[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 (91)

Search Parameters:
Keywords = GOCI

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 8875 KiB  
Article
Exploring the Green Tide Transport Mechanisms and Evaluating Leeway Coefficient Estimation via Moderate-Resolution Geostationary Images
by Menghao Ji, Xin Dou, Chengyi Zhao and Jianting Zhu
Remote Sens. 2024, 16(16), 2934; https://doi.org/10.3390/rs16162934 - 10 Aug 2024
Viewed by 631
Abstract
The recurring occurrence of green tides as an ecological disaster has been reported annually in the Yellow Sea. While remote sensing technology effectively tracks the scale, extent, and duration of green tide outbreaks, there is limited research on the underlying driving mechanisms of [...] Read more.
The recurring occurrence of green tides as an ecological disaster has been reported annually in the Yellow Sea. While remote sensing technology effectively tracks the scale, extent, and duration of green tide outbreaks, there is limited research on the underlying driving mechanisms of green tide drift transport and the determination of the leeway coefficient. This study investigates the green tide transport mechanism and evaluates the feasibility of estimating the leeway coefficient by analyzing green tide drift velocities obtained from Geostationary Ocean Color Imager-II (GOCI-II) images using the maximum cross-correlation (MCC) technique and leeway method across various time intervals alongside ocean current and wind speed data. The results reveal the following: (1) Significant spatial variations in green tide movement, with a distinct boundary at 34°40′N. (2) Short-term green tide transport is primarily influenced by tidal forces, while wind and ocean currents, especially the combined Ekman and geostrophic current component, predominantly govern net transport. (3) Compared to 1, 3, and 7 h intervals, estimating the leeway coefficient with a 25 h interval is feasible for moderate-resolution geostationary images, yielding values consistent with previous studies. This study offers new insights into exploring the transport mechanisms of green tides through remote sensing-driven velocity. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Figure 1

Figure 1
<p>The map delineates the study area in the Yellow Sea where green tides reoccur. (<b>a</b>,<b>b</b>) The location of the study area and the M2 cotidal graph (the values of the contour lines represent the phase in degrees), respectively. (<b>c</b>,<b>d</b>) The average tidal current patterns from 10:30 to 13:30 local time during the periods of 4–10 June and 11–19 June 2021, respectively. The blue dashed line, positioned at 34°40′N, serves as the boundary for the drift velocity of the green tide.</p>
Full article ">Figure 2
<p>The distribution of green tides extracted at various time intervals: (<b>a</b>) 1 h interval, (<b>b</b>) 3 h interval, (<b>c</b>) 7 h interval, and (<b>d</b>) 25 h interval.</p>
Full article ">Figure 3
<p>Green tide drift velocities extracted at different time intervals is depicted in subplots (<b>a</b>–<b>f</b>). Subplots (<b>a</b>–<b>d</b>) illustrate the result for 1 h intervals, (<b>b</b>–<b>e</b>) for 3 h intervals, and (<b>c</b>–<b>f</b>) for 7 h intervals. The top panel represents the meridional direction, while the bottom panel represents the zonal direction. The red dashed line indicates 34°40′N, and the blue line is the trend line. North and east direction values are positive.</p>
Full article ">Figure 4
<p>The rose diagram of green tide drift velocities at 1, 3, and 7 h time intervals is depicted in subplots (<b>a</b>–<b>f</b>). The upper panel corresponds to the area south of 34°40′N, while the bottom panel corresponds to the region north of 34°40′N. Colors represent the speed, while the length of the bars indicates the frequency.</p>
Full article ">Figure 5
<p>Cumulative bar chart of green tide drift velocities under varying ocean currents and wind speeds at 1, 3, and 7 h intervals. Ulva represents the green tide drift velocity, while the names of ocean current components (including tides, Stokes, and Ekman plus geostrophic) represent their velocities in either the zonal or meridional directions. The wind multiplied by 2% represents the direct contribution of wind to the green tide. Blue triangles indicate the central latitude of the green tide patches. The left panel (<b>a</b>,<b>c</b>,<b>e</b>) and right panel (<b>b</b>,<b>d</b>,<b>f</b>) represent the meridional and zonal directions, respectively. North and east direction values are positive.</p>
Full article ">Figure 6
<p>Cumulative bar chart of green tide drift velocities under varying ocean currents and wind speeds at 25 h intervals. The left panel (<b>a</b>) and right panel (<b>b</b>) represent the meridional and zonal directions, respectively. The legend in the figure is consistent with that in <a href="#remotesensing-16-02934-f005" class="html-fig">Figure 5</a>. N represents the sample size.</p>
Full article ">Figure 7
<p>Stokes and leeway coefficients relative to wind calculated based on green tide drift velocities extracted at 25 h intervals. Mn represents mean value and SD represents standard deviation. The top panel (<b>a</b>) represents the meridional direction, while the bottom panel (<b>b</b>) represents the zonal direction.</p>
Full article ">Figure 8
<p>Box plots of the estimated leeway and Stokes coefficients at various time intervals. The yellow box represents the meridional direction, while the green box represents the zonal direction. The top panel shows the contribution of Stokes drift, while the bottom panel shows the contribution of wind.</p>
Full article ">Figure 9
<p>The rose diagram of tidal currents at 1, 3, and 7 h time intervals is depicted in subplots (<b>a</b>–<b>f</b>). The top panel corresponds to the area south of 34°40′N, while the bottom panel corresponds to the region north of 34°40′N. Colors represent the speed, while the length of the bars indicates the frequency.</p>
Full article ">Figure 10
<p>Wind rose of velocities for the green tide (<b>a</b>), Ekman plus geostrophic current (<b>b</b>), and wind (<b>c</b>) at 25 h intervals. Colors represent the speed, while the length of the bars indicates the frequency.</p>
Full article ">Figure 11
<p>Comparison of green tide drift velocity extracted from high-spatial-resolution images with GOCI-II images. (<b>a</b>,<b>b</b>) Comparison of green tide drift velocity extracted at 1 h intervals by GOCI-II and the Landsat 8 and Sentinel-2 image pair on 7 June 2021. (<b>c</b>,<b>d</b>) Comparison of green tide drift velocity extracted at 3 h intervals by GOCI-II and HY-1C/D on 6 June 2021. South and east direction values are positive.</p>
Full article ">Figure 12
<p>The distribution of green tides at 25 h intervals extracted from various satellite images. The left panel displays the GOCI-II image, while the right panel shows images from Landsat 8 and Sentinel-2.</p>
Full article ">
22 pages, 12746 KiB  
Article
Monitoring the Vertical Variations in Chlorophyll-a Concentration in Lake Chaohu Using the Geostationary Ocean Color Imager
by Hanhan Li, Xiaoqi Wei, Zehui Huang, Haoze Liu, Ronghua Ma, Menghua Wang, Minqi Hu, Lide Jiang and Kun Xue
Remote Sens. 2024, 16(14), 2611; https://doi.org/10.3390/rs16142611 - 17 Jul 2024
Viewed by 621
Abstract
Due to the external environment and the buoyancy of cyanobacteria, the inhomogeneous vertical distribution of phytoplankton in eutrophic lakes affects remote sensing reflectance (Rrs) and the inversion of surface chlorophyll-a concentration (Chla). In this study, vertical profiles [...] Read more.
