E (<b>a</b>,<b>d</b>,<b>g</b>), IMERG-L (<b>b</b>,<b>e</b>,<b>h</b>), and IMERG-F (<b>c</b>,<b>f</b>,<b>i</b>) against the Asfezari dataset. The color represents the occurrence frequency.</p>
Full article ">Figure 18
<p>Temporal variation in the fixed threshold indices generated from Asfezari, IMERG-E, IMERG-L, and IMERG-F over 2001–2020.</p>
Full article ">Figure 19
<p>Temporal variation in the grid-related threshold indices generated from Asfezari, IMERG-E, IMERG-L, and IMERG-F over 2001–2020.</p>
Full article ">Figure 20
<p>Temporal variation in the non-threshold indices generated from Asfezari, IMERG-E, IMERG-L, and IMERG-F over 2001–2020.</p>
Full article ">Figure 21
<p>Relevant error box plots for Asfezari, IMERG-E, IMERG-L, and IMERG-F for fixed threshold indices (<b>a</b>–<b>d</b>), grid-related threshold indices (<b>e</b>–<b>h</b>), and non-threshold indices (<b>i</b>–<b>k</b>). The whiskers denote the maximum and minimum values in the data. The boxes extending from Q1 to Q3 show the median, while the red + symbols show outliers.</p>
Full article ">
23 pages, 4390 KiB  
Article
Forecasting Meteorological Drought Conditions in South Korea Using a Data-Driven Model with Lagged Global Climate Variability
by Seonhui Noh and Seungyub Lee
Sustainability 2024, 16(15), 6485; https://doi.org/10.3390/su16156485 - 29 Jul 2024
Viewed by 440
Abstract
Drought prediction is crucial for early risk assessment, preventing negative impacts and the timely implementation of mitigation measures for sustainable water management. This study investigated the relationship between climate variations in three seas and the prediction of December meteorological droughts in South Korea, [...] Read more.
Drought prediction is crucial for early risk assessment, preventing negative impacts and the timely implementation of mitigation measures for sustainable water management. This study investigated the relationship between climate variations in three seas and the prediction of December meteorological droughts in South Korea, using the Standardized Precipitation Evapotranspiration Index (SPEI). Climate indices with multiple time lags were integrated into multiple linear regression (MLR) and Random Forest (RF) models and evaluated using Pearson’s correlation coefficients (PCCs) and the Root Mean Square Error (RMSE). The results indicated that the MLR model outperformed RF model in the western inland region with a PCC of 0.52 for predicting SPEI-2. On the other hand, the RF model effectively predicted drought states of ‘moderate drought’ or worse (SPEI < −1) nationwide, achieving an average hit rate of 47.17% and Heidke skill score (HSS) of 0.56, particularly excelling in coastal areas. Nino 3.4 turned out to be the most influential factor for short-period extreme droughts (SPEI-2) with a three-month lag, contributed by the Pacific, Atlantic, and Indian Oceans. For periods of four months or longer, climate variations had a lower predictive value. However, integrating autocorrelation functions to account for the previous month’s drought status improved the accuracy. A HYBRID model, which blends linear and nonlinear approaches, further enhanced reliability, making the proposed model more applicable for drought forecasting in neighboring countries and valuable for South Korea’s drought monitoring system to support sustainable water management. Full article
Show Figures

Figure 1

Figure 1
<p>General illustration of (<b>a</b>) study area and spatial distribution of average annual precipitation anomalies and temporal anomalies and trend of (<b>b</b>) the precipitation and (<b>c</b>) the temperature in South Korea during the period 1973–2022.</p>
Full article ">Figure 2
<p>Time series of 50 years (1973–2022) of December SPEI-2 for all K-Hidra grids across South Korea. The box for each year reflects the distribution of SPEI values all station. The lower and upper hinges correspond to the first and third quartiles (the 25th and 75th percentiles). Values below y = −1 indicate ‘moderate drought’ and are demarcated by a blue dotted line. The average values from 205 observation stations are connected with a red solid line. Outliers are dark orange, X-shaped.</p>
