[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

Search Results (169)

Search Parameters:
Keywords = CDOM

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
9 pages, 2084 KiB  
Article
Research Regarding the Autochthonous Dissolved Organic Carbon to Recalcitrant Dissolved Organic Carbon Transformation Mechanism in a Typical Surface Karst River
by Jiabin Li, Qiong Xiao, Qiufang He, Yurui Cheng, Fang Liu, Peiling Zhang, Yifei Liu, Daoxian Yuan and Shi Yu
Water 2024, 16(18), 2584; https://doi.org/10.3390/w16182584 - 12 Sep 2024
Viewed by 253
Abstract
Autochthonic recalcitrant organic carbon is the most stable component in karst aquatic systems. Still, the processes of its generation and transformation remain unclear, which hinders the study of the mechanisms and quantitative calculations of carbon sinks in karst aquatic systems. This study collected [...] Read more.
Autochthonic recalcitrant organic carbon is the most stable component in karst aquatic systems. Still, the processes of its generation and transformation remain unclear, which hinders the study of the mechanisms and quantitative calculations of carbon sinks in karst aquatic systems. This study collected water samples from the Li River, a typical surface karst river in Southwest China. Through in situ microbial cultivation and the chromophoric dissolved organic matter (CDOM) spectrum, changes in organic carbon components and their contents during the transformation of autochthonic dissolved organic carbon (Auto-DOC) to autochthonic dissolved recalcitrant organic carbon (Auto-RDOC) were analyzed to investigate the inert transformation processes of endogenous organic carbon. This study found that microbial carbon pumps (MCPs) promote the tyrosine-like component condensed into microbial-derived fulvic and humic components via heterotrophic bacteria metabolism, forming Auto-RDOC. During the dry season, the high level of Auto-DOC provides abundant organic substrates for heterotrophic bacteria, resulting in significantly higher Auto-RDOC production compared to the rainy season. This study provides fundamental information on the formation mechanisms of Auto-DOC in karst aquatic systems, which contributes to the assessment of carbon sinks in karst aquatic systems. Full article
Show Figures

Figure 1

Figure 1
<p>Geological and hydrological maps of Li River watershed and sampling sites.</p>
Full article ">Figure 2
<p>Seasonal fluorescence and absorption variations in the Auto-DOC to Auto-RDOC process.</p>
Full article ">Figure 3
<p>Seasonal CDOM fluorescence intensity variation in the Li River basin.</p>
Full article ">Figure 4
<p>Relative correlation heatmap of CDOM components and spectrum index.</p>
Full article ">
11 pages, 5748 KiB  
Article
The Influence of Groundwater Migration on Organic Matter Degradation and Biological Gas Production in the Central Depression of Qaidam Basin, China
by Jixian Tian, Qiufang He, Zeyu Shao and Fei Zhou
Water 2024, 16(15), 2163; https://doi.org/10.3390/w16152163 - 31 Jul 2024
Viewed by 522
Abstract
For insight into the productive and storage mechanisms of biogas in the Qaidam Basin, efforts were made to investigate the groundwater recharge and the processes of hydrocarbon generation by CDOM-EEM (fluorescence excitation-emission matrix of Chromophoric dissolved organic matter) spectrum, hydrogen and oxygen isotopes, [...] Read more.
For insight into the productive and storage mechanisms of biogas in the Qaidam Basin, efforts were made to investigate the groundwater recharge and the processes of hydrocarbon generation by CDOM-EEM (fluorescence excitation-emission matrix of Chromophoric dissolved organic matter) spectrum, hydrogen and oxygen isotopes, and geochemical characters in the central depression of the Qaidam Basin, China. The samples contain formation water from three gas fields (TN, SB, and YH) and surrounding surface water (fresh river and brine lake). The results indicate that modern precipitation significantly controls the salinity distribution and organic matter leaching in the groundwater system of the central depression of the Qaidam Basin. Higher salinity levels inhibit microbial activity, which leads to organic matter degradation and to gas generation efficiency being limited in the groundwater. The inhabitation effect is demonstrated by the notable negative correlation between the extent of organic matter degradation and its concentration with hydrogen and oxygen isotopes. The conclusion of this study indicated that modern precipitation emerges as a crucial factor affecting the biogas production and storage in the Qaidam Basin by influencing the ultimate salinity and organic matter concentration in the formation, which provides theoretical insight for the maintenance of modern gas production wells and the assessment of gas production potential. Full article
(This article belongs to the Section Hydrogeology)
Show Figures