Due to the external environment and the buoyancy of cyanobacteria, the inhomogeneous vertical distribution of phytoplankton in eutrophic lakes affects remote sensing reflectance (Rrs) and the inversion of surface chlorophyll-a concentration (Chla). In this study, vertical profiles of Chla(z) (where z is the water depth) and field Rrs (Rrs_F) were collected and utilized to retrieve the vertical profiles of Chla in Lake Chaohu in China. Chla(z) was categorized into vertically uniform (Type 1: N = 166) and vertically non-uniform (Type 2: N = 58) types. Based on the validation of the atmospheric correction performance of the Geostationary Ocean Color Imager (GOCI), a Chla(z) inversion model was developed for Lake Chaohu from 2011 to 2020 using GOCI Rrs data (Rrs_G). (1) Five functions of non-uniform Chla(z) were compared, and the best result was found for Chla(z) = a × exp(b × z) + c (R2 = 0.98, RMSE = 38.15 μg/L). (2) A decision tree of Chla(z) was established with the alternative floating algae index (AFAIRrs), the fluorescence line height (FLH), and wind speed (WIN), where the overall accuracy was 89% and the Kappa coefficient was 0.79. The Chla(z) inversion model for Type 1 was established using the empirical relationship between Chla (z = surface) and AFAIRrs (R2 = 0.58, RMSE = 10.17 μg/L). For Type 2, multivariate regression models were established to estimate the structural parameters of Chla(z) combined with Rrs_G and environmental parameters (R2 = 0.75, RMSE = 72.80 μg/L). (3) There are obvious spatial variations in Chla(z), especially from the water surface to a depth of 0.1 m; the largest diurnal variations were observed at 12:16 and 13:16 local time. The Chla(z) inversion method can determine Chla in different layers of each pixel, which is important for the scientific assessment of phytoplankton biomass and lake carbon and can provide vertical information for the short-term prediction of algal blooms (and the generation of corresponding warnings) in lake management. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Lake Chaohu in China. The spatial distributions of samplings are shown.</p>
Full article ">Figure 2
<p>Number of samplings conducted between May 2013 and November 2017.</p>
Full article ">Figure 3
<p>Number of GOCI images of Lake Chaohu in different months.</p>
Full article ">Figure 4
<p>Validation results for GOCI <span class="html-italic">R<sub>rs</sub>_G</span> versus in situ <span class="html-italic">R<sub>rs</sub>_F</span> at the relevant bands: (<b>a</b>) 660 nm, (<b>b</b>) 680 nm, (<b>c</b>) 745 nm, and (<b>d</b>) 865 nm.</p>
Full article ">Figure 5
<p>Measured Chl<span class="html-italic">a</span>(z) profiles in Lake Chaohu for (<b>a</b>) Type 1 model showing vertical uniformity and (<b>b</b>) Type 2 model showing vertical non-uniformity (Different color represent each of the different vertical profiles).</p>
Full article ">Figure 6
<p>Decision tree of identifying Chl<span class="html-italic">a</span> vertical profile classes based on AFAI<sub>Rrs</sub>, FLH, and wind speed, where AFAI<sub>Rrs</sub> and FLH are the indices derived from the in situ measurements and “WIN” represents wind speed.</p>
Full article ">Figure 7
<p>Sensitivity analysis of the threshold value of the decision tree for (<b>a</b>) classification accuracy (%) with changing WIN, (<b>b</b>) classification accuracy (%) with changing FLH, (<b>c</b>) classification accuracy (%) with changing AFAI<sub>Rrs1</sub> (WIN &gt; 3 m/s), and (<b>d</b>) classification accuracy (%) with changing AFAI<sub>Rrs2</sub> (WIN ≤ 3 m/s). The red markers (WIN = 3.0 m/s, FLH = −0.004, AFAI<sub>Rrs1</sub> = −0.002, and AFAI<sub>Rrs2</sub> = 0.002) are the thresholds with the highest classification accuracy. “OA” is overall accuracy.</p>
Full article ">Figure 8
<p>Training and validation of the Type 1 inversion model for: (<b>a</b>) the training results from the exponential equation with AFAI<sub>Rrs</sub> and (<b>b</b>) the validation results from the exponential model.</p>
Full article ">Figure 9
<p>Relationships between the <span class="html-italic">R<sub>rs</sub></span> at the single band, meteorological data, index factors, and the vertical structural parameters <span class="html-italic">a</span>, <span class="html-italic">b</span>, and <span class="html-italic">c</span>. Note that “Tem” is temperature and “WIN” is wind speed.</p>
Full article ">Figure 10
<p>Validation results for the Type 2 Chl<span class="html-italic">a</span>(z) model for (<b>a</b>) structure parameter <span class="html-italic">a</span>, (<b>b</b>) structure parameter <span class="html-italic">b</span>, and (<b>c</b>) structure parameter <span class="html-italic">c</span>, as well as (<b>d</b>) a scatter plot of measured Chl<span class="html-italic">a</span>(z) versus predicted Type 2 Chl<span class="html-italic">a</span>(z). Note that the log-scale was used in plots and statistics were calculated using the Chl<span class="html-italic">a</span> values.</p>
Full article ">Figure 11
<p>Yearly mean Chl<span class="html-italic">a</span>(z) from 2011 to 2020 in the three vertical layers (surface, 0.1 m, and 0.3 m) of Lake Chaohu.</p>
Full article ">Figure 12
<p>Monthly mean Chl<span class="html-italic">a</span> (z = surface, 0.1 m, and 0.3 m) from 2011 to 2020 in Lake Chaohu.</p>
Full article ">Figure 13
<p>Spatial distribution of monthly mean Chl<span class="html-italic">a</span> (z = surface, 0.1 m, and 0.3 m) from 2011 to 2020 in Lake Chaohu.</p>
Full article ">Figure 14
<p>Hourly mean Chl<span class="html-italic">a</span>(z) distribution results from 2011 to 2020 for (<b>a</b>–<b>c</b>) spatial distribution results at surface, 0.1 m, and 0.3 m; (<b>d</b>) the diurnal variations in three layers; and (<b>e</b>–<b>g</b>) the vertical structural parameters <span class="html-italic">a</span>, <span class="html-italic">b</span>, and <span class="html-italic">c</span>.</p>
Full article ">Figure 15
<p>Spatial distribution of the mean values of structural parameters for (<b>a</b>–<b>c</b>) the structural parameters <span class="html-italic">a</span>, <span class="html-italic">b</span>, and <span class="html-italic">c</span>, respectively.</p>
Full article ">Figure 16
<p>Vertical profiles of the mean Chl<span class="html-italic">a</span>(z) for (<b>a</b>) seasonal variations and (<b>b</b>) diurnal variations.</p>
Full article ">Figure 17
<p>Algorithm verification result based on the ABI-Bio algorithm (ABI-Bio) data.</p>
Full article ">Figure 18
<p>Vertical distribution of Chl<span class="html-italic">a</span>(z) in different layers at the surface and at 0.3 m, 0.6 m, and 1.0 m for the case of (<b>a</b>) vertical uniformity on 4 December 2014 and (<b>b</b>) vertical non-uniformity on 24 October 2014.</p>
Full article ">Figure 19
<p>Sensitivity analysis of the structural parameter model conducted by changing the parameters of WIN, GHI, B86, and AFAI<sub>Rrs</sub>/X ±20%, ±15%, ±10%, and ±5%, respectively, showing results of a specific variable for (<b>a</b>) WIN, (<b>b</b>) GHI, (<b>c</b>) B86, and (<b>d</b>) AFAI<sub>Rrs</sub>/X for parameters <span class="html-italic">a</span>, <span class="html-italic">b</span>, and <span class="html-italic">c</span>.</p>
Full article ">
20 pages, 12311 KiB  
Article
Evaluation of Rayleigh-Corrected Reflectance on Remote Detection of Algal Blooms in Optically Complex Coasts of East China Sea
by Chengxin Zhang, Bangyi Tao, Yunzhou Li, Libo Ai, Yixian Zhu, Liansong Liang, Haiqing Huang and Changpeng Li
Remote Sens. 2024, 16(13), 2304; https://doi.org/10.3390/rs16132304 - 24 Jun 2024
Viewed by 600
Abstract
This study used GOCI-II data to systematically evaluate the feasibility of Rayleigh-corrected reflectance (Rrc) to detect algal blooms in the complex optical environment of the East China Sea (ECS). Based on long-term in situ remote sensing reflectance (Rrs [...] Read more.