Full article ">Figure 3
<p>Pearson’s correlation coefficients between SPEI-2 and lagged large-scale climate indices for K-Hidra grid data across South Korea.</p>
Full article ">Figure 4
<p>Model performance metrics for predicting December SPEI-2 using 10 lagged climate indices. (<b>a</b>,<b>b</b>) display PCC values, while (<b>c</b>,<b>d</b>) show RMSE values. The left panels represent the MLR model, and the right panels represent the RF model. Redder colors indicate better performance.</p>
Full article ">Figure 5
<p>Comparison between observed values of SPEI-2 and predicted model in December 2015. (<b>a</b>) is the observed value, (<b>b</b>) is the predicted value of the MLR model, and (<b>c</b>) is the predicted value of the RF model.</p>
Full article ">Figure 6
<p>Time series comparison of MLR and RF models for SPEI-2 prediction at grids. Note that observed SPEI-2 values are in gray, MLR predictions in blue, and RF predictions in red. The time series presented are the results of the MLR (blue) and RF (red) models. The observed SPEI-2 is shown in gray. (<b>a</b>,<b>c</b>) are time series comparisons on the best and worst skill between MLR model predictions and observations, and (<b>b</b>,<b>d</b>) are best and worst skill for RF respectively.</p>
Full article ">Figure 7
<p>Model proficiency map for predicting December SPEI-2 drought conditions. Panels (<b>a</b>,<b>b</b>) show hit rates, while panels (<b>c</b>,<b>d</b>) present HSS values. The left panels represent the MLR model, and the right panels represent the RF model. Redder colors indicate better performance in both hit rates and HSS.</p>
Full article ">Figure 8
<p>Predictor count of selection out of 10 variables across 205 grids for MLR and RF models.</p>
Full article ">Figure 9
<p>Predictor proportion of selection out of 110 possible selections (11 lag times for 10 variables). Predictors include predefined teleconnection indices, color-coded by ocean group: red for Pacific, green for Atlantic, and purple for Indian Ocean.</p>
Full article ">Figure 10
<p>Boxplots comparing the skill of MLR and RF models for predicting SPEI-2, SPEI-4, and SPEI-6. The upper panel shows the PCC, and the lower panel shows the RMSE. The bars are colored according to the model: MLR (green), RF (light blue). C.I. stands for climate index, a model using 10 climate indices, and C.I + ACF stands for autocorrelation function plus climate index, a model including 10 climate indices, and the SPEI’s one-month prior autocorrelation variable.</p>
Full article ">Figure 11
<p>Barplots of the ACF of December’s SPEI with lagged SPEIs occurring 1 to 11 months ahead. lag1 is the correlation between December and November of the SPEI value of that year, and lag2 is the correlation of December and October of the SPEI value of that year. lag11 is December and January of the SPEI of that year.</p>
Full article ">Figure 12
<p>Proficiency map of the HYBRID model for predicting drought in SPEI-2 (December). The left map (<b>a</b>) shows hit rate percentage, and the right map (<b>b</b>) shows HSS. The redder the map color for hit rates and HSS, the better the performance.</p>
Full article ">Figure 13
<p>Boxplots comparing the proficiency of MLR, RF, and HYBRID models for predicting SPEI-2, SPEI-4, and SPEI-6. The upper panel shows hit rate percentage box plots, and the lower panel shows Heidke skill score (HSS). Bars are colored according to the model: MLR (green), RF (light blue), HYBRID (light red).</p>
Full article ">
16 pages, 3823 KiB  
Article
Remote Sensing of Chlorophyll-a and Water Quality over Inland Lakes: How to Alleviate Geo-Location Error and Temporal Discrepancy in Model Training
by Jongmin Park, Sami Khanal, Kaiguang Zhao and Kyuhyun Byun
Remote Sens. 2024, 16(15), 2761; https://doi.org/10.3390/rs16152761 - 29 Jul 2024
Viewed by 508
Abstract
Harmful algal blooms (HABs) threaten lake ecosystems and public health. Early HAB detection is possible by monitoring chlorophyll-a (Chl-a) concentration. Ground-based Chl-a data have limited spatial and temporal coverage but can be geo-registered with temporally coincident satellite imagery to calibrate a remote sensing-based [...] Read more.