Figure 1

Figure 1
<p>The schematic hydrogeology (<b>a</b>) and Quaternary Stratigraphic Column map (<b>b</b>) of Qaidam Basin.</p>
Full article ">Figure 2
<p>Stable isotope distribution of hydrogen and oxygen in the formation and surface water of central depression in Qaidam Basin (The data of brine lake and river water are cited from Li et al. [<a href="#B11-water-16-02163" class="html-bibr">11</a>]).</p>
Full article ">Figure 3
<p>The organic matter contents and CDOM index variation in the surface and formation water of Qaidam Basin.</p>
Full article ">Figure 4
<p>Fitting graphs of organic matter content in formation water, water isotopes, and CDOM spectral index in the central depression area of the Qaidam Basin; (<b>a</b>) Linear fit of total organic carbon (TOC) content with hydrogen (filled blue rhombus) and oxygen isotope (filled purple triangle) content; (<b>b</b>) Fitting of total organic carbon (TOC) content with humification index (HIX, filled green circle) and spectral index of aromaticity (SUV254, filled square).</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 486
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 ">
19 pages, 6891 KiB  
Article
Enhancing Machine Learning Performance in Estimating CDOM Absorption Coefficient via Data Resampling
by Jinuk Kim, Jin Hwi Kim, Wonjin Jang, JongCheol Pyo, Hyuk Lee, Seohyun Byeon, Hankyu Lee, Yongeun Park and Seongjoon Kim
Remote Sens. 2024, 16(13), 2313; https://doi.org/10.3390/rs16132313 - 25 Jun 2024
Viewed by 493
Abstract
Chromophoric dissolved organic matter (CDOM) is a mixture of various types of organic matter and a useful parameter for monitoring complex inland surface waters. Remote sensing has been widely utilized to detect CDOM in various studies; however, in many cases, the dataset is [...] Read more.
Chromophoric dissolved organic matter (CDOM) is a mixture of various types of organic matter and a useful parameter for monitoring complex inland surface waters. Remote sensing has been widely utilized to detect CDOM in various studies; however, in many cases, the dataset is relatively imbalanced in a single region. To address these concerns, data were acquired from hyperspectral images, field reflection spectra, and field monitoring data, and the imbalance problem was solved using a synthetic minority oversampling technique (SMOTE). Using the on-site reflectance ratio of the hyperspectral images, the input variables Rrs (452/497), Rrs (497/580), Rrs (497/618), and Rrs (684/618), which had the highest correlation with the CDOM absorption coefficient aCDOM (355), were extracted. Random forest and light gradient boosting machine algorithms were applied to create a CDOM prediction algorithm via machine learning, and to apply SMOTE, low-concentration and high-concentration datasets of CDOM were distinguished by 5 m−1. The training and testing datasets were distinguished at a 75%:25% ratio at low and high concentrations, and SMOTE was applied to generate synthetic data based on the training dataset, which is a sub-dataset of the original dataset. Datasets using SMOTE resulted in an overall improvement in the algorithmic accuracy of the training and test step. The random forest model was selected as the optimal model for CDOM prediction. In the best-case scenario of the random forest model, the SMOTE algorithm showed superior performance, with testing R2, absolute error (MAE), and root mean square error (RMSE) values of 0.838, 0.566, and 0.777 m−1, respectively, compared to the original algorithm’s test values of 0.722, 0.493, and 0.802 m−1. This study is anticipated to resolve imbalance problems using SMOTE when predicting remote sensing-based CDOM. It is expected to produce and implement a machine learning model with improved reliable performance. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Baekje reservoir (BJR) in the Geum River Basin and sampling points for each monitoring period.</p>
Full article ">Figure 2
<p>Airborne hyperspectral reflectance spectra of the sampling stations for seven campaigns in the Baekje reservoir (BJR).</p>
Full article ">Figure 3
<p>Scheme of the synthetic minority oversampling technique (SMOTE) application method to construct the random forest model [<a href="#B20-remotesensing-16-02313" class="html-bibr">20</a>].</p>
Full article ">Figure 4
<p>Distribution and histogram of CDOM data: (<b>a</b>) daily distribution of CDOM data; (<b>b</b>) histogram and section count of CDOM data and 5 m<sup>−1</sup>, which is the standard for class distinction, is indicated by a red line.</p>
Full article ">Figure 5
<p>R<sup>2</sup> heatmap by hyperspectral band ratio combinations (<span class="html-italic">X</span>-axis/<span class="html-italic">Y</span>-axis wavelength reflectance) versus <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mo>(</mo> <mn>355</mn> <mo>)</mo> </mrow> </semantics></math>. The red circle indicates a high R<sup>2</sup> region and shows the denominator/numerator wavelength of the highest R<sup>2</sup> value. The grey circle exhibits symmetry with the red circle and has a relatively lower R<sup>2</sup> value than that of the red circle.</p>
Full article ">Figure 6
<p>Correlation analysis between observed <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mo>(</mo> <mn>355</mn> <mo>)</mo> </mrow> </semantics></math> and simulated <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mo>(</mo> <mn>355</mn> <mo>)</mo> </mrow> </semantics></math> calculated using random forest: (<b>a</b>) training/testing selected as R<sup>2</sup> in the original dataset; (<b>b</b>) training/testing selected as MAE/RMSE in the original dataset; (<b>c</b>) training/testing selected as R<sup>2</sup> in the new dataset; (<b>d</b>) training/testing selected as MAE/RMSE in the new dataset. (<b>a</b>–<b>d</b>) are reclassified into Class 1 (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>355</mn> </mrow> </mfenced> <mo>&lt;</mo> <mn>5</mn> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>) and Class 2 (<math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mfenced separators="|"> <mrow> <mn>355</mn> </mrow> </mfenced> <mo>≥</mo> <mn>5</mn> <msup> <mrow> <mi mathvariant="normal">m</mi> </mrow> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </mrow> </semantics></math>), respectively, and the correlation and performance for each class are calculated and expressed as (<b>e</b>–<b>h</b>). The blue line represents the trend line in Train dataset, and the orange line represents the trend line in test dataset in (<b>a</b>–<b>d</b>). The red line represents the trend line in Class 2, and the green line represents the trend line in Class 1 in (<b>e</b>–<b>h</b>).</p>
Full article ">Figure 7
<p>Spatial distribution analysis of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mo>(</mo> <mn>355</mn> <mo>)</mo> </mrow> </semantics></math> at three points in the high-concentration section using hyperspectral imaging: hyperspectral images of (<b>a</b>) 12 August 2016, (<b>d</b>) 24 August 2016, and (<b>g</b>) 14 October 2016. (<b>b</b>,<b>e</b>,<b>h</b>) showed the CDOM spatial distribution constructed through the random forest algorithm from the original dataset, and (<b>c</b>,<b>f</b>,<b>i</b>) showed the CDOM spatial distribution constructed through the random forest algorithm from the new dataset.</p>
Full article ">Figure 8
<p>Distribution of data generated using SMOTE in the best-case scenario.</p>
Full article ">Figure 9
<p>Rainfall, temperature, and runoff time series data from 2016 to 2017 at the BJR and range, mean value, and standard deviation of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> </msub> <mo>(</mo> <mn>355</mn> <mo>)</mo> </mrow> </semantics></math> obtained from spatial distribution in sampling date.</p>
Full article ">
28 pages, 13381 KiB  
Article
Retrieval of Total Suspended Matter Concentration Based on the Iterative Analysis of Multiple Equations: A Case Study of a Lake Taihu Image from the First Sustainable Development Goals Science Satellite’s Multispectral Imager for Inshore
by Xueke Hu, Jiaguo Li, Yuan Sun, Yunfei Bao, Yonghua Sun, Xingfeng Chen and Yueguan Yan
Remote Sens. 2024, 16(8), 1385; https://doi.org/10.3390/rs16081385 - 14 Apr 2024
Cited by 1 | Viewed by 1011
Abstract
Inland waters consist of multiple concentrations of constituents, and solving the interference problem of chlorophyll-a and colored dissolved organic matter (CDOM) can help to accurately invert total suspended matter concentration (Ctsm). In this study, according to the characteristics of the [...] Read more.
Inland waters consist of multiple concentrations of constituents, and solving the interference problem of chlorophyll-a and colored dissolved organic matter (CDOM) can help to accurately invert total suspended matter concentration (Ctsm). In this study, according to the characteristics of the Multispectral Imager for Inshore (MII) equipped with the first Sustainable Development Goals Science Satellite (SDGSAT-1), an iterative inversion model was established based on the iterative analysis of multiple linear regression to estimate Ctsm. The Hydrolight radiative transfer model was used to simulate the radiative transfer process of Lake Taihu, and it analyzed the effect of three component concentrations on remote sensing reflectance. The characteristic band combinations B6/3 and B6/5 for multiple linear regression were determined using the correlation of the three component concentrations with different bands and band combinations. By combining the two multiple linear regression models, a complete closed iterative inversion model for solving Ctsm was formed, which was successfully verified by using the modeling data (R2 = 0.97, RMSE = 4.89 g/m3, MAPE = 11.48%) and the SDGSAT-1 MII image verification data (R2 = 0.87, RMSE = 3.92 g/m3, MAPE = 8.13%). And it was compared with iterative inversion models constructed based on other combinations of feature bands and other published models. Remote sensing monitoring Ctsm was carried out using SDGSAT-1 MII images of Lake Taihu in 2022–2023. This study can serve as a technical reference for the SDGSAT-1 satellite in terms of remote sensing monitoring of Ctsm, as well as monitoring and improving the water environment. Full article
(This article belongs to the Special Issue Remote Sensing Band Ratios for the Assessment of Water Quality)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area and measured sample point distribution.</p>
Full article ">Figure 2
<p>Absorption spectrum.</p>
Full article ">Figure 3
<p>Scattering spectrum.</p>
Full article ">Figure 4
<p>SDGSAT-1 MII spectral response function.</p>
Full article ">Figure 5
<p>The simulated <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> is fitted with the measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>: (<b>a</b>) in the 400–800 nm and (<b>b</b>) at band center wavelength.</p>