This study used GOCI-II data to systematically evaluate the feasibility of Rayleigh-corrected reflectance (Rrc) to detect algal blooms in the complex optical environment of the East China Sea (ECS). Based on long-term in situ remote sensing reflectance (Rrs), Rrc spectra demonstrated the similar capability of reflecting the water condition under various atmospheric conditions, and the baseline indices (BLIs) derived from Rrc and Rrs showed good consistency (R2 > 0.98). The effectiveness of five Rrc-based BLIs (SS490, CI, DI, FLH, and MCI) for algal bloom detection was assessed, among which SS490 and MCI showed better performances. A synthetic bloom detection algorithm based on the BLIs of Rrc was then developed to avoid the impact of turbid water. The validation of the BLI algorithm was carried out based on the in situ algal abundance data from 2021 to 2023. Specifically, SS490 showed the best bloom detection result (F-measure coefficient, FM = 0.97), followed by MCI (FM = 0.88). Since the 709 nm bands used in MCI were missing in many ocean color satellites, the SS490 algorithm was more useful in application. Compared to Rrs based bloom detection algorithms, synthetical Rrc BLI proposed in this paper provides more effective observation results and even better algal bloom detection performance. In conclusion, the study confirmed the feasibility of utilizing Rrc for algal bloom detection in the coastal areas of the ECS, and recognized the satisfactory performance of synthetical SS490 by comparing with the other BLIs. Full article
Show Figures

Figure 1

Figure 1
<p>Schematic of the study region, the majority of regions within the ECS are encompassed by optically complex waters. The circles represent data collected by the Wenzhou Marine Center between 2021 and 2023, with the center of the circles indicate the location of algal bloom (white) and non-bloom (green) sites, respectively.</p>
Full article ">Figure 2
<p>(<b>a</b>) Dongou oceanographic platform, (<b>b</b>) SeaPRISM radiometer, (<b>c</b>) Runjiang No. 1 experimental ship, (<b>d</b>) Cruise AOP observation system.</p>
Full article ">Figure 3
<p><span class="html-italic">R</span><sub>rc</sub> spectra of typical clean water, algal bloom water, medium turbid water, and turbid water, (<b>a</b>) original <span class="html-italic">R</span><sub>rc</sub> spectra and (<b>b</b>) normalized <span class="html-italic">R</span><sub>rc</sub> spectra.</p>
Full article ">Figure 4
<p>Scatter plots showing BLIs derived from (<b>a</b>–<b>c</b>) in situ measured or (<b>d</b>–<b>f</b>) GOCI-II <span class="html-italic">R</span><sub>rs</sub> with BLIs derived from GOCI-II <span class="html-italic">R</span><sub>rc</sub> on the Dongou platform in 2021.</p>
Full article ">Figure 5
<p>(<b>a</b>) Schematic diagram illustrated the underway observation section of the LORCE on 19 August 2021, and the yellow line represented the track. (<b>b</b>–<b>f</b>) BLIs comparison between the underway measured <span class="html-italic">R</span><sub>rs</sub> and GOCI-II-derived <span class="html-italic">R</span><sub>rc</sub>. Three GOCI-II images from the closest times to the measured <span class="html-italic">R</span><sub>rs</sub> acquisition times were selected for matching: 11:15, 12:15, and 13:15, respectively. In each scatter plot, the three parts separated by dashed lines correspond to matching results at different times, respectively.</p>
Full article ">Figure 6
<p>Comparison between long-term <span class="html-italic">R</span><sub>rc</sub> and <span class="html-italic">R</span><sub>rs</sub> calculated BLIs obtained by GOCI-II at Dongou platform location in 2021, (<b>a</b>–<b>e</b>) showing comparison results of SS490, CI, DI, FLH, and MCI, respectively.</p>
Full article ">Figure 7
<p>Distribution of BLIs (SS490, CI, DI, FLH, MCI) for four large algal blooms as well as one typical non-bloom day from 2021 to 2023. In situ stations are marked with yellow circles in the figure, and the corresponding algal cell abundance is labeled nearby. The first to fourth rows (labeled <b>a</b>–<b>x</b>) show the distribution of different baseline indices for two <span class="html-italic">P. donhaiense</span> blooms, one diatom bloom, and one <span class="html-italic">A. sanguinea</span> bloom, and the fifth row (labeled <b>y</b>–<b>D</b>) shows the distribution of different baseline indices for typical non-bloom in the winter, with the time of acquisition of the images labeled in the first subplot of each row. The columns of the figure show the RGB true-color composite image and the distribution images of SS490, CI, DI, FLH, and MCI, respectively, and the color bar of each baseline index is plotted at the bottom of each column. (Data from the boxed areas of the map will be used to plot the scatter distribution of the different baseline indices against the turbid water index, see <a href="#remotesensing-16-02304-f008" class="html-fig">Figure 8</a>).</p>
Full article ">Figure 8
<p>Scatterplots showing the distributions of different BLIs with TI, with each subplot corresponding to the selected regions boxed in <a href="#remotesensing-16-02304-f007" class="html-fig">Figure 7</a>, (<b>a</b>–<b>y</b>) correspond to <a href="#remotesensing-16-02304-f007" class="html-fig">Figure 7</a> (<b>b</b>–<b>l</b>,<b>n</b>–<b>r</b>,<b>t</b>–<b>x</b>,<b>z</b>–<b>D</b>), respectively. In each subplot, the scatter plots of SS490, CI, DI, FLH, and MCI with the TI were arranged from left to right. Blue, red, green, and brown circles were used to, respectively, denote clear water, algal bloom, medium turbid water, and turbid water units in the figure.</p>
Full article ">Figure 9
<p>(<b>a</b>) Schematic procedure of GOCI-II synthetical SS490 method for detecting algal blooms. Take a <span class="html-italic">P. donghaiense</span> bloom on 1 May 2023 as an example, (<b>b</b>) RGB true-color composite map, (<b>c</b>) original SS490 map, (<b>d</b>) SS490 distribution after cloud mask, (<b>e</b>) SS490 distribution after turbid correction, and (<b>f</b>) algal bloom area determined by synthetical SS490.</p>
Full article ">Figure 10
<p>Scatterplots showing measured algal cell abundance versus different <span class="html-italic">R</span><sub>rc</sub>-derived (<b>a</b>) SS490, (<b>b</b>) CI, (<b>c</b>) DI, (<b>d</b>) FLH and (<b>e</b>) MCI, respectively.</p>
Full article ">Figure 11
<p>Comparative validation plot of algal bloom detection accuracy based on different <span class="html-italic">R</span><sub>rc</sub>-derived (<b>a</b>) LHR, (<b>b</b>) RAB, (<b>c</b>) RI and (<b>d</b>) SS510, respectively.</p>
Full article ">Figure 12