Harmful algal blooms (HABs) threaten lake ecosystems and public health. Early HAB detection is possible by monitoring chlorophyll-a (Chl-a) concentration. Ground-based Chl-a data have limited spatial and temporal coverage but can be geo-registered with temporally coincident satellite imagery to calibrate a remote sensing-based predictive model for regional mapping over time. When matching ground and satellite data, positional and temporal discrepancies are unavoidable due particularly to dynamic lake surfaces, thereby biasing the model calibration. This limitation has long been recognized but so far has not been addressed explicitly. To mitigate such effects of data mismatching, we proposed an Akaike Information Criterion (AIC)-like weighted regression algorithm that relies on an error-based heuristic to automatically favor “good” data points and downplay “bad” points. We evaluated the algorithm for estimating Chl-a over inland lakes in Ohio using Harmonized Landsat Sentinel-2. The AIC-like weighted regression estimates showed superior performance with an R2 of 0.91 and an error variance (σE2) of 0.29 μg/L, outperforming linear regression (R2 = 0.34, σE2 = 2.34 μg/L) and random forest (R2 = 0.82, σE2 = 0.92 μg/L). We also noticed the poorest performance occurred in the spring due to low reflectance variation in clear water and low Chl-a concentration. Our weighted regression scheme is adaptive and generically applicable. Future studies may adopt our scheme to tackle other remote sensing estimation problems (e.g., terrestrial applications) for alleviating the adverse effects of geolocation errors and temporal discrepancies. Full article
(This article belongs to the Special Issue Multi-Source Remote Sensing Data in Hydrology and Water Management)
Show Figures

Figure 1

Figure 1
<p>Geographic information of the study area. The blue dots represent the ground-based stations utilized in this study, and the dotted line delineates the boundary of the Hydrologic Unit Code (HUC)-4.</p>
Full article ">Figure 2
<p>Overall framework of the proposed AIC-like weighted linear regression model.</p>
Full article ">Figure 3
<p>(<b>a</b>) Temporal behavior of Chl-a concentration and (<b>b</b>) annually averaged concentration and standard deviation of Chl-a concentration over inland lakes in Ohio.</p>
Full article ">Figure 4
<p>Temporal variation in chlorophyll-a concentration over Lake Erie (<b>a</b>) from 2000 to 2009 and (<b>b</b>) 2010 to 2020.</p>
Full article ">Figure 5
<p>Scatterplot of Chl-a estimates from (<b>a</b>) multivariate regression and (<b>b</b>) AIC-like weighted regression.</p>
Full article ">Figure 6
<p>Seasonal scatterplot of observed and predicted Chl-a during (<b>a</b>) spring, (<b>b</b>) summer, and (<b>c</b>) autumn.</p>
Full article ">Figure 7
<p>Example of Chl-a estimation with AIC-like weighted scheme over spring period.</p>
Full article ">Figure 8
<p>Boxplot of weight depending on different spatial windows along with number of datasets used for (<b>a</b>) Lake Erie and (<b>b</b>) other inland lakes across Ohio.</p>
Full article ">Figure 9
<p>Boxplot of weight depending on different temporal windows for (<b>a</b>) Lake Erie and (<b>b</b>) other inland lakes across Ohio.</p>
Full article ">
21 pages, 4337 KiB  
Article
Optimizing Crop Spatial Structure to Improve Water Use Efficiency and Ecological Sustainability in Inland River Basin
by Zihan Wu, Sunxun Zhang, Baoying Shan, Fan Zhang and Xi Chen
Agronomy 2024, 14(8), 1645; https://doi.org/10.3390/agronomy14081645 - 27 Jul 2024
Viewed by 332
Abstract
Inland arid basins face the challenge of ecological deterioration due to insufficient water availability. The irrigation water consumption depletes the water flowing into the downstream tailrace ecological wetland, leading to increasing ecological deterioration. It is urgent to optimize the management of irrigation water [...] Read more.