Full article ">Figure 6
<p>The measured concentration data of 54 sample points and the correlation coefficient between the simulated <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and the measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> for 54 sample points.</p>
Full article ">Figure 7
<p><math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> under single-component concentration change. (<b>a</b>–<b>c</b>) are <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> obtained under different <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi mathvariant="italic">cdom</mi> </mrow> </msub> <mrow> <mo>(</mo> <mn>440</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>d</b>–<b>f</b>) are <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> obtained under different <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi mathvariant="italic">cdom</mi> </mrow> </msub> <mrow> <mo>(</mo> <mn>440</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p><math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> under single-component concentration change, λ = 448 nm, 509 nm, 569 nm, 668 nm, and 773 nm. (<b>a</b>–<b>c</b>) are <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> obtained under different <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi mathvariant="italic">cdom</mi> </mrow> </msub> <mrow> <mo>(</mo> <mn>440</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>; (<b>d</b>–<b>f</b>) are <math display="inline"><semantics> <mrow> <msubsup> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> <mrow> <mi mathvariant="sans-serif">Δ</mi> </mrow> </msubsup> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> obtained under different <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math>, and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>a</mi> </mrow> <mrow> <mi mathvariant="italic">cdom</mi> </mrow> </msub> <mrow> <mo>(</mo> <mn>440</mn> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure 9
<p>Bands and band combinations that are not related to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">cdom</mi> </mrow> </msub> </mrow> </semantics></math> but to <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 10
<p>The relationship between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and the single-component remote sensing reflectance contributed by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>: (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>3</mn> </mrow> </mrow> </msub> <mo>)</mo> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msup> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>5</mn> </mrow> </mrow> </msub> <mo>)</mo> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 11
<p>The relationship between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msub> </mrow> </semantics></math> and the single-component remote sensing reflectance contributed by <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>: (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>3</mn> </mrow> </mrow> </msub> <mo>)</mo> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msup> </mrow> </semantics></math> and (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <mo>(</mo> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>5</mn> </mrow> </mrow> </msub> <mo>)</mo> </mrow> <mrow> <mi mathvariant="italic">chla</mi> </mrow> </msup> </mrow> </semantics></math>.</p>
Full article ">Figure 12
<p>The change in the difference between each set of output value <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mo>(</mo> <mi>m</mi> <mrow> <mo>+</mo> <mn>1</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> and input value <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mo>(</mo> <mi>m</mi> <mo>)</mo> </mrow> </msup> </mrow> </semantics></math> during iteration.: (<b>a</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 1 g/m<sup>3</sup>; (<b>b</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 10 g/m<sup>3</sup>; (<b>c</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 50 g/m<sup>3</sup>; (<b>d</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 100 g/m<sup>3</sup>; (<b>e</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 200 g/m<sup>3</sup>; and (<b>f</b>) <math display="inline"><semantics> <mrow> <msup> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> <mrow> <mrow> <mo>(</mo> <mn>0</mn> <mo>)</mo> </mrow> </mrow> </msup> </mrow> </semantics></math> = 500 g/m<sup>3</sup>.</p>
Full article ">Figure 13
<p>The change in RE in the process of iterative convergence.</p>
Full article ">Figure 14
<p>The measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> are fitted to the retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> of the iterative model.</p>
Full article ">Figure 15
<p>The measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> is fitted to the retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> from the SDGSAT-1 MII image, based on the iterative model.</p>
Full article ">Figure 16
<p>Change in the difference between the output value and the input value during the iteration process: (<b>a</b>) based on <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mn>6</mn> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>5</mn> </mrow> </mrow> </msub> </mrow> </semantics></math>; (<b>b</b>) based on <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>2</mn> </mrow> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>5</mn> </mrow> </mrow> </msub> </mrow> </semantics></math>; and (<b>c</b>) based on <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>4</mn> </mrow> </mrow> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mn>6</mn> <mo>/</mo> <mn>5</mn> </mrow> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure 17
<p>Fitting of measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> when models are applied to the modeled data.</p>
Full article ">Figure 18
<p>Fitting of measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> when models are applied to the SDGSAT-1 MII image of 27 July 2022.</p>
Full article ">Figure 19
<p>Fitting of measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> when the adapted models are applied to the modeled data [<a href="#B18-remotesensing-16-01385" class="html-bibr">18</a>,<a href="#B47-remotesensing-16-01385" class="html-bibr">47</a>,<a href="#B48-remotesensing-16-01385" class="html-bibr">48</a>].</p>
Full article ">Figure 20
<p>Fitting of measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> and retrieved <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> when the adapted models are applied to the SDGSAT-1 MII image of 27 July 2022 [<a href="#B18-remotesensing-16-01385" class="html-bibr">18</a>,<a href="#B47-remotesensing-16-01385" class="html-bibr">47</a>,<a href="#B48-remotesensing-16-01385" class="html-bibr">48</a>].</p>
Full article ">Figure 21
<p>Spatial distribution of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math> in Lake Taihu, 2022–2023.</p>
Full article ">Figure 22
<p>Quarterly statistics of the proportion of pixels in each range of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>C</mi> </mrow> <mrow> <mi mathvariant="italic">tsm</mi> </mrow> </msub> </mrow> </semantics></math>.</p>
Full article ">Figure A1
<p>Comparison between simulated <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> and measured <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>. (<b>a1</b>–<b>a8</b>), (<b>b1</b>–<b>b8</b>), (<b>c1</b>–<b>c8</b>), (<b>d1</b>–<b>d8</b>), (<b>e1</b>–<b>e8</b>), (<b>f1</b>–<b>f8</b>), and (<b>g1</b>–<b>g6</b>) are the comparisons of 54 groups of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi mathvariant="italic">rs</mi> </mrow> </msub> <mrow> <mo>(</mo> <mi mathvariant="sans-serif">λ</mi> <mo>)</mo> </mrow> </mrow> </semantics></math>.</p>
Full article ">Figure A2
<p>Correlation analysis results: (<b>a</b>) correlation between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>i</mi> </mrow> </msub> </mrow> </semantics></math> and the concentrations of the three components; (<b>b</b>) correlation between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mi>i</mi> <mo>/</mo> <mi>j</mi> </mrow> </msub> </mrow> </semantics></math> and the concentrations of the three components; and (<b>c</b>) correlation between <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>B</mi> </mrow> <mrow> <mrow> <mi>i</mi> <mo> </mo> </mrow> <mtext>-</mtext> <mi>j</mi> </mrow> </msub> </mrow> </semantics></math> and the concentrations of the three components.</p>
Full article ">
16 pages, 4706 KiB  
Article
Drought Offsets the Controls on Colored Dissolved Organic Matter in Lakes
by Enass Said. Al-Kharusi, Geert Hensgens, Abdulhakim M. Abdi, Tiit Kutser, Jan Karlsson, David E. Tenenbaum and Martin Berggren
Remote Sens. 2024, 16(8), 1345; https://doi.org/10.3390/rs16081345 - 11 Apr 2024
Viewed by 914
Abstract
The concentration of colored dissolved organic matter (CDOM) in lakes is strongly influenced by climate, land cover, and topographic settings, but it is not known how drought may affect the relative importance of these controls. Here, we evaluate the controls of CDOM during [...] Read more.
The concentration of colored dissolved organic matter (CDOM) in lakes is strongly influenced by climate, land cover, and topographic settings, but it is not known how drought may affect the relative importance of these controls. Here, we evaluate the controls of CDOM during two summers with strongly contrasting values of the Palmer drought index (PDI), indicating wet vs. dry conditions. We hypothesized that lake CDOM during a wet summer season is regulated mainly by the surrounding land cover to which the lakes are hydrologically connected, while, during drought, the lakes are disconnected from the catchment and CDOM is regulated by climatic and morphometric factors that govern the internal turnover of CDOM in the lakes. A suite of climate, land cover, and morphometric variables was assembled and used to explain remotely sensed CDOM values for 255 boreal lakes distributed across broad environmental and geographic gradients in Sweden and Norway. We found that PDI explained the variability in CDOM among lakes in a dry year, but not in a wet year, and that severe drought strongly decreased CDOM during the dry year. Large lakes, especially, with a presumed high degree of catchment uncoupling, showed low CDOM during the dry year. However, in disagreement with our hypothesis, climate, land cover, and morphometry all showed a stronger impact on lake CDOM in wet vs. dry years. Thus, drought systematically weakened the predictability of CDOM variations at the same time as CDOM was offset toward lower values. Our results show that drought not only has a direct effect on CDOM, but also acts indirectly by changing the spatial regulation of CDOM in boreal lakes. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Figure 1