<p>(<b>a</b>–<b>c</b>) The RGB image and (<b>d</b>–<b>f</b>) <span class="html-italic">R</span><sub>rc</sub>, <span class="html-italic">R</span><sub>rs</sub> spectra corresponding to the green points with detection errors in <a href="#remotesensing-16-02304-f011" class="html-fig">Figure 11</a>.</p>
Full article ">
17 pages, 6389 KiB  
Article
Continuity and Enhancements in Sea Surface Salinity Estimation in the East China Sea Using GOCI and GOCI-II: Challenges and Further Developments
by Eunna Jang, Jong-Kuk Choi and Jae-Hyun Ahn
Remote Sens. 2024, 16(12), 2111; https://doi.org/10.3390/rs16122111 - 11 Jun 2024
Cited by 1 | Viewed by 685
Abstract
During the summer, substantial freshwater discharge from the Changjiang River into the East China Sea (ECS) results in extensive low-salinity water (LSW) plumes that significantly affect regions along the southern Korean Peninsula and near Jeju Island. Previous research developed an empirical equation to [...] Read more.
During the summer, substantial freshwater discharge from the Changjiang River into the East China Sea (ECS) results in extensive low-salinity water (LSW) plumes that significantly affect regions along the southern Korean Peninsula and near Jeju Island. Previous research developed an empirical equation to estimate sea surface salinity (SSS) in the ECS during the summer season using remote-sensing reflectance (Rrs) data from bands 3–6 (490, 555, 660, and 680 nm) of the Geostationary Ocean Color Imager (GOCI). With the conclusion of the GOCI mission in March 2021, this study aims to ensure the continuity of SSS estimation in the ECS by transitioning to its successor, the GOCI-II. This transition was facilitated through two approaches: applying the existing GOCI-based equation and introducing a new machine learning method using a random forest model. Our analysis demonstrated a high correlation between SSS estimates derived from the GOCI and GOCI-II when applying the equation developed for the GOCI to both satellites, as indicated by a robust R2 value of 0.984 and a low RMSD of 0.8465 psu. This study successfully addressed the challenge of maintaining continuous SSS estimation in the ECS post-GOCI mission and evaluated the accuracy and limitations of the GOCI-II-derived SSS, proposing future strategies to enhance its effectiveness. Full article
(This article belongs to the Section Ocean Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The East China Sea (ECS) study area and the in situ data collection sites used in this study. The green circle represents the Ieodo Ocean Research Station’s location (32.12°N and 125.18°S), and the blue circles depict the distribution of serial oceanographic observation data collected in August 2020 and 2021.</p>
Full article ">Figure 2
<p>Flow chart for the procedural steps for estimating the SSS from GOCI-II data. Choi’s equation is represented as a multilinear regression equation used to estimate the SSS in the ECS using GOCI data, as explained in <a href="#sec3dot2-remotesensing-16-02111" class="html-sec">Section 3.2</a>. ‘GII-C SSS’ is the SSS derived from GOCI-II using Choi’s equation, whereas ‘GII-RF SSS’ indicates the SSS derived from GOCI-II using the machine learning model.</p>
Full article ">Figure 3
<p>Scatterplots of the adjusted GOCI-II R<sub>rs</sub> versus GOCI R<sub>rs</sub> for bands at (<b>a</b>) 443 nm, (<b>b</b>) 555 nm, (<b>c</b>) 660 nm, and (<b>d</b>) 680 nm. The color bar depicts data density on a logarithmic scale, visually depicting the alignment of the GOCI-II R<sub>rs</sub> with the GOCI R<sub>rs</sub> across the examined spectral bands.</p>
Full article ">Figure 4
<p>Scatterplots comparing the GOCI-derived SSS with GII-C SSS, distinguishing between SSS estimates derived from adjusted GOCI-II R<sub>rs</sub> data (<b>a</b>,<b>c</b>) and those using original GOCI-II R<sub>rs</sub> data (<b>b</b>,<b>d</b>). The results for all August 2020 pixels are shown in (<b>a</b>,<b>b</b>), while (<b>c</b>,<b>d</b>) exclude GOCI-II pixels that were flagged as turbid pixels. The color bar depicts data density on a logarithmic scale.</p>
Full article ">Figure 5
<p>Modeling results from the random forest (RF) model, using in situ data for calibration (<b>a</b>) and validation (<b>b</b>), demonstrating the model’s performance in estimating SSS.</p>
Full article ">Figure 6
<p>Scatterplots comparing in situ data with the GII-C SSS (<b>a</b>), GII-RF SSS (<b>b</b>), GOCI-derived SSS (<b>c</b>), SMAP L2B SSS (<b>d</b>), and SMAP L3 SSS (<b>e</b>), distinguishing between I-ORS data (green circles) and NIFS data (blue circles).</p>
Full article ">Figure 7
<p>The daily SSS comparison between the in situ I-ORS data and various satellite-derived SSS products for August 2020 and 2021, with the X-axis marking the days of August.</p>
Full article ">Figure 8
<p>Scatterplots comparing the GOCI-II-derived SSS with the SMAP L2B SSS (<b>a</b>–<b>c</b>) and SMAP L3 SSS (<b>d</b>–<b>f</b>), covering all pixels from August 2020 to 2021. (<b>a</b>,<b>b</b>,<b>d</b>,<b>e</b>) include the entire dataset, whereas (<b>c</b>,<b>f</b>) focus specifically on pixels from GOCI-II data not classified as turbid. The color bar depicts data density on a logarithmic scale.</p>
Full article ">Figure 9
<p>Spatial distribution maps of SSS as derived from GII-C SSS (<b>a</b>), GOCI-derived SSS (<b>b</b>), SMAP L2B SSS (<b>c</b>), and SMAP L3 SSS (<b>d</b>) data at UTC03 on 30 August 2020, with the I-ORS highlighted by a black circle.</p>
Full article ">Figure 10
<p>Spatial distribution of GOCI-II R<sub>rs</sub> (<b>a</b>–<b>d</b>) and the resultant GII-C SSS (<b>e</b>) for UTC03 on 30 August 2020, illustrating the close alignment between R<sub>rs</sub> and SSS estimations.</p>
Full article ">Figure 11
<p>Spatial distribution of the GOCI-II R<sub>rs</sub> at 555 nm (<b>a</b>), 660 nm (<b>b</b>), and GOCI-II-derived SSS (<b>c</b>) at UTC03 on 15 August 2020. (<b>b</b>) Delineates each GOCI-II slot with matching boundaries and slot numbers, highlighting the spatial variability across the slots.</p>
Full article ">
32 pages, 7440 KiB  
Review
A Systematic Review of the Application of the Geostationary Ocean Color Imager to the Water Quality Monitoring of Inland and Coastal Waters
by Shidi Shao, Yu Wang, Ge Liu and Kaishan Song
Remote Sens. 2024, 16(9), 1623; https://doi.org/10.3390/rs16091623 - 1 May 2024
Viewed by 1642
Abstract
In recent decades, eutrophication in inland and coastal waters (ICWs) has increased due to anthropogenic activities and global warming, thus requiring timely monitoring. Compared with traditional sampling and laboratory analysis methods, satellite remote sensing technology can provide macro-scale, low-cost, and near real-time water [...] Read more.