Inland arid basins face the challenge of ecological deterioration due to insufficient water availability. The irrigation water consumption depletes the water flowing into the downstream tailrace ecological wetland, leading to increasing ecological deterioration. It is urgent to optimize the management of irrigation water resources in the middle reaches and improve the ecological sustainability of the lower reaches. To ensure sustainable development, improving water use efficiency and preserving the health of basin ecosystems should be simultaneously considered in the agricultural water management of these regions. Therefore, a 0–1 integer multi-objective programming approach was proposed to optimize midstream crop planting. This method has advantages in (1) effectively balancing ecological sustainability, agricultural production, and water-saving goals; (2) linking irrigation district management with grid geographic information to develop land use strategies; and (3) obtaining optimal solutions for multi-objective synergies. The proposed approach is applied to a typical inland river basin in China, the Heihe River Basin in Gansu Province. Results indicate that the optimization schemes can increase agricultural benefits, crop suitability, water use efficiency, and ecological quality by 12.37%, 6.82%, 13.00%, and 8.04% (compared to 2022), respectively, while irrigation water can be saved about 7.53%. The optimization results and proposed approach can help decision-makers manage water resources in the Heihe River Basin and similar regions. Full article
Show Figures

Figure 1

Figure 1
<p>Study framework.</p>
Full article ">Figure 2
<p>Geolocation of the study area.</p>
Full article ">Figure 3
<p>Distribution of suitability for cultivation of maize, wheat, and cash crops.</p>
Full article ">Figure 4
<p>Reference crop evapotranspiration distribution.</p>
Full article ">Figure 5
<p>Spatial and temporal distribution of MEQI in the Heihe River Basin from 2000 to 2022.</p>
Full article ">Figure 6
<p>The relationship between the ecological quality index of lower Heihe River and runoff from the Zhengyi Gorge.</p>
Full article ">Figure 7
<p>Comparison of model optimization results (<b>a</b>) and actual cropping structure (<b>b</b>) in 2022.</p>
Full article ">Figure 8
<p>Values (<b>a</b>) and proportions (<b>b</b>) of area shifts with the optimization model compared to 2022.</p>
Full article ">Figure 9
<p>Comparison of the optimized model with the base year indicators.</p>
Full article ">Figure 10
<p>Comparison of the performance of different models in terms of crop suitability (<b>a</b>), irrigation water use efficiency (<b>b</b>), and ecological quality (<b>c</b>).</p>
Full article ">
17 pages, 3445 KiB  
Article
Comparative Study of In Situ Chlorophyll-a Measuring Methods and Remote Sensing Techniques Focusing on Different Applied Algorithms in an Inland Lake
by János Grósz, Veronika Zsófia Tóth, István Waltner, Zoltán Vekerdy and Gábor Halupka
Water 2024, 16(15), 2104; https://doi.org/10.3390/w16152104 - 25 Jul 2024
Viewed by 387
Abstract
Water conservation efforts and studies receive special attention, versatile and constantly developing remote sensing methods especially so. The quality and quantity of algae fundamentally influence the ecosystems of water bodies. Inland lakes are less-frequently studied despite their essential ecological role compared to ocean [...] Read more.
Water conservation efforts and studies receive special attention, versatile and constantly developing remote sensing methods especially so. The quality and quantity of algae fundamentally influence the ecosystems of water bodies. Inland lakes are less-frequently studied despite their essential ecological role compared to ocean and sea waters. One of the reasons for this is the small-scale surface extension, which poses challenges during satellite remote sensing. In this study, we investigated the correlations between remote-sensing- (via Seninel-2 satellite) and laboratory-based results in different chlorophyll-a concentration ranges. In the case of low chlorophyll-a concentrations, the measured values were between 15 µg L−1 and 35 µg L−1. In the case of medium chlorophyll-a concentrations, the measured values ranged between 35 and 80 µg L−1. During high chlorophyll-a concentrations, the results were higher than 80 µg L−1. Finally, under extreme environmental conditions (algal bloom), the values were higher than 180 µg L−1. We also studied the accuracy and correlation and the different algorithms applied through the Acolite (20231023.0) processing software. The chl_re_mishra algorithm of the Acolite software gave the highest correlation. The strong positive correlations prove the applicability of the Sentinel-2 images and the Acolite software in the indication of chlorophyll-a. Because of the high CDOM concentration of Lake Naplás, the blue–green band ratio underestimated the concentration of chlorophyll-a. In summer, higher chlorophyll-a was detected in both laboratory and satellite investigations. In the case of extremely high chlorophyll-a concentrations, it is significantly underestimated by satellite remote sensing. This study proved the applicability of remote sensing to detect chlorophyll-a content but also pointed out the current limitations, thus assigning future development and research directions. Full article
Show Figures

Figure 1

Figure 1
<p>Locations of sampling points (source: Google Earth).</p>
Full article ">Figure 2
<p>Q-Q plot of chl_re_mishra algorithm.</p>
Full article ">Figure 3
<p>Q-Q plot of chl_re_moses3b740.</p>
Full article ">Figure 4
<p>Chlorophyll-a concentration (laboratory and Acolite results).</p>
Full article ">Figure 5
<p>Vertical distribution of phytoplankton in Case 1 (12 October 2017).</p>
Full article ">Figure 6
<p>Vertical distribution of phytoplankton in Case 2 (5 August 2022).</p>
Full article ">Figure 7
<p>Chlorophyll-a concentration at the sampling points (laboratory and satellite).</p>
Full article ">Figure 8
<p>Chlorophyll-a map (chl_re_mishra) on 5 August 2022.</p>
Full article ">Figure 9
<p>Chlorophyll-a map (chl_re_mishra) on 12 October 2017.</p>
Full article ">Figure 10
<p>Seasonal distribution of chlorophyll-a.</p>
Full article ">
18 pages, 5889 KiB  
Article
How Useful Are Moderate Resolution Imaging Spectroradiometer Observations for Inland Water Temperature Monitoring and Warming Trend Assessment in Temperate Lakes in Poland?