Figure 1
<p>Locations of the study sites in Sweden and Norway. Sentinel-2 image tiles are highlighted in gray.</p>
Full article ">Figure 2
<p>Relationship between ln-transformed CDOM (m<sup>−1</sup>) and Palmer drought index (y = 0.1677x + 0.0652; R<sup>2</sup> = 0.11; <span class="html-italic">n</span> = 255; 2-tail <span class="html-italic">p</span> &lt; 0.0001) for all regions in the dry summer season of 2018 (solid line). In the wet year of 2016, there was no significant correlation, hence a dashed line showing the average value. Note that nearby lakes can have identical PDI values due to the course spatial resolution of the downloaded PDI data.</p>
Full article ">Figure 3
<p>Variance partitioning of ln(CDOM) for the categories of catchment morphometry, land cover, and climate (16, 11, and 3 variables, respectively) shown for (<b>A</b>) 2016 and (<b>B</b>) 2018 (see list of variables in <a href="#app1-remotesensing-16-01345" class="html-app">Table S3</a>). Total adjusted R<sup>2</sup> is 0.55 for 2016 and 0.41 for 2018. Partitioning of adjusted R<sup>2</sup> is shown in the graph. Insignificant values are not shown.</p>
Full article ">Figure 4
<p>(<b>A</b>) Performance and (<b>B</b>) variables of importance in a PLS model for natural logarithm of CDOM (m<sup>−1</sup>) during the <span class="html-italic">wet year of 2016</span>. The model explained 60% of the variance in CDOM, using 76% of the variance in the predictor variables (<a href="#remotesensing-16-01345-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 5
<p>(<b>A</b>) Performance and (<b>B</b>) variables of importance in a PLS model for ln-transformed CDOM (m<sup>−1</sup>) regressed from predictor variables in the <span class="html-italic">dry year of 2018</span>. The model explained 43% of the variance in CDOM with using 71% of the data variance of the predictor variables. See explanation of selected variables in <a href="#remotesensing-16-01345-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 6
<p>(<b>A</b>) Correlation between ln-transformed CDOM (m<sup>−1</sup>) and air temperature (°C) in the 2016 wet year (y = 0.150x−0.804; R<sup>2</sup> = 0.29; <span class="html-italic">p</span> &lt; 0.0001) and 2018 dry year (y = 0.118x−1.397; R<sup>2</sup> = 0.09; <span class="html-italic">p</span> &lt; 0.0001). (<b>B</b>) Correlation between ln-transformed CDOM (m<sup>−1</sup>) and ln-transformed lake area (m<sup>2</sup>) for the (2018) dry summer (y = −0.117x + 1.495; R² = 0.10). (<b>C</b>) Correlation between ln-transformed CDOM (m<sup>−1</sup>) and relative catchment cover of coniferous forest area in (2016) wet season (y = 0.664x + 0.576, R<sup>2</sup> = 0.19; <span class="html-italic">p</span> &lt; 0.0001).</p>
Full article ">Figure 7
<p>(<b>A</b>,<b>B</b>,<b>D</b>,<b>E</b>) The spatial distribution of the strength of the coefficients (temperature and SDTI) and the significance of the spatial points are shown for the respective variables in the wet year 2016. (<b>C</b>,<b>F</b>) The spatial distribution of the strength of the coefficients lake area and the significance of the spatial points are shown for the dry year 2018.</p>
Full article ">
25 pages, 133890 KiB  
Article
Characterizing Water Composition with an Autonomous Robotic Team Employing Comprehensive In Situ Sensing, Hyperspectral Imaging, Machine Learning, and Conformal Prediction
by John Waczak, Adam Aker, Lakitha O. H. Wijeratne, Shawhin Talebi, Ashen Fernando, Prabuddha M. H. Dewage, Mazhar Iqbal, Matthew Lary, David Schaefer and David J. Lary
Remote Sens. 2024, 16(6), 996; https://doi.org/10.3390/rs16060996 - 12 Mar 2024
Cited by 2 | Viewed by 1227
Abstract
Inland waters pose a unique challenge for water quality monitoring by remote sensing techniques due to their complicated spectral features and small-scale variability. At the same time, collecting the reference data needed to calibrate remote sensing data products is both time consuming and [...] Read more.
Inland waters pose a unique challenge for water quality monitoring by remote sensing techniques due to their complicated spectral features and small-scale variability. At the same time, collecting the reference data needed to calibrate remote sensing data products is both time consuming and expensive. In this study, we present the further development of a robotic team composed of an uncrewed surface vessel (USV) providing in situ reference measurements and an unmanned aerial vehicle (UAV) equipped with a hyperspectral imager. Together, this team is able to address the limitations of existing approaches by enabling the simultaneous collection of hyperspectral imagery with precisely collocated in situ data. We showcase the capabilities of this team using data collected in a northern Texas pond across three days in 2020. Machine learning models for 13 variables are trained using the dataset of paired in situ measurements and coincident reflectance spectra. These models successfully estimate physical variables including temperature, conductivity, pH, and turbidity as well as the concentrations of blue–green algae, colored dissolved organic matter (CDOM), chlorophyll-a, crude oil, optical brighteners, and the ions Ca2+, Cl, and Na+. We extend the training procedure to utilize conformal prediction to estimate 90% confidence intervals for the output of each trained model. Maps generated by applying the models to the collected images reveal small-scale spatial variability within the pond. This study highlights the value of combining real-time, in situ measurements together with hyperspectral imaging for the rapid characterization of water composition. 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>Configuration of the USV: (<b>a</b>) Frontal view of the USV showing the Eureka Manta + 40 multiprobes mounted on the underside of the boat. (<b>b</b>) The USV deployed in the water.</p>
Full article ">Figure 2
<p>Configuration of the UAV: (<b>a</b>) The hyperspectral imager and acquisition computer. (<b>b</b>) The assembled UAV with secondary processing computer and (upward facing) downwelling irradiance spectrometer.</p>
Full article ">Figure 3
<p>Hyperspectral image processing: Hyperspectral data cubes are collected one scan-line at a time (<b>left</b>). By utilizing downwelling irradiance spectra, we convert each pixel from spectral radiance to reflectance. By using orientation and position data from the on-board GPS and INS, we georeference each pixel to assign it a latitude and longitude on the ground. The final data product is the georectified hyperspectral reflectance data cube (<b>right</b>) visualized as a pseudo-color image with reflectance as a function of wavelength along the z-axis.</p>
Full article ">Figure 4
<p>A georectified reflectance data cube is visualized (center) with the <math display="inline"><semantics> <msub> <mo form="prefix">log</mo> <mn>10</mn> </msub> </semantics></math> reflectance along the z-axis and a pseudo-color image on the top. In the top left, we visualize the downwelling irradiance spectrum (the incident light). The surrounding plots showcase exemplar pixel reflectance spectra for open water, dry grass, algae, and a rhodamine dye plume used to test the system.</p>
Full article ">Figure 5
<p>The pond in Montague, Texas, where the robot team was deployed. The pond includes multiple distinct regions separated by small islands and grasses.</p>
Full article ">Figure 6
<p>Distribution of total downwelling intensity during each of the three HSI collection flights. The multi-modal nature of these distributions reflects the impact of the relative orientation of the drone to the sun as well as potential occlusion due to the presence of clouds.</p>
Full article ">Figure 7
<p>Scatter diagrams (<b>left</b>) and quantile–quantile plots (<b>right</b>) for the hyperparameter-optimized RFR models for the physical variables measured by the USV.</p>
Full article ">Figure 8
<p>Ranked permutation importance for each feature in the physical variable models. Permutation importance measured the decrease in the model’s <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math> value after replacing each feature in the prediction set with a random permutation of its values.</p>
Full article ">Figure 9
<p>Maps generated by applying each of the physical variable models to the hyperspectral data cubes collected on 23 November. Overlaid over the predictions are color-filled squares showing the associated in situ reference data for the same collection period. The size of the squares has been exaggerated for visualization. We note that there is good agreement between the model predictions and the reference data.</p>
Full article ">Figure 10
<p>Scatter diagrams (<b>left</b>) and quantile–quantile plots (<b>right</b>) for the hyperparameter-optimized RFR models for the ion measurements made by the USV.</p>
Full article ">Figure 11
<p>Ranked permutation importance for the top 25 features of the ion models. The permutation importance measures the decrease in the model’s <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math> value when each feature is replaced by a random permutation of its values.</p>
Full article ">Figure 12
<p>Maps generated by applying the trained ion models to the data cubes collected on 23 November. Overlaid on the maps are the in situ reference measurements for the same collection period. The size of the squares has been exaggerated for the visualization. We note that there is good agreement between the generated map and the reference data.</p>
Full article ">Figure 13
<p>Scatter plots (<b>left</b>) and quantile–quantile plots (<b>right</b>) for the final hyperparameter-optimized models for the biochemical targets blue–green algae (phycoerythrin), CDOM, chlorophyll-a, and blue–green algae (phycocyanin).</p>
Full article ">Figure 14
<p>Ranked permutation importance for each feature in the trained biochemical models. The permutation importance measures the decrease in the model’s <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math> value after replacing each feature with a random permutation of its values.</p>
Full article ">Figure 15
<p>Maps generated by applying the trained biochemical models to the data cubes collected on 23 November. Overlaid are the in situ reference data for the same collection period. The size of the squares has been exaggerated for the visualization. We note there is good agreement between the predicted map and the reference data.</p>
Full article ">Figure 16
<p>Scatter diagrams (<b>left</b>) and quantile–quantile plots (<b>right</b>) for the hyperparameter-optimized RFR models for the chemical variables measured by the USV.</p>
Full article ">Figure 17
<p>Ranked permutation importance for the top 25 features of the chemical models. The permutation importance measures the decrease in the model’s <math display="inline"><semantics> <msup> <mi>R</mi> <mn>2</mn> </msup> </semantics></math> value after replacing each feature in the prediction set with a random permutation of its values.</p>
Full article ">Figure 18
<p>Maps generated by applying the trained chemical variable models to the hyperspectral data cubes collected on 23 November. Overlaid are color-filled squares showing the in situ reference data for the same collection period. The size of the squares is exaggerated for the visualization. We note that there is good agreement between the model predictions and reference data.</p>
Full article ">
19 pages, 2875 KiB  
Article
Estimating the Colored Dissolved Organic Matter in the Negro River, Amazon Basin, with In Situ Remote Sensing Data
by Rogério Ribeiro Marinho, Jean-Michel Martinez, Tereza Cristina Souza de Oliveira, Wagner Picanço Moreira, Lino A. Sander de Carvalho, Patricia Moreira-Turcq and Tristan Harmel
Remote Sens. 2024, 16(4), 613; https://doi.org/10.3390/rs16040613 - 6 Feb 2024
Viewed by 1314
Abstract
Dissolved organic matter (DOM) is a crucial component of continental aquatic ecosystems. It plays a vital role in the carbon cycle by serving as a significant source and reservoir of carbon in water. DOM provides energy and nutrients to organisms, affecting primary productivity, [...] Read more.
Dissolved organic matter (DOM) is a crucial component of continental aquatic ecosystems. It plays a vital role in the carbon cycle by serving as a significant source and reservoir of carbon in water. DOM provides energy and nutrients to organisms, affecting primary productivity, organic composition, and the food chain. This study presents empirical bio-optical models for estimating the absorption of colored dissolved organic matter (aCDOM) in the Negro River using in situ remote sensing reflectance (Rrs) data. Physical–chemical data (TSS, DOC, and POC) and optical data (aCDOM and Rrs) were collected from the Negro River, its tributaries, and lakes and empirical relationships between aCDOM at 440 nm, single band, and the ratio bands of Rrs were assessed. The analysis of spectral slope shows no statistically significant correlations with DOC concentration or aCDOM absorption coefficient. However, strong relationships were observed between DOC and aCDOM (R2 = 0.72), aCDOM and Rrs at 650 nm (R2 > 0.80 and RMSE < 1.75 m−1), as well as aCDOM and the green/red band ratio (R2 > 0.80 and RMSE < 2.30 m−1). aCDOM displayed large spatial and temporal variations, varying from 1.9 up to 20.1 m−1, with higher values in rivers of the upper course of the Negro basin and lower values in rivers with total solids suspended > 10 mg·L−1. Environmental factors that influence the production of dissolved organic matter include soil type, dense forest cover, high precipitation, and low erosion rates. This study demonstrated that aCDOM can serve as an indicator of DOC, and Rrs can serve as an indicator of aCDOM in the Negro basin. Our findings offer a starting point for future research on the optical properties of Amazonian black-water rivers. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of the study area and sampling sections.</p>
Full article ">Figure 2
<p>Examples of Rrs collected in different sections of the Negro River (black-water), its tributaries (black- and clear-water), and the Amazon River (white-water). (<b>a</b>) High-water period; (<b>b</b>) low-water period.</p>
Full article ">Figure 3
<p>aCDOM of the Negro River (<b>a</b>) and tributaries (<b>b</b>). Note that the <span class="html-italic">Y</span>-axis scale in (<b>a</b>) is different from the <span class="html-italic">Y</span>-axis scale in <a href="#remotesensing-16-00613-f003" class="html-fig">Figure 3</a>b. Refer to <a href="#remotesensing-16-00613-t001" class="html-table">Table 1</a> for the names of the sections in (<b>b</b>).</p>
Full article ">Figure 4
<p>DOC versus aCDOM in the Negro River Basin. The dashed black line indicates a 1:1 relationship. Please refer to <a href="#remotesensing-16-00613-t001" class="html-table">Table 1</a> for the names of the sections.</p>
Full article ">Figure 5
<p>Relationship between aCDOM at 440 nm and the ratio of Rrs for MSI bands simulated at 560 nm and 665 nm in the high-water period (<b>a</b>) and low-water period (<b>b</b>). The blue lines are 95% confidence intervals.</p>
Full article ">Figure 6
<p>The relationship between aCDOM at 440 nm and the Rrs for MSI band simulated at 665 nm for the high-water period (<b>a</b>) and low-water period (<b>b</b>). The blue lines represent the 95% confidence intervals.</p>
Full article ">Figure 7
<p>Spatial distribution of aCDOM at 440 nm in the Negro basin, Amazon River, and the coast of French Guiana. Source: This study and GLORIA [<a href="#B73-remotesensing-16-00613" class="html-bibr">73</a>].</p>
Full article ">
28 pages, 4379 KiB  
Article
Estimation of the Biogeochemical and Physical Properties of Lakes Based on Remote Sensing and Artificial Intelligence Applications
by Kaire Toming, Hui Liu, Tuuli Soomets, Evelyn Uuemaa, Tiina Nõges and Tiit Kutser
Remote Sens. 2024, 16(3), 464; https://doi.org/10.3390/rs16030464 - 25 Jan 2024
Cited by 2 | Viewed by 1467
Abstract
Lakes play a crucial role in the global biogeochemical cycles through the transport, storage, and transformation of different biogeochemical compounds. Their regulatory service appears to be disproportionately important relative to their small areal extent, necessitating continuous monitoring. This study leverages the potential of [...] Read more.
Lakes play a crucial role in the global biogeochemical cycles through the transport, storage, and transformation of different biogeochemical compounds. Their regulatory service appears to be disproportionately important relative to their small areal extent, necessitating continuous monitoring. This study leverages the potential of optical remote sensing sensors, specifically Sentinel-2 Multispectral Imagery (MSI), to monitor and predict water quality parameters in lakes. Optically active parameters, such as chlorophyll a (CHL), total suspended matter (TSM), and colored dissolved matter (CDOM), can be directly detected using optical remote sensing sensors. However, the challenge lies in detecting non-optically active substances, which lack direct spectral characteristics. The capabilities of artificial intelligence applications can be used in the identification of optically non-active compounds from remote sensing data. This study aims to employ a machine learning approach (combining the Genetic Algorithm (GA) and Extreme Gradient Boost (XGBoost)) and in situ and Sentinel-2 Multispectral Imagery data to construct inversion models for 16 physical and biogeochemical water quality parameters including CHL, CDOM, TSM, total nitrogen (TN), total phosphorus (TP), phosphate (PO4), sulphate, ammonium nitrogen, 5-day biochemical oxygen demand (BOD5), chemical oxygen demand (COD), and the biomasses of phytoplankton and cyanobacteria, pH, dissolved oxygen (O2), water temperature (WT) and transparency (SD). GA_XGBoost exhibited strong predictive capabilities and it was able to accurately predict 10 biogeochemical and 2 physical water quality parameters. Additionally, this study provides a practical demonstration of the developed inversion models, illustrating their applicability in estimating various water quality parameters simultaneously across multiple lakes on five different dates. The study highlights the need for ongoing research and refinement of machine learning methodologies in environmental monitoring, particularly in remote sensing applications for water quality assessment. Results emphasize the need for broader temporal scopes, longer-term datasets, and enhanced model selection strategies to improve the robustness and generalizability of these models. In general, the outcomes of this study provide the basis for a better understanding of the role of lakes in the biogeochemical cycle and will allow the formulation of reliable recommendations for various applications used in the studies of ecology, water quality, the climate, and the carbon cycle. Full article
Show Figures