In recent decades, eutrophication in inland and coastal waters (ICWs) has increased due to anthropogenic activities and global warming, thus requiring timely monitoring. Compared with traditional sampling and laboratory analysis methods, satellite remote sensing technology can provide macro-scale, low-cost, and near real-time water quality monitoring services. The Geostationary Ocean Color Imager (GOCI), aboard the Communication Ocean and Meteorological Satellite (COMS) from the Republic of Korea, marked a significant milestone as the world’s inaugural geostationary ocean color observation satellite. Its operational tenure spanned from 1 April 2011 to 31 March 2021. Over ten years, the GOCI has observed oceans, coastal waters, and inland waters within its 2500 km × 2500 km target area centered on the Korean Peninsula. The most attractive feature of the GOCI, compared with other commonly used water color sensors, was its high temporal resolution (1 h, eight times daily from 0 UTC to 7 UTC), providing an opportunity to monitor ICWs, where their water quality can undergo significant changes within a day. This study aims to comprehensively review GOCI features and applications in ICWs, analyzing progress in atmospheric correction algorithms and water quality monitoring. Analyzing 123 articles from the Web of Science and China National Knowledge Infrastructure (CNKI) through a bibliometric quantitative approach, we examined the GOCI’s strength and performance with different processing methods. These articles reveal that the GOCI played an essential role in monitoring the ecological health of ICWs in its observation coverage (2500 km × 2500 km) in East Asia. The GOCI has led the way to a new era of geostationary ocean satellites, providing new technical means for monitoring water quality in oceans, coastal zones, and inland lakes. We also discuss the challenges encountered by Geostationary Ocean Color Sensors in monitoring water quality and provide suggestions for future Geostationary Ocean Color Sensors to better monitor the ICWs. Full article
Show Figures

Figure 1

Figure 1
<p>Structure of this paper.</p>
Full article ">Figure 2
<p>The regional observation area of GOCI (<a href="https://kosc.kiost.ac.kr/index.nm?menuCd=43&amp;lang=en" target="_blank">https://kosc.kiost.ac.kr/index.nm?menuCd=43&amp;lang=en</a>, accessed on 25 January 2024).</p>
Full article ">Figure 3
<p>Number and countries of papers published related to GOCI between 2010 and 2023.</p>
Full article ">Figure 4
<p>Map of the study area.</p>
Full article ">Figure 5
<p>Journals and number of publications.</p>
Full article ">Figure 6
<p>Keyword mapping from the Web of Science, searched by author name.</p>
Full article ">Figure 7
<p>Overview of the application of GOCI retrieved by subject terms from CNKI.</p>
Full article ">Figure 8
<p>Summary and proportion of the applications of GOCI in inland and coastal waters.</p>
Full article ">Figure 9
<p>Schematic diagram of integrated ground–air space.</p>
Full article ">Figure 10
<p>Schematic diagram of image fusion, taking GOCI-II and Himawari8/9 as an example.</p>
Full article ">
17 pages, 32322 KiB  
Article
Automatic Detection of Floating Ulva prolifera Bloom from Optical Satellite Imagery
by Hailong Zhang, Quan Qin, Deyong Sun, Xiaomin Ye, Shengqiang Wang and Zhixin Zong
J. Mar. Sci. Eng. 2024, 12(4), 680; https://doi.org/10.3390/jmse12040680 - 19 Apr 2024
Cited by 1 | Viewed by 1133
Abstract
Annual outbreaks of floating Ulva prolifera blooms in the Yellow Sea have caused serious local environmental and economic problems. Rapid and effective monitoring of Ulva blooms from satellite observations with wide spatial-temporal coverage can greatly enhance disaster response efforts. Various satellite sensors and [...] Read more.