by Mariusz Sojka, Mariusz Ptak, Katarzyna Szyga-Pluta and Senlin Zhu
Remote Sens. 2024, 16(15), 2727; https://doi.org/10.3390/rs16152727 - 25 Jul 2024
Viewed by 360
Abstract
Continuous software development and widespread access to satellite imagery allow for obtaining increasingly accurate data on the natural environment. They play an important role in hydrosphere research, and one of the most frequently addressed issues in the era of climate change is the [...] Read more.
Continuous software development and widespread access to satellite imagery allow for obtaining increasingly accurate data on the natural environment. They play an important role in hydrosphere research, and one of the most frequently addressed issues in the era of climate change is the thermal dynamics of its components. Interesting research opportunities in this area are provided by the utilization of data obtained from the moderate resolution imaging spectroradiometer (MODIS). These data have been collected for over two decades and have already been used to study water temperature in lakes. In the case of Poland, there is a long history of studying the thermal regime of lakes based on in situ observations, but so far, MODIS data have not been used in these studies. In this study, the available products, such as 1-day and 8-day MODIS land surface temperature (LST), were validated. The obtained data were compared with in situ measurements, and the reliability of using these data to estimate long-term thermal changes in lake waters was also assessed. The analysis was conducted based on the example of two coastal lakes located in Poland. The results of 1-day LST MODIS generally showed a good fit compared to in situ measurements (average RMSE 1.9 °C). However, the analysis of long-term trends of water temperature changes revealed diverse results compared to such an approach based on field measurements. This situation is a result of the limited number of satellite data, which is dictated by environmental factors associated with high cloud cover reaching 60% during the analysis period. Full article
Show Figures

Figure 1

Figure 1
<p>Location of the study lakes.</p>
Full article ">Figure 2
<p>Minimum (blue), average (red), and maximum (green) in situ water temperatures of Łebsko (<b>a</b>) and Gardno (<b>b</b>) lakes from April to October during the years of 2003–2022.</p>
Full article ">Figure 3
<p>Daily water temperature deviations measured in situ from the values obtained from the MODIS sensor for Lake Łebsko.</p>
Full article ">Figure 4
<p>Daily water temperature deviations measured in situ from the values obtained from the MODIS sensor for Lake Gardno.</p>
Full article ">Figure 5
<p>Scatter plot of daily water temperatures measured in situ and 1 day from the MODIS sensor for Lake Łebsko.</p>
Full article ">Figure 6
<p>Scatter plot of daily water temperatures measured in situ and 1 day from the MODIS sensor for Lake Gardno.</p>
Full article ">Figure 7
<p>One-day LST values obtained from MODIS sensors compared to in situ measurements for Lake Łebsko for the period from April to October 2019.</p>
Full article ">Figure 8
<p>Average deviations between 1-day LST values recorded by MODIS on Aqua and Terra satellites during the day and night.</p>
Full article ">Figure 9
<p>The slope of the regression lines obtained from in situ measurements (blue color) and MODIS (red color) for Lakes Łebsko and Gardno over the years of 2003–2022.</p>
Full article ">Figure 9 Cont.