Figure 1

Figure 1
<p>Study area, 45 lakes of the input data for the GA_XGBoost model (Lakes (In situ), blue dots); and 180 Estonian lakes (&gt;0.1 km<sup>2</sup>), whose biogeochemical and physical water quality parameters were retrieved using the GA_XGBoost models and Sentinel-2 data (Lakes (Sentinel-2), red dots).</p>
Full article ">Figure 2
<p>Heatmap of Pearson correlations between biogeochemical and physical water quality parameters. Statistically significant correlations (<span class="html-italic">p</span>-value &lt; 0.05) are colored either red (positive) or blue (negative), while correlations that were not significant (<span class="html-italic">p</span> &gt; 0.05) are marked as grey.</p>
Full article ">Figure 3
<p>The mean Sentinel-2 MSI atmospherically corrected reflectance spectra sorted by the trophic state of study lakes. Sentinel-2 data are derived from each match-up point. Thick lines show the mean value, and the semitransparent area shows the standard error (±) of the mean.</p>
Full article ">Figure 4
<p>Scatter plots of training, test, and validation datasets produced using the best GA_XGB_model for deriving biogeochemical and physical water quality parameters from Sentinel-2 data along with the ideal model (1:1 line). The figure starts on the previous page.</p>
Full article ">Figure 4 Cont.
<p>Scatter plots of training, test, and validation datasets produced using the best GA_XGB_model for deriving biogeochemical and physical water quality parameters from Sentinel-2 data along with the ideal model (1:1 line). The figure starts on the previous page.</p>
Full article ">Figure 5
<p>The boxplots of the mean values of chlorophyll a (CHL, µg/L), colored dissolved organic matter (CDOM, mg/L), total suspended matter (TSM, mg/L), total nitrogen (TN, mg/L), total phosphorus (TP, mg/L), PO<sub>4</sub> (mg/L), TN:TP ratio, BOD<sub>5</sub> (mg O<sub>2</sub>/L), COD (mg O<sub>2</sub>/L), pH, Secchi depth (SD, m), water temperature (WT, C°), and O<sub>2</sub> (mg/L) in 180 Estonian lakes &gt; 0.1 km<sup>2</sup> on five different dates using Sentinel-2 data. On the plots the line indicates the median, the circle is the mean, the box shows the interquartile range, and the upper and lower whiskers are the maximum and minimum, respectively.</p>
Full article ">
20 pages, 14821 KiB  
Article
Estimation of Dissolved Organic Carbon Using Sentinel-2 in the Eutrophic Lake Ebinur, China
by Naixin Cao, Xingwen Lin, Changjiang Liu, Mou Leong Tan, Jingchao Shi, Chi-Yung Jim, Guanghui Hu, Xu Ma and Fei Zhang
Remote Sens. 2024, 16(2), 252; https://doi.org/10.3390/rs16020252 - 9 Jan 2024
Cited by 1 | Viewed by 1344
Abstract
Dissolved organic carbon (DOC) in lakes, as a regulatory agent and light-absorbing compound, is a key component of the global carbon cycling in lacustrine ecosystems. Hence, continuous monitoring of the DOC concentration in arid regions is extremely important. This study utilizes the QAA-CDOM [...] Read more.
Dissolved organic carbon (DOC) in lakes, as a regulatory agent and light-absorbing compound, is a key component of the global carbon cycling in lacustrine ecosystems. Hence, continuous monitoring of the DOC concentration in arid regions is extremely important. This study utilizes the QAA-CDOM semi-analytical model, which has good accuracy in retrieving the CDOM (colored dissolved organic matter) concentration of Lake Ebinur. We chose to invert the CDOM time-series data from May to October during the 2018–2022 period. A DOC estimation model was then established using the linear regression approach based on the CDOM inversion data and the field DOC measurements. In general, the DOC concentration in Lake Ebinur exhibited an increasing trend from 2018 to 2022, typically lower in May and higher in June. When comparing the average values of DOC in Lake Ebinur for the same months across different years, it can be observed that the month of September exhibits the greatest variability, whereas June shows the least variability. In sum, this study successfully retrieved CDOM concentrations for a saline lake within an arid region and developed a DOC estimation model, thereby providing a reference for investigating carbon cycling in typical lakes of arid areas. Full article
(This article belongs to the Section Ecological Remote Sensing)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Xinjiang, China, (<b>b</b>) sampling points over the main lake area, (<b>c</b>) average wind direction and speed in Lake Ebinur from 2014 to 2022, and (<b>d</b>) field sampling weather condition statistics.</p>
Full article ">Figure 2
<p>Flowchart of this study.</p>
Full article ">Figure 3
<p>Absorption coefficients of (<b>a</b>) CDOM and (<b>b</b>) SPM.</p>
Full article ">Figure 4
<p>Comparison between the absorption coefficients of dewatering and pure water ((<b>a</b>) represents the absorption coefficient of dewatering, (<b>b</b>) represents a comparison between the absorption coefficient of pure water and that of non-water, where the blue curve represents the average value of the measured absorption coefficient of non-water).</p>
Full article ">Figure 5
<p>CDOM inversion model validation of the QAA-CDOM.</p>
Full article ">Figure 6
<p>Accuracy verification of DOC estimation model ((<b>a</b>) linear regression model, (<b>b</b>) quadratic regression model, (<b>c</b>) exponential regression model, (<b>d</b>) logarithmic regression model).</p>
Full article ">Figure 7
<p>Monthly DOC inversion results for Lake Ebinur from 2018 to 2022.</p>
Full article ">Figure 8
<p>Interannual DOC changes in Lake Ebinur from 2018 to 2021 (considering that the area of Lake Ebinur varies greatly, the minimum lake area for different months in the same year was selected as the boundary when the annual average data of Lake Ebinur from 2018 to 2021 were synthesized).</p>
Full article ">Figure 9
<p>Seasonal DOC changes in Lake Ebinur from 2018 to 2021 (considering the large seasonal variation in Lake Ebinur’s area, the minimum lake area of the same month in different years was selected as the boundary when compiling the monthly average data of Lake Ebinur from 2018 to 2021).</p>
Full article ">Figure 10
<p>Average monthly DOC statistics of Lake Ebinur from 2018 to 2022. ((<b>a</b>) Line chart of the monthly DOC concentration in Lake Ebinur; (<b>b</b>) box chart of the average DOC concentration in Lake Ebinur in different months; (<b>c</b>) variation coefficient of DOC in different regions from 2018 to 2022; (<b>d</b>) variation coefficient of DOC in different regions from May to October.).</p>
Full article ">Figure 11
<p>Accuracy verification of linear regression model.</p>
Full article ">
23 pages, 8649 KiB  
Article
Links between Land Cover and In-Water Optical Properties in Four Optically Contrasting Swedish Bays
by Susanne Kratzer and Martin Allart
Remote Sens. 2024, 16(1), 176; https://doi.org/10.3390/rs16010176 - 31 Dec 2023
Cited by 1 | Viewed by 921
Abstract
The optical complexity of coastal waters is mostly caused by the water discharged from land carrying optical components (such as dissolved and particulate matter) into coastal bays and estuaries, and increasing the attenuation of light. This paper aims to investigate the links between [...] Read more.
The optical complexity of coastal waters is mostly caused by the water discharged from land carrying optical components (such as dissolved and particulate matter) into coastal bays and estuaries, and increasing the attenuation of light. This paper aims to investigate the links between in-water optical properties in four Swedish bays (from the northern Baltic proper up to the Bothnian bay) and the land use and land cover (LULC) in the respective catchment of each bay. The optical properties were measured in situ over the last decade by various research and monitoring groups while the LULC in each bay was classified using the Copernicus Land Monitoring Service based on Landsat 8/OLI data. The absorption coefficient of colored dissolve organic matter (CDOM) at 440 nm, aCDOM (440), was significantly correlated to Wetlands which may act as sources of CDOM, while Developed areas (Agricultural and Urban classes) were negatively correlated. The Agriculture class was also negatively related to suspended particulate organic matter (SPOM), whilst Coniferous Forests and Mixed Forests as well as Meadows were positively correlated. SPOM seems thus to mostly originate from Natural classes, possibly due to the release of pollen and other organic matter. Overall, the methods applied here allow for a better understanding of effects of land use and land cover on the bio-optical properties, and thus coastal water quality, on a macroscopic scale. Full article
(This article belongs to the Special Issue Oceans from Space V)
Show Figures