Annual outbreaks of floating Ulva prolifera blooms in the Yellow Sea have caused serious local environmental and economic problems. Rapid and effective monitoring of Ulva blooms from satellite observations with wide spatial-temporal coverage can greatly enhance disaster response efforts. Various satellite sensors and remote sensing methods have been employed for Ulva detection, yet automatic and rapid Ulva detection remains challenging mainly due to complex observation scenarios present in different satellite images, and even within a single satellite image. Here, a reliable and fully automatic method was proposed for the rapid extraction of Ulva features using the Tasseled-Cap Greenness (TCG) index from satellite top-of-atmosphere reflectance (RTOA) data. Based on the TCG characteristics of Ulva and Ulva-free targets, a local adaptive threshold (LAT) approach was utilized to automatically select a TCG threshold for moving pixel windows. When tested on HY1C/D-Coastal Zone Imager (CZI) images, the proposed method, termed the TCG-LAT method, achieved over 95% Ulva detection accuracy though cross-comparison with the TCG and VBFAH indexes with a visually determined threshold. It exhibited robust performance even against complex water backgrounds and under non-optimal observing conditions with sun glint and cloud cover. The TCG-LAT method was further applied to multiple HY1C/D-CZI images for automatic Ulva bloom monitoring in the Yellow Sea in 2023. Moreover, promising results were obtained by applying the TCG-LAT method to multiple optical satellite sensors, including GF-Wide Field View Camera (GF-WFV), HJ-Charge Coupled Device (HJ-CCD), Sentinel2B-Multispectral Imager (S2B-MSI), and the Geostationary Ocean Color Imager (GOCI-II). The TCG-LAT method is poised for integration into operational systems for disaster monitoring to enable the rapid monitoring of Ulva blooms in nearshore waters, facilitated by the availability of near-real-time satellite images. Full article
(This article belongs to the Special Issue New Advances in Marine Remote Sensing Applications)
Show Figures

Figure 1

Figure 1
<p>Workflow for the automatic <span class="html-italic">Ulva</span> detection from optical satellite images.</p>
Full article ">Figure 2
<p>HY1C-CZI false-color RGB image on 13 June 2023 (<b>a</b>), and the distribution of <span class="html-italic">R</span><sub>TOA,red</sub> (<b>b</b>). The bright targets (i.e., other class) were extracted using the BT_red approach (<b>c</b>). The diagram for the automatic Th<sub>red</sub> selection (<b>d</b>).</p>
Full article ">Figure 3
<p>The diagrams for the frequency distribution histogram of TCG: without <span class="html-italic">Ulva</span> (<b>a</b>) and with <span class="html-italic">Ulva</span> pixels (<b>b</b>); the corresponding example of satellite images (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 4
<p>The distribution of the <span class="html-italic">Ulva</span> and <span class="html-italic">Ulva</span>-free pixel samples in the CIE color space (<b>a</b>). Histograms of CIE-<span class="html-italic">α</span> (<b>b</b>) and CIE-<span class="html-italic">x</span> (<b>c</b>) for the <span class="html-italic">Ulva</span> and water pixels.</p>
Full article ">Figure 5
<p>Satellite images of HY1C/D-CZI during 2020 to 2022 (<b>top panel</b>). The FRGB (NIR–red–green) images of the selected sub-images (<b>middle panel</b>), where their locations marked by white boxes were shown in top panel, and their TCG distributions (<b>bottom panel</b>).</p>
Full article ">Figure 6
<p>Satellite HY1C-CZI image collected on 23 June 2023 showing <span class="html-italic">Ulva</span> bloom in the Yellow Sea (<b>a</b>). The TCG images for the selected sub-regions (<b>b</b>), where their locations marked by white boxes are shown in (<b>a</b>), and the pixel-wise distribution of TCG for four regions (<b>c</b>) with individual regional thresholds marked by blue dashed lines.</p>
Full article ">Figure 7
<p>The HY1C-CZI FRGB image on 6 June 2021 and the locations of the selected sub-regions marked by white boxes (<b>a</b>). For four sub-regions, the FRGB images (<b>b</b>), TCG images (<b>c</b>), <span class="html-italic">Ulva</span> detection results (<b>d</b>), and the TH thresholds determined by the TCG-LAT method (<b>e</b>).</p>
Full article ">Figure 8
<p>The HY1C-CZI FRGB image on 6 June 2021 (<b>a</b>) and the detected <span class="html-italic">Ulva</span> result using the TCG-LAT method (<b>b</b>).</p>
Full article ">Figure 9
<p>Cross-comparison between satellite <span class="html-italic">Ulva</span> detection using different three methods for three selected sub-regions as depicted in <a href="#jmse-12-00680-f007" class="html-fig">Figure 7</a>. HY1C-CZI FRGB images (<b>a</b>) and the <span class="html-italic">Ulva</span> results detected by the TCG-VDT method (<b>b</b>) and TCG-LAT method (<b>c</b>). HJ2B-CCD FRGB images (<b>e</b>) and the <span class="html-italic">Ulva</span> results detected by the VBFAH-VDT method (<b>d</b>).</p>
Full article ">Figure 10
<p>The performance analysis of the TCG-LAT method under different observing conditions based on HY1C-CZI image (<b>a</b>) and HY1D-CZI image (<b>b</b>): clear water (R1), turbid water (R2), sun glint (R3), and cloud cover (R4).</p>
Full article ">Figure 11
<p><span class="html-italic">Ulva</span> distribution and coverage area for the 2023 <span class="html-italic">Ulva</span> bloom event detected from HY1C/D-CZI images.</p>
Full article ">Figure 12
<p>The <span class="html-italic">Ulva</span> results detected by the TCG-LAT method and their runtimes with different window sizes.</p>
Full article ">Figure 13
<p>Satellite FRGB images and the <span class="html-italic">Ulva</span> detection results using the TCG-LAT method from HJ-CCD image (<b>a-1</b>,<b>a-2</b>), GF1-WFV image (<b>b-1</b>,<b>b-2</b>), Sentinel2B-MSI (<b>c-1</b>,<b>c-2</b>), and GOCI (<b>d-1</b>,<b>d-2</b>).</p>
Full article ">
19 pages, 13159 KiB  
Article
An Improved Data Interpolating Empirical Orthogonal Function Method for Data Reconstruction: A Case Study of the Chlorophyll-a Concentration in the Bohai Sea, China
by Tongfang Hong, Rufu Qin and Zhounan Xu
Appl. Sci. 2024, 14(7), 2803; https://doi.org/10.3390/app14072803 - 27 Mar 2024
Viewed by 834
Abstract
Chlorophyll-a (chl-a) serves as a key indicator in water quality and harmful algal blooms (HABs) research. While satellite ocean color data have greatly advanced chl-a research and HABs monitoring, missing data caused by cloud cover and other factors limit the spatiotemporal continuity and [...] Read more.