<p>The slope of the regression lines obtained from in situ measurements (blue color) and MODIS (red color) for Lakes Łebsko and Gardno over the years of 2003–2022.</p>
Full article ">
29 pages, 9666 KiB  
Article
Diatoms’ Diversity in the Assessment of the Impact of Diamond and Oil and Gas Mining on Aquatic Ecosystems of the Central Yakut Plain (Eastern Siberia, Yakutia) Using Bioindication and Statistical Mapping Methods
by Sophia Barinova, Viktor Gabyshev, Sergey Genkal and Olga Gabysheva
Diversity 2024, 16(8), 440; https://doi.org/10.3390/d16080440 - 24 Jul 2024
Viewed by 700
Abstract
Diamond and oil and gas production carries risks to the aquatic ecosystem. In Eastern Siberia, on the territory of the Central Yakut Plain, mining development of the Yakut diamond-bearing province and Tas-Yuryakh oil and gas condensate field has been underway for several decades. [...] Read more.
Diamond and oil and gas production carries risks to the aquatic ecosystem. In Eastern Siberia, on the territory of the Central Yakut Plain, mining development of the Yakut diamond-bearing province and Tas-Yuryakh oil and gas condensate field has been underway for several decades. But the problem of studying negative consequences in the region is covered only from the point of view of the impact on terrestrial ecosystems. The purpose of this study was to assess the impact of diamond and oil and gas production on the aquatic ecosystems of the region using the bioindicative properties of diatoms. The work used previously widely tested methods of ecological mapping, JASP, and species–environments relationship analysis. The results of chemical analysis of water showed that in oil and gas production areas, there is no pollution with petroleum products, but the concentration of silicon and zinc is increased. The study identified key pollutants in the Central Yakut Plain and demonstrated the effectiveness of diatoms as bioindicators. Elevated chloride levels were found in diamond mining areas, and increased copper levels were noted in oil production regions. In the diatom flora of the region, 144 species were identified, of which 137 are indicator species. Natural and anthropogenic clusters of environmental factors are identified, characterized by a specific effect on the species composition and taxonomic structure of diatom communities. The results obtained are suitable for assessing the level of anthropogenic impact on aquatic communities of photosynthetic microorganisms in permafrost conditions. Full article
(This article belongs to the Special Issue Freshwater Biodiversity Hotspots in 2024)
Show Figures

Figure 1

Figure 1
<p>Map with green dots indicating the sampling points on the studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia) and world map with a red point showing the geographic location of the study area. The violet-colored area is the territory of diamond mining. The red-colored area is the territory of oil and gas production. White dots with black outline show the settlements. Yellow lines show the highways. Blue arrows show the river flow direction.</p>
Full article ">Figure 2
<p>Natural landscape of sampling station areas. Aerial photo of taiga and Ulakhan-Murbayi River, swamp (station 7) (<b>a</b>). Satellite image of artificial water reservoir (station 6) (<b>b</b>). Aerial photo of Tustakh River, small stream (station 11) (<b>c</b>). Sampling process, swamp (station 10) (<b>d</b>). Satellite image of abandoned diamond quarry named after XXIII Party Congress (station 3) (<b>e</b>). Panoramic photo of abandoned nameless diamond quarry (station 4) (<b>f</b>).</p>
Full article ">Figure 3
<p>Bioindicators’ distribution in the 11 studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022: (<b>a</b>) habitat preferences (P—planktonic; P-B—plankto-benthic; B—benthic); (<b>b</b>) temperature (cool—cool water; temp—temperate; eterm—eurythermic; warm—warm water); (<b>c</b>) oxygen (st—standing water; str—streaming water; st-str—low streaming water; aer—aerophiles); (<b>d</b>) salinity (hb—oligohalobes-halophobes; i—oligohalobes-indifferent; hl—halophiles; mh—mesohalobes).</p>
Full article ">Figure 4