Figure 1

Figure 1
<p>Overview of the Baltic Sea and its location within Europe. Zones I and II are areas of interest for this study. Source: Baltic Sea catchment area and HELCOM Sub-basins 2018 [<a href="#B20-remotesensing-16-00176" class="html-bibr">20</a>] Country boundaries: Natural Earth [<a href="#B21-remotesensing-16-00176" class="html-bibr">21</a>], European coastline shapefile: EEA [<a href="#B22-remotesensing-16-00176" class="html-bibr">22</a>], SMHI’s sub-basin division (havs-områden_SVAR_2016 [<a href="#B23-remotesensing-16-00176" class="html-bibr">23</a>]). LD refers to Landsort Deep, the deepest part in the Baltic Sea (459 m).</p>
Full article ">Figure 2
<p>The four Swedish bays, their sub-basins and sampling stations (different coloured circles in different bays) (<b>a</b>) Bråviken, (<b>b</b>) Himmerfjärden<b>,</b> (<b>c</b>)<b>,</b> Östhammar (<b>d</b>) Råneå. Note (<b>a</b>–<b>c</b>) are situated in Zone I (see also <a href="#remotesensing-16-00176-f001" class="html-fig">Figure 1</a> and <a href="#remotesensing-16-00176-f003" class="html-fig">Figure 3</a>a) and (<b>d</b>) is situated in Zone II (<a href="#remotesensing-16-00176-f001" class="html-fig">Figure 1</a> and <a href="#remotesensing-16-00176-f003" class="html-fig">Figure 3</a>b). Maps generated in QGIS using predefined shapefiles (European coastline shapefile [<a href="#B22-remotesensing-16-00176" class="html-bibr">22</a>]), SMHI’s sub-basin division (havs-områden_SVAR_2016 [<a href="#B23-remotesensing-16-00176" class="html-bibr">23</a>]). Recurrence layer—Global Surface Water, Surface elevation—Copernicus EU-DEM- v1.1; Sentinel Hub [<a href="#B25-remotesensing-16-00176" class="html-bibr">25</a>]).</p>
Full article ">Figure 3
<p>Water course network and catchments of (<b>a</b>) Zone I: Bråviken bay (light and dark green), Himmerfjärden bay (light and dark purple), Östhammar bay (orange); (<b>b</b>) Råneå bay, located in Zone II in Swedish Lapland (see also <a href="#remotesensing-16-00176-f001" class="html-fig">Figure 1</a> for situation of Zones I and II). Maps generated in QGIS using predefined shapefiles (European coastline shapefile [<a href="#B22-remotesensing-16-00176" class="html-bibr">22</a>]). Recurrence layer—Global Surface Water, Surface elevation—Copernicus EU-DEM- v1.1; Sentinel Hub [<a href="#B25-remotesensing-16-00176" class="html-bibr">25</a>]), Water course network [<a href="#B26-remotesensing-16-00176" class="html-bibr">26</a>].</p>
Full article ">Figure 4
<p>Workflow diagram of the data processing method.</p>
Full article ">Figure 5
<p>Changes in the monthly discharge averaged over the years 1991–2021 for (<b>a</b>) Himmerfärden and Östhammar bays, and (<b>b</b>) for Råneå and Bråviken bay. The values were derived from the SMHI interactive map (SMHI, Hydrologiskt Nuläge, Vattenwebb—in English: Current Hydrological Status, Water Web), available online [<a href="#B26-remotesensing-16-00176" class="html-bibr">26</a>] by summing up all the contributions of the smaller watersheds and greater catchments contributing to the discharge in the considered bay.</p>
Full article ">Figure 6
<p>Image of the closer Himmerfjärden catchment area before and after LULC analysis. Maps generated in QGIS using predefined shapefiles (European coastline shapefile [<a href="#B22-remotesensing-16-00176" class="html-bibr">22</a>]. Recurrence layer—Global Surface Water, Surface elevation—Copernicus EU-DEM- v1.1, downloaded from the Sentinel Hub [<a href="#B25-remotesensing-16-00176" class="html-bibr">25</a>]; Water course network (vattendragslinjer nätverk): SVAR (Svenskt Vat-tenARkiv [<a href="#B26-remotesensing-16-00176" class="html-bibr">26</a>]).</p>
Full article ">Figure 7
<p>Landcover maps for the greater catchment areas of (<b>a</b>) the Råneå bay, (<b>b</b>) the Östhammar bay<b>,</b> (<b>c</b>) the Himmerfjärden bay, and for (<b>d</b>) the Bråviken bay. Note that the same figures are shown enlarged in the <a href="#app1-remotesensing-16-00176" class="html-app">Supplementary Figure S2a–d</a> (1 map per page). This allows for a more detailed assessment of the LULC in the greater catchment of each bay.</p>
Full article ">Figure 8
<p>CDOM absorption at 440 nm, <span class="html-italic">a<sub>CDOM</sub></span>, in four Swedish bays: (<b>a</b>) Råneå bay, (<b>b</b>) Östhammar bay (<b>c</b>) Bråviken bay and (<b>d</b>) Himmerfärden bay. Values are plotted against horizontal distance from the outlet (in km). All transects were plotted within each bay regardless of the season. The different colors in each plot refer to different days for the respective in situ transects.</p>
Full article ">Figure 9
<p>(<b>a</b>) CDOM absorption at 440 nm, a<span class="html-italic"><sub>CDOM</sub></span>, against the LULC category <span class="html-italic">Ratio*</span> (<span class="html-italic">Dev/Nat)</span>, i.e., <span class="html-italic">the</span> ratio of developed to natural classes; all transects for all seasons. (<b>b</b>) Averaged slope factor per transects against <span class="html-italic">Ratio*</span> (<span class="html-italic">Dev/Nat</span>) per bay; Råneå (Rå), Östhammar (Östh), Bråviken (Br) and Himmerfjärden (Hfj)—all transects for all seasons.</p>
Full article ">Figure 10
<p>(<b>a</b>) CDOM absorption at 440 nm, <span class="html-italic">a<sub>CDOM</sub></span>, against the LULC category <span class="html-italic">Ratio*</span> (<span class="html-italic">Dev/Nat)</span>, i.e., <span class="html-italic">the</span> ratio of developed to natural classes; average value per transect and bay in summer, and (<b>b</b>) the average value per transect and bay in spring; Råneå (Rå), Östhammar (Östh), Bråviken (Br) and Himmerfjärden (Hfj).</p>
Full article ">Figure 11
<p>Chl<span class="html-italic">-a</span> concentration in four Swedish bays—Råneå (Rå), Östhammar (Östh), Bråviken (BR) and Himmerfjärden (Hfj)—against the ratio of <span class="html-italic">Developed</span> to <span class="html-italic">Natural</span> (<span class="html-italic">Dev/Nat</span>) LULC.</p>
Full article ">
20 pages, 3676 KiB  
Article
Improving Colored Dissolved Organic Matter (CDOM) Retrievals by Sentinel2-MSI Data through a Total Suspended Matter (TSM)-Driven Classification: The Case of Pertusillo Lake (Southern Italy)
by Emanuele Ciancia, Alessandra Campanelli, Roberto Colonna, Angelo Palombo, Simone Pascucci, Stefano Pignatti and Nicola Pergola
Remote Sens. 2023, 15(24), 5718; https://doi.org/10.3390/rs15245718 - 13 Dec 2023
Cited by 1 | Viewed by 1100
Abstract
Colored dissolved organic matter (CDOM) is a significant constituent of aquatic systems and biogeochemical cycles. Satellite CDOM retrievals are challenging in inland waters, due to overlapped absorption properties of bio-optical parameters, like Total Suspended Matter (TSM). In this framework, we defined an accurate [...] Read more.
Colored dissolved organic matter (CDOM) is a significant constituent of aquatic systems and biogeochemical cycles. Satellite CDOM retrievals are challenging in inland waters, due to overlapped absorption properties of bio-optical parameters, like Total Suspended Matter (TSM). In this framework, we defined an accurate CDOM model using Sentinel2-MSI (S2-MSI) data in Pertusillo Lake (Southern Italy) adopting a classification scheme based on satellite TSM data. Empirical relationships were established between the CDOM absorption coefficient, aCDOM (440), and reflectance band ratios using ground-based measurements. The Green-to-Red (B3/B4 and B3/B5) and Red-to-Blue (B4/B2 and B5/B2) band ratios showed good relationships (R2 ≥ 0.75), which were further improved according to sub-region division (R2 up to 0.93). The best accuracy of B3/B4 in the match-ups between S2-MSI-derived and in situ band ratios proved the exportability on S2-MSI data of two B3/B4-based aCDOM (440) models, namely the fixed (for the whole PL) and the switching one (according to sub-region division). Although they both exhibited good agreements in aCDOM (440) retrievals (R2 ≥ 0.69), the switching model showed the highest accuracy (RMSE of 0.0155 m−1). Finally, the identification of areas exposed to different TSM patterns can assist with refining the calibration/validation procedures to achieve more accurate aCDOM (440) retrievals. Full article
Show Figures