Chlorophyll-a (chl-a) serves as a key indicator in water quality and harmful algal blooms (HABs) research. While satellite ocean color data have greatly advanced chl-a research and HABs monitoring, missing data caused by cloud cover and other factors limit the spatiotemporal continuity and the utility of remote sensing data products. The Data Interpolating Empirical Orthogonal Function (DINEOF) method, widely used to reconstruct missing values in remote sensing datasets, is open to improvement in terms of computational accuracy and efficiency. We propose an improved method called Concentration-Stratified DINEOF (CS-DINEOF), which uses a coordinate–value correlative data division strategy to stratify the study area into several subregions based on annual average chl-a concentration. The proposed method clusters data points with similar spatiotemporal patterns, allowing for more targeted and effective reconstruction in each sub-dataset. The feasibility and advantage of the proposed method are tested and evaluated in the experiments of chl-a data reconstruction in the water of the Bohai Sea. Compared with the ordinary DINEOF method, the CS-DINEOF method improves the reconstruction accuracy, with an average Root Mean Square Error (RMSE) reduction of 0.0281 mg/m3, and saves computational time by 228.9%. Furthermore, the gap-free images generated from CS-DINEOF are able to illustrate small variations and details of the chl-a distribution in local areas. We can conclude that the proposed CS-DINEOF method is superior in providing significant insights for water quality and HABs studies in the Bohai Sea region. Full article
(This article belongs to the Section Marine Science and Engineering)
Show Figures

Figure 1

Figure 1
<p>A satellite remote sensing image of the study area in the Bohai Sea region.</p>
Full article ">Figure 2
<p>Spatial distribution of the pixels TMR in 2019 chl-a original dataset of the Bohai Sea region.</p>
Full article ">Figure 3
<p>Spatial distribution of average chl-a concentration in the Bohai Sea in 2019. (<b>a</b>) A stretch color map of chl-a concentration with a blue to red color ramp, in which blue indicates low concentration and red represents high concentration; (<b>b</b>) a graduated map divided into 10 levels with different intervals.</p>
Full article ">Figure 4
<p>Overall workflow of CS-DINEOF algorithm.</p>
Full article ">Figure 5
<p>Statistics of data missing from 2019 chl-a refined dataset of Bohai Sea region. (<b>a</b>) SMR of each chl-a image in the refined dataset. (<b>b</b>) Spatial distribution of TMR for each pixel point in the refined dataset. (<b>c</b>) Comparison of OMR for each sub-dataset before and after missing-data checks.</p>
Full article ">Figure 6
<p>Four original chl-a images on (<b>a</b>) 15 February, (<b>b</b>) 26 April, (<b>c</b>) 22 September, and (<b>d</b>) 28 November in 2019, with SMRs of 78.47%, 31.66%, 18.98%, and 53.46%, respectively.</p>
Full article ">Figure 7
<p>The reconstruction results using (<b>a</b>) ordinary DINEOF and (<b>b</b>) CS-DINEOF on 15 February 2019. Specific local area is marked with red box in (<b>a</b>,<b>b</b>), and the corresponding zoomed-in views are presented in panels (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 8
<p>The reconstruction results using (<b>a</b>) ordinary DINEOF and (<b>b</b>) CS-DINEOF on 26 April 2019. Specific local area is marked with red box in (<b>a</b>,<b>b</b>), and the corresponding zoomed-in views are shown in panels (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 9
<p>The reconstruction results using (<b>a</b>) ordinary DINEOF and (<b>b</b>) CS-DINEOF on 22 September 2019. Specific local area is marked with red box in (<b>a</b>,<b>b</b>), and the corresponding zoomed-in views are shown in panels (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 10
<p>The reconstruction results using (<b>a</b>) ordinary DINEOF and (<b>b</b>) CS-DINEOF on 28 November 2019. Specific local area is marked with red box in (<b>a</b>,<b>b</b>), and the corresponding zoomed-in views are shown in panels (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 11
<p>Density scatterplots of original and reconstructed chl-a values by the CS-DINEOF method at cross-validation points on (<b>a</b>) 15 February, (<b>b</b>) 26 April, (<b>c</b>) 22 September, and (<b>d</b>) 28 November in 2019.</p>
Full article ">Figure 12
<p>RMSEs obtained from CS-DINEOF and ordinary DINEOF methods of different data SMRs of all images in the chl-a dataset.</p>
Full article ">Figure 13
<p>RMSEs obtained from CS-DINEOF and DINEOF methods of different sub-datasets.</p>
Full article ">
21 pages, 15824 KiB  
Article
Remote Sensing Observations of a Coastal Water Environment Based on Neural Network and Spatiotemporal Fusion Technology: A Case Study of Hangzhou Bay
by Rugang Tang, Xiaodao Wei, Chao Chen, Rong Jiang and Fang Shen
Remote Sens. 2024, 16(5), 800; https://doi.org/10.3390/rs16050800 - 25 Feb 2024
Viewed by 1149
Abstract
The coastal environment is characterized by high, multi-scale dynamics and the corresponding observations from a single remote sensing sensor are still facing challenges in achieving both high temporal and spatial resolution. This study proposed a spatiotemporal fusion model for coastal environments, which could [...] Read more.
The coastal environment is characterized by high, multi-scale dynamics and the corresponding observations from a single remote sensing sensor are still facing challenges in achieving both high temporal and spatial resolution. This study proposed a spatiotemporal fusion model for coastal environments, which could fully enhance the efficiency of remote sensing data use and overcome the shortcomings of traditional spatiotemporal models that are insensitive to small-scale disturbances. The Enhanced Deep Super-Resolution Network (EDSR) was used to reconstruct spatial features in the lower spatial resolution GOCI-II data. The spatial features obtained instead of GOCI-II data were fed into the spatiotemporal fusion model, which enabled the fusion data to achieve an hour-by-hour observation of the water color and morphology information changes at 30 m resolution, including the changes in the spatial and temporal distributions of suspended particulate matter (SPM), the characterization of the vortex street caused by the bridge piers, the inundation process of the tidal flats, and coastline changes. In addition, this study analyzed the various factors affecting fusion accuracy, including spectral difference, errors in both temporal difference and location distance, and the structure of the EDSR model on the fusion accuracy. It is demonstrated that the location distance error and the spectral difference have the most significant impact on the fusion data, which may lead to the introduction of some ambiguous or erroneous spatial features. Full article
(This article belongs to the Section Remote Sensing in Geology, Geomorphology and Hydrology)
Show Figures

Figure 1

Figure 1
<p>The study area and the SPM distribution based on Landsat-9 data on 20 December 2022. A.1, A.2, and A.3 are the regions of interest for water color and morphological dynamics observations. The star (★) signifies the locations of the Zhapu (S.1) and Daishan (S.2) hydrological stations, which supplied measured water levels for the regions of interest.</p>
Full article ">Figure 2
<p>Structure of the EDSR model.</p>
Full article ">Figure 3
<p>The production process of high spatiotemporal fusion data.</p>
Full article ">Figure 4