<p>Bioindicators’ distribution in the 11 studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022: (<b>a</b>) pH (alb—alkalibiontes; alf—alkaliphiles; ind—indifferent; acf—acidophiles; acb—acidobiontes); (<b>b</b>) organic pollution indicators according to T. Watanabe et al. [<a href="#B48-diversity-16-00440" class="html-bibr">48</a>] (sx—saproxenes; es—eurysaprobes; sp—saprophiles); (<b>c</b>) nitrogen uptake metabolism [<a href="#B15-diversity-16-00440" class="html-bibr">15</a>] (ats—nitrogen-autotrophic taxa, tolerating very small concentrations of organically bound nitrogen; ate—nitrogen-autotrophic taxa, tolerating elevated concentrations of organically bound nitrogen; hne—facultative nitrogen-heterotrophic taxa, needing periodically elevated concentrations of organically bound nitrogen); (<b>d</b>) trophic state indicators [<a href="#B15-diversity-16-00440" class="html-bibr">15</a>] (ot—oligotraphentic; om—oligomesotraphentic; m—mesotraphentic; me—mesoeutraphentic; e—eutraphentic; o-e—oligo to hypereutraphentic).</p>
Full article ">Figure 5
<p>Distribution of water quality class indicator species based on their species-specific index s related to class in the 11 studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure 6
<p>Statistical maps of the chemical variables’ distribution in the 11 studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022: (<b>a</b>) water temperature; (<b>b</b>) pH; (<b>c</b>) Cl; (<b>d</b>) Pt-Co; (<b>e</b>) N-NO<sub>3</sub>; (<b>f</b>) Cu.</p>
Full article ">Figure 7
<p>Statistical maps of the species number (<b>a</b>) and bioindicator variables; (<b>b</b>) eurythermic species; (<b>c</b>) alkaliphilic species; (<b>d</b>) mesohalobes; (<b>e</b>) oligo to hypereutraphentic species; (<b>f</b>) class 4 water quality distribution in the 11 studied waterbodies of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure 8
<p>Tree of similarity of the chemical and bioindicator variables in studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure 9
<p>Tree of similarity of the chemical variables in studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure 10
<p>Tree of similarity of the bioindicator variables in studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure 11
<p>JASP (Jeffreys’s Amazing Statistics Program) plots of similarities of total bioindicators and chemical variables (<b>a</b>) and chemistry only (<b>b</b>) in studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022. Bold lines show largest similarity in the type of analysis, “Huge” correlation &gt; 0.5.</p>
Full article ">Figure 12
<p>JASP (Jeffreys’s Amazing Statistics Program) plot of bioindicator variables only showing similarities in the studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022. Bold lines show largest similarity in the type of analysis, “Huge” correlation &gt; 0.5. Red lines—negative correlation; blue lines—positive correlation.</p>
Full article ">Figure 13
<p>RDA triplots of the species number, bioindicator groups, and environmental variables (<b>a</b>) and the taxonomic groups and environmental variables’ (<b>b</b>) relationships in the studied sites of the Central Yakut Plain (Eastern Siberia, Yakutia), 2022.</p>
Full article ">Figure A1
<p>Electron micrographs of diatoms of studied waterbodies: (<b>a</b>)—<span class="html-italic">Aulacoseira italica</span>, (<b>b</b>)—<span class="html-italic">Boreozonacola hustedtii</span>, (<b>c</b>)—<span class="html-italic">Brachysira brebissonii</span>, (<b>d</b>)—<span class="html-italic">Cyclotella distinguenda</span>, (<b>e</b>)—<span class="html-italic">Cymbella cymbiformis</span>, (<b>f</b>)—<span class="html-italic">Eunotia superbidens</span>, (<b>g</b>)—<span class="html-italic">Gyrosigma acuminatum</span>, (<b>h</b>)—<span class="html-italic">Gomphonema brebissonii</span>, (<b>i</b>)—<span class="html-italic">G. micropus</span>, (<b>j</b>)—<span class="html-italic">Lindavia comta</span>, (<b>k</b>)—<span class="html-italic">Pinnularia</span> cf. <span class="html-italic">globiceps var. linearis</span>, (<b>l</b>)—<span class="html-italic">P. ovata</span>, (<b>m</b>)—<span class="html-italic">P. sinistra</span>, (<b>n</b>)—<span class="html-italic">P. subrostrata</span>, (<b>o</b>)—<span class="html-italic">Rhoicosphenia abbreviata</span>, (<b>p</b>)—<span class="html-italic">Sellaphora pseudopupula</span>.</p>
Full article ">
Back to TopTop