Figure 1

Figure 1
<p>On the left is the PL location and on the right is the magnification of the study area (reported in the WGS 84 Coordinate Reference System (CRS)). The main rivers/tributaries (continuous black lines) and the bathymetry (in blue tones) are depicted. The villages (black triangles) and the onshore oil field (i.e., Centro Olio Val d’Agri) close to PL are also reported.</p>
Full article ">Figure 2
<p>Map of the PL sub-regions derived from the ISODATA unsupervised classification (adapted from Ciancia et al. [<a href="#B40-remotesensing-15-05718" class="html-bibr">40</a>]). Dots represent the locations of the sampling stations designed for the in situ measurement campaigns. The yellow dots (from S1 to S6) are those falling into the PL western side while the red ones (from S7 to S10) are within the PL eastern side.</p>
Full article ">Figure 3
<p>Scatter plot of estimated versus measured a<sub>CDOM</sub> (440) for (<b>a</b>) the empirical [<a href="#B14-remotesensing-15-05718" class="html-bibr">14</a>] and (<b>b</b>) the semi-analytical [<a href="#B27-remotesensing-15-05718" class="html-bibr">27</a>] algorithms. The regression and 1:1 lines are depicted by the continuous and dashed lines, respectively.</p>
Full article ">Figure 4
<p>Most performing calibration models (in terms of R<sup>2</sup> and RMSE) for the whole PL (<b>a</b>), the West (<b>b</b>), and East (<b>c</b>) subsets, respectively. The equations in the plots represent the mathematical “best-fitting” functions (dashed red line).</p>
Full article ">Figure 5
<p>S2-MSI band ratios versus in situ ones for (<b>a</b>) B3/B4, (<b>b</b>) B3/B5, (<b>c</b>) B4/B2, and (<b>d</b>) B5/B2. Note that the analysis was performed without distinction between sub-regions (i.e., West and East). The regression and 1:1 lines are depicted by the continuous and dashed lines, respectively.</p>
Full article ">Figure 6
<p>S2-MSI a<sub>CDOM</sub> (440) maps of 14 June 2017 for the CDOM fixed model (<b>a</b>) and the switching one (<b>b</b>), respectively. The black and red dots are the validation sampling stations falling into the West and East subsets, respectively.</p>
Full article ">Figure 7
<p>S2-MSI a<sub>CDOM</sub> (440) maps of 12 October 2017 for the CDOM fixed model (<b>a</b>) and the switching one (<b>b</b>), respectively. The black and red dots are the validation sampling stations falling into the West and East subsets, respectively.</p>
Full article ">Figure 8
<p>Validation match-ups of S2-MSI a<sub>CDOM</sub> (440) models based on a fixed PL-tuned algorithm (<b>a</b>) and on a switching PL-tuned one (<b>b</b>). The regression and 1:1 lines are depicted by the continuous and dashed lines, respectively.</p>
Full article ">
19 pages, 5868 KiB  
Article
Remote Sensing Estimation of CDOM and DOC with the Environmental Implications for Lake Khanka
by Sining Qiang, Kaishan Song, Yingxin Shang, Fengfa Lai, Zhidan Wen, Ge Liu, Hui Tao and Yunfeng Lyu
Remote Sens. 2023, 15(24), 5707; https://doi.org/10.3390/rs15245707 - 13 Dec 2023
Viewed by 1308
Abstract
Chromophoric dissolved organic matter (CDOM) is a significant contributor to the biogeochemical cycle and energy dynamics within aquatic ecosystems. Hence, the implementation of a systematic and comprehensive monitoring and governance framework for the CDOM in inland waters holds significant importance. This study conducted [...] Read more.
Chromophoric dissolved organic matter (CDOM) is a significant contributor to the biogeochemical cycle and energy dynamics within aquatic ecosystems. Hence, the implementation of a systematic and comprehensive monitoring and governance framework for the CDOM in inland waters holds significant importance. This study conducted the retrieval of CDOM in Lake Khanka. Specifically, we use the GBDT (R2 = 0.84) algorithm which performed best in retrieving CDOM levels and an empirical relationship based on the situ data between CDOM and dissolved organic carbon (DOC) to indicate the distribution of DOC indirectly. The performance of the CDOM-DOC retrieval scheme was reasonably good, achieving an R2 value of 0.69. The empirical algorithms were utilized for the analysis of Sentinel-3 datasets from the period 2016 to 2020 in Lake Khanka. The potential factors that contributed to the sources of DOM were also analyzed with the humification index (HIX). The significant relationship between CDOM and DOC (HIX and chemical oxygen demand (COD)) indicated the potential remote sensing application of water quality monitoring for water management. An analysis of our findings suggests that the water quality of the Great Khanka is superior to that of the Small Khanka. Moreover, the distribution of diverse organic matter exhibits a pattern where concentrations are generally higher along the shoreline compared to the center of the lake. Efficient measures should be promptly implemented to safeguard the water resources in international boundary lakes such as Lake Khanka and comprehensive monitoring systems including DOM distribution, DOM sources, and water quality management would be essential for water resource protection and government management. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>–<b>e</b>) Pseudo-color displays Sentinel-3 OLCI images showing the landscape and sample point locations of the study area, (<b>f</b>,<b>g</b>) Lake Khanka, and surrounding drainage basin in the world.</p>
Full article ">Figure 2
<p>Correlation heat map of lake environmental water chemical parameters.</p>
Full article ">Figure 3
<p>Modeling and validating scatterplots of eXtreme Gradient Boosting.</p>
Full article ">Figure 4
<p>Modeling and validating scatterplots of Support Vector Regression.</p>
Full article ">Figure 5
<p>Modeling and validating scatterplots of Back Propagation.</p>
Full article ">Figure 6
<p>Modeling and validating scatterplots of Random Forest.</p>
Full article ">Figure 7
<p>Modeling and validating scatterplots of Gradient Boosting Decision Tree.</p>
Full article ">Figure 8
<p>Construction and verification of CDOM absorption coefficient and DOC concentration model.</p>
Full article ">Figure 9
<p>Annual mean value fluctuation and reserves of DOC in lake Khanka during 2016–2022.</p>
Full article ">Figure 10
<p>Inversion results of DOC concentration distribution in lake Khanka during 2016–2022.</p>
Full article ">Figure 11
<p>Contribution rate of natural factors (PRCP, WDSP, STP, and TEMP) to DOC changes in 1 day and 30 days in multiple regression.</p>
Full article ">Figure 12
<p>CDOM and HIX average values in the inversion results of Lake Khanka images in chronological order.</p>
Full article ">
24 pages, 50912 KiB  
Article
Estimation of Water Quality Parameters in Oligotrophic Coastal Waters Using Uncrewed-Aerial-Vehicle-Obtained Hyperspectral Data
by Morena Galešić Divić, Marija Kvesić Ivanković, Vladimir Divić, Mak Kišević, Marko Panić, Predrag Lugonja, Vladimir Crnojević and Roko Andričević
J. Mar. Sci. Eng. 2023, 11(10), 2026; https://doi.org/10.3390/jmse11102026 - 22 Oct 2023
Cited by 2 | Viewed by 1392
Abstract
Water quality monitoring in coastal areas and estuaries poses significant challenges due to the intricate interplay of hydrodynamic, chemical, and biological processes, regardless of the chosen monitoring methods. In this study, we analyzed the applicability of different monitoring sources using in situ data, [...] Read more.
Water quality monitoring in coastal areas and estuaries poses significant challenges due to the intricate interplay of hydrodynamic, chemical, and biological processes, regardless of the chosen monitoring methods. In this study, we analyzed the applicability of different monitoring sources using in situ data, uncrewed-aerial-vehicle (UAV)-mounted hyperspectral sensing, and Sentinel-2-based multispectral imagery. In the first part of the study, we evaluated the applicability of existing empirical algorithms for water quality (WQ) parameter retrieval using hyperspectral, simulated multispectral, and satellite multispectral datasets and in situ measurements. In particular, we focused on three optically active WQ parameters: chlorophyll a (Chl,a), turbidity (TUR), and colored dissolved organic matter (CDOM) in oligotrophic coastal waters. We observed that most existing algorithms performed poorly when applied to different reflectance datasets, similar to previous findings in small and optically complex water bodies. Hence, we proposed a novel set of locally based empirical algorithms tailored for determining water quality parameters, which constituted the second part of our study. The newly developed regression-based algorithms utilized all possible combinations of spectral bands derived from UAV-generated hyperspectral data and exhibited coefficients of determination exceeding 0.9 for the three considered WQ parameters. The presented two-part approach was demonstrated in the semi-enclosed area of Kaštela Bay and the Jadro River estuary in the Central Eastern Adriatic Sea. This study introduces a promising and efficient screening method for UAV-based water quality monitoring in coastal areas worldwide. Such an approach may support decision-making processes related to coastal management and ultimately contribute to the conservation of coastal water ecosystems. Full article
(This article belongs to the Special Issue Remote Sensing Techniques in Marine Environment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Flowchart of implemented approach.</p>
Full article ">Figure 2
<p>Case study location.</p>
Full article ">Figure 3
<p>Representation for three reflectance datasets used in this study which illustrates the varying spectral density between multispectral (MS) and hyperspectral (HS) bands.</p>
Full article ">Figure 4
<p>Seasonally grouped box–whisker plots depicting measured chlorophyll <math display="inline"><semantics> <mrow> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math>), turbidity (TUR), and colored dissolved organic matter (CDOM) at the Jadro River estuary.</p>
Full article ">Figure 5
<p>Temporal and spatial variations of chlorophyll <math display="inline"><semantics> <mrow> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math>), turbidity (TUR), and colored dissolved organic matter (CDOM) obtained in field campaigns.</p>
Full article ">Figure 6
<p>Hyperspectral reflectances at measurement points, each season distinguished by a unique color palette and line style.</p>
Full article ">Figure 7
<p>Comparison of the <math display="inline"><semantics> <msubsup> <mi>θ</mi> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> <mn>8</mn> </msubsup> </semantics></math> algorithm performance derived using (<b>a</b>) UAV HS, (<b>b</b>) UAV MS, and (<b>c</b>) SAT MS datasets to in situ chlorophyll <math display="inline"><semantics> <mrow> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math>) measurements.</p>
Full article ">Figure 8
<p>Comparison of the <math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>θ</mi> <mo>˜</mo> </mover> <mrow> <mi>T</mi> <mi>U</mi> <mi>R</mi> </mrow> <mn>4</mn> </msubsup> </semantics></math> algorithm performance derived using (<b>a</b>) UAV HS, (<b>b</b>) UAV MS, and (<b>c</b>) SAT MS datasets to in situ turbidity (TUR) measurements.</p>
Full article ">Figure 9
<p>Comparison of the <math display="inline"><semantics> <msubsup> <mover accent="true"> <mi>θ</mi> <mo>˜</mo> </mover> <mrow> <mi>C</mi> <mi>D</mi> <mi>O</mi> <mi>M</mi> </mrow> <mn>2</mn> </msubsup> </semantics></math> algorithm performance derived using (<b>a</b>) UAV HS, (<b>b</b>) UAV MS, and (<b>c</b>) SAT MS datasets to in situ colored dissolved organic matter (CDOM) measurements.</p>
Full article ">Figure 10
<p>Best performing local algorithm for chlorophyll <math display="inline"><semantics> <mrow> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math> (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math>) retrieval.</p>
Full article ">Figure 11
<p>Best performing local algorithm for turbidity (TUR) retrieval.</p>
Full article ">Figure 12
<p>Best-performing local algorithm for colored dissolved organic matter (CDOM) retrieval.</p>
Full article ">Figure 13
<p>Chlorophyll-a (<math display="inline"><semantics> <mrow> <mi>C</mi> <mi>h</mi> <mi>l</mi> <mspace width="0.166667em"/> <mi>a</mi> </mrow> </semantics></math>) map and its distribution over the UAV polygon for all field campaigns. (<b>a</b>) Summer, (<b>b</b>) autumn, (<b>c</b>) winter, (<b>d</b>) spring.</p>
Full article ">
22 pages, 9627 KiB  
Article
Regional Models for Sentinel-2/MSI Imagery of Chlorophyll a and TSS, Obtained for Oligotrophic Issyk-Kul Lake Using High-Resolution LIF LiDAR Data
by Vadim Pelevin, Ekaterina Koltsova, Aleksandr Molkov, Sergei Fedorov, Salmor Alymkulov, Boris Konovalov, Mairam Alymkulova and Kubanychbek Jumaliev
Remote Sens. 2023, 15(18), 4443; https://doi.org/10.3390/rs15184443 - 9 Sep 2023
Viewed by 1144
Abstract
The development of regional satellite bio-optical models for natural waters with high temporal and spatial variability, such as inland seas, reservoirs, and coastal ocean waters, requires the implementation of an intermediate measuring link in the chain, “water sampling—bio-optical models”, and this link must [...] Read more.
The development of regional satellite bio-optical models for natural waters with high temporal and spatial variability, such as inland seas, reservoirs, and coastal ocean waters, requires the implementation of an intermediate measuring link in the chain, “water sampling—bio-optical models”, and this link must have certain intermediate characteristics. The most crucial of them are the high-precision measurements of the main water quality parameters, such as the concentration of chlorophyll a (Chl a), colored dissolved organic matter (CDOM), and total suspended sediments (TSS) in the upper water layer, together with a high operational rate and the ability to cover a large water area in a short time, which corresponds to a satellite overpass. A possible solution is to utilize laser-induced fluorescence (LIF) of water constituents measured by a marine LiDAR in situ with a high sampling rate from a high-speed vessel. This allows obtaining a large ground-truth dataset of the main water quality parameters simultaneously with the satellite overpass within the time interval determined by NASA protocols. This method was successfully applied to the oligotrophic Issyk-Kul Lake in Kyrgyzstan, where we obtained more than 4000 and 1000 matchups for the Chl a and TSS, respectively. New preliminary regional bio-optical models were developed on the basis of a one-day survey and tested for archive Sentinel-2A data for 2022. This approach can be applied for regular monitoring and further correction in accordance with seasonal variability. The obtained results, together with previously published similar studies for eutrophic coastal and productive inland waters, emphasize the applicability of the presented method for the development or adjustment of regional bio-optical models for water bodies of a wide trophic range. Full article
Show Figures

Figure 1

Figure 1
<p>Issyk-Kul Lake: (<b>a</b>) location on the world map (NASA “Blue Marble” image (<a href="https://visibleearth.nasa.gov/collection/1484/blue-marble" target="_blank">https://visibleearth.nasa.gov/collection/1484/blue-marble</a>, accessed on 1 August 2023) and (<b>b</b>) the satellite RGB-composite image (Sentinel-3B, accessed on 17 July 2022).</p>
Full article ">Figure 2
<p>Map of Issyk-Kul Lake (<a href="https://matplotlib.org/basemap/" target="_blank">https://matplotlib.org/basemap/</a>, accessed on 20 December 2022) with schematized bathymetry contours of 200, 300, 400, and 500 m [<a href="#B7-remotesensing-15-04443" class="html-bibr">7</a>]. The red dots denote the deepest areas. The green dashed lines in the eastern part of the lake indicate the palaeochanels of the Jyrgalan and Tyup Rivers. The dark blue line in the eastern part of the lake is a ship path.</p>
Full article ">Figure 3
<p>Photos of the studied region from a motorboat: (<b>a</b>) highly turbid waters of Jyrgalan River, (<b>b</b>) Tyup River, and (<b>c</b>) clean waters of the open lake between Jyrgalan and Tyup mouths.</p>
Full article ">Figure 4
<p>The ultraviolet fluorescence LiDAR UFL-9 on the motorboat.</p>
Full article ">Figure 5
<p>Results of comparison of LiDAR signals at stations and water samples analyzed in the laboratory according to concentrations of (<b>a</b>) Chl <span class="html-italic">a</span> and (<b>b</b>) TSS. Black lines correspond to the best calibration matches (1) and (2), respectively.</p>
Full article ">Figure 6
<p>The RGB-composite image of Issyk-Kul Lake: (<b>a</b>) basic map with the motorboat path divided into water zones with river outflows (blue (1), dark blue (6), and light green (7) colors), shore water zones (red (2), green (3), and orange (5) colors), and zone with deep water (purple (4) color), yellow dots represent stations, (<b>b</b>) map with filtered data for Chl <span class="html-italic">a</span> estimation, and (<b>c</b>) map with filtered data for TSS estimation. The white satellite logo represents the moment of the Sentinel-2A overpass (11:36 a.m. local time (UTC+6)).</p>
Full article ">Figure 7
<p>Examples of <math display="inline"><semantics> <mrow> <msub> <mrow> <mi>R</mi> </mrow> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> </mrow> </semantics></math> spectra measured in Lake Issyk-Kul on 16 July 2022: (<b>a</b>) spectra included in the temporal-synchronous dataset for calibration /validation of Chl <span class="html-italic">a</span> and TSS models; and (<b>b</b>) spectra excluded from the analysis.</p>
Full article ">Figure 8
<p>Results of (<b>a</b>) calibration (N = 2057) and (<b>b</b>) validation (N = 2054) of OC2 algorithm with coefficients a = [−0.3103; −2.3868; −3.7861; 17.8521; −13.8532] for Chl <span class="html-italic">a</span> in Issyk-Kul Lake. The colors of the dots correspond to the zones shown in <a href="#remotesensing-15-04443-f006" class="html-fig">Figure 6</a>.</p>
Full article ">Figure 9
<p>Results of (<b>a</b>) calibration (N = 536) and (<b>b</b>) validation (N = 535) of linear B3 algorithm (TSS = 22.02⋅B3 + 0.41) for TSS in Issyk-Kul Lake. The dashed line represents the 1:1 line.</p>
Full article ">Figure 10
<p>Results of the comparison of (<b>a</b>) NASA OC2 standard algorithm for Chl <span class="html-italic">a</span> and (<b>b</b>) Nechad algorithm for TSS in Issyk-Kul Lake.</p>
Full article ">Figure 11
<p>Spatial distribution of (<b>a</b>) Chl <span class="html-italic">a</span> and (<b>b</b>) TSS in Issyk-Kul Lake on 16 July 2022 obtained by LiDAR.</p>
Full article ">Figure 12
<p>Spatial distribution of Chl <span class="html-italic">a</span> in Issyk-Kul Lake on the start and end tracks on (<b>a</b>) 16 July 2022 (from 8:50 to 09:50), and (<b>b</b>) 17 July 2022 (from 17:00 to 18:00).</p>
Full article ">Figure 13
<p>Spatial distribution of TSS in Issyk-Kul Lake on the start and end tracks on (<b>a</b>) 16 July 2022 (from 08:50 to 09:50), and (<b>b</b>) 17 July 2022 (from 17:00 to 18:00).</p>
Full article ">Figure 13 Cont.
<p>Spatial distribution of TSS in Issyk-Kul Lake on the start and end tracks on (<b>a</b>) 16 July 2022 (from 08:50 to 09:50), and (<b>b</b>) 17 July 2022 (from 17:00 to 18:00).</p>
Full article ">Figure A1
<p>Remote-sensing reflectance on bands 490 nm (<b>a</b>) and 560 nm (<b>b</b>). Blue lines are measured reflectance, orange ones—satellite retrieved reflectance.</p>
Full article ">Figure A2
<p>Satellite retrieved remote-sensing reflectance (y-axes) against measured remote-sensing reflectance (x-axes) for different spectral bands.</p>
Full article ">
Back to TopTop