<p>Cross-comparison of <span class="html-italic">R</span><sub>rs</sub> in the blue, green, red, and near-infrared bands of GOCI-II and Landsat 8/9. From top to bottom, each row corresponds to the comparison results on 27 February 2022 (<b>a</b>–<b>d</b>), 15 March 2022 (<b>e</b>–<b>h</b>), 20 December 2022 (<b>i</b>–<b>l</b>), and 29 January 2023 (<b>m</b>–<b>p</b>).</p>
Full article ">Figure 5
<p>The feature reconstruction capability of multiple upscaling models on observed data. (<b>a</b>) Observations of the model with a spatial resolution of 240 m; (<b>b</b>–<b>i</b>) high spatial resolution data, predicted based on the EDSR model, CSI, and their combination models, with a target spatial resolution of 30 m. The model is named in the format of [upscaling method_origin spatial resolution_target spatial resolution]. For example, [EDSR_240_120, CSI_120_30] refers to the model trained by EDSR, which reconstructs 240 m spatial resolution data into 120 m spatial resolution data, and then upscales it to 30 m using CSI; (<b>j</b>) the true values of the data, with a spatial resolution of 30 m, were used to evaluate the feature reconstruction effects of the models.</p>
Full article ">Figure 6
<p>Water levels at Daishan Hydrological Station obtained from the Global Tide Prediction Service Platform (<a href="http://global-tide.nmdis.org.cn/" target="_blank">http://global-tide.nmdis.org.cn/</a>, accessed on 1 October 2023), hollow circles are hourly predicted water levels, and the gray columns correspond to the observation periods of GOCI-II and fusion data in <a href="#remotesensing-16-00800-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 7
<p>True-color observations of GOCI-II, Landsat-8, and fusion data on the north shore tidal flats of Daishan County on 29 January 2023, from 9:30 to 15:30. (<b>a</b>–<b>g</b>) Disturbances in the inundation area of the east–west tidal flats observed by GOCI-II; (<b>h</b>,<b>j</b>–<b>n</b>) predicted results of spatiotemporal fusion at the corresponding moments of GOCI-II; (<b>i</b>) observation data of Landsat-8; (<b>h.1</b>–<b>n.1</b>) spatiotemporal variations of the western tidal flats, corresponding to the yellow box in (<b>h</b>–<b>n</b>); (<b>h.2</b>–<b>n.2</b>) spatiotemporal variations of the eastern tidal flats, corresponding to the green box in (<b>h</b>–<b>n</b>).</p>
Full article ">Figure 8
<p>Spatial distribution of SPM in the central part of Hangzhou Bay on 15 March 2022. (<b>a</b>) Landsat-8 true-color composite image with two sections and a region of interest detailed in <a href="#remotesensing-16-00800-f009" class="html-fig">Figure 9</a> and <a href="#remotesensing-16-00800-f010" class="html-fig">Figure 10</a>; (<b>b</b>–<b>h</b>) variation of the spatial distribution of SPM in central Hangzhou Bay derived from fusion data.</p>
Full article ">Figure 9
<p>Spatial distribution of SPM concentrations near Hangzhou Bay Bridge from GOCI-II and fusion data inversion. (<b>a</b>–<b>g</b>) SPM concentrations derived from GOCI-II data near Hangzhou Bay Bridge on 15 March 2022, 9:30–15:30; (<b>h</b>,<b>j</b>–<b>n</b>) SPM concentration derived from spatiotemporal fusion data near Hangzhou Bay Bridge, the black dotted box in (<b>h</b>) represents the SPM mutation area near the bridge pier due to vortex streets; (<b>i</b>) SPM concentrations derived from Landsat-8 data near Hangzhou Bay Bridge on 15 March 2022, 10:30.</p>
Full article ">Figure 10
<p>SPM distribution on March 15 2022 at the north and south cross-sections of the Hangzhou Bay Bridge. (<b>a</b>) SPM distribution at Section 1 in the northern part of the Hangzhou Bay Bridge from 9:30 to 15:30. A data gap at 10 km was caused by the bridge cover-up; (<b>b</b>) SPM distribution at Section 2 in the southern part of the Hangzhou Bay Bridge from 9:30 to 15:30; (<b>c</b>) water level at Zhapu station (S.1 in <a href="#remotesensing-16-00800-f001" class="html-fig">Figure 1</a>), hollow circles are predicted tide heights and gray columns are the observation periods of the GOCI-II.</p>
Full article ">Figure 11
<p>Spatiotemporal variation in the coastline on the southern coast of Hangzhou Bay observed by GOCI-II and fusion data. (<b>a</b>,<b>b</b>) GOCI-II and fusion data observation at 9:30 on 27 February 2022; (<b>c</b>,<b>d</b>) GOCI-II and fusion data observation at 10:30 on 20 December 2022. (<b>e</b>) Coastal deposit (orange) and erosion (green) areas identified by fusion data.</p>
Full article ">Figure 12
<p>Cross-comparison between GOCI-II and Landsat-8 on 29 January 2023, in the A.2 area (the tidal flats in <a href="#remotesensing-16-00800-f001" class="html-fig">Figure 1</a>). Warmer colors represent areas with higher point density, while cooler colors indicate lower density regions. (<b>a</b>–<b>d</b>) scatter plot of the <span class="html-italic">L</span><sub>TOA</sub>; (<b>e</b>–<b>h</b>) scatter plot of the <span class="html-italic">R</span><sub>rs</sub>.</p>
Full article ">Figure 13
<p>Model performance evaluation and error analysis based on simulated data. (<b>a</b>) Low spatial resolution image at time t<sub>1</sub>; (<b>b</b>) high spatial resolution image at time t<sub>1</sub>; (<b>c</b>) low spatial resolution image at time t<sub>2</sub>; (<b>d</b>) high spatial resolution image at time t<sub>2</sub>, serving as the true value for the fusion data; (<b>e</b>) fusion data predicted based on the STARFM; (<b>f</b>) fusion data predicted based on the EDSR model and the STARFM under theoretical conditions; (<b>g</b>) fusion data predicted based on the EDSR model and the STARFM considering location distance error (250m); (<b>h</b>) fusion data predicted based on the EDSR model and the STARFM considering temporal difference error (0.14% random noise); (<b>i</b>) fusion data predicted based on the EDSR model and the STARFM considering a spectral difference with a 20% offset added to the low spatial resolution image.</p>
Full article ">Figure 14
<p>True-color fusion data of the tidal flats along the northern side of Daishan County at 14:30 on 29 January 2023. (<b>a</b>) Spatiotemporal fusion data generated by <span class="html-italic">L</span><sub>TOA</sub> of GOCI-II and Landsat-8; (<b>b</b>,<b>c</b>) western and eastern tidal flats in the fused data, incorrect spatial features are marked by red boxes; (<b>d</b>,<b>e</b>) corresponding fusion data of (<b>b</b>,<b>c</b>) generated using the <span class="html-italic">R</span><sub>rs</sub>.</p>
Full article ">
19 pages, 14918 KiB  
Article
Ocean Colour Atmospheric Correction for Optically Complex Waters under High Solar Zenith Angles: Facilitating Frequent Diurnal Monitoring and Management
by Yongquan Wang, Huizeng Liu, Zhengxin Zhang, Yanru Wang, Demei Zhao, Yu Zhang, Qingquan Li and Guofeng Wu
Remote Sens. 2024, 16(1), 183; https://doi.org/10.3390/rs16010183 - 31 Dec 2023
Cited by 2 | Viewed by 1375
Abstract
Accurate atmospheric correction (AC) is one fundamental and essential step for successful ocean colour remote-sensing applications. Currently, most ACs and the associated ocean colour remote-sensing applications are restricted to solar zenith angles (SZAs) lower than 70°. The ACs under high SZAs present degraded [...] Read more.
Accurate atmospheric correction (AC) is one fundamental and essential step for successful ocean colour remote-sensing applications. Currently, most ACs and the associated ocean colour remote-sensing applications are restricted to solar zenith angles (SZAs) lower than 70°. The ACs under high SZAs present degraded accuracy or even failure problems, rendering the satellite retrievals of water quality parameters more challenging. Additionally, the complexity of the bio-optical properties of the coastal waters and the presence of complex aerosols add to the difficulty of AC. To address this challenge, this study proposed an AC algorithm based on extreme gradient boosting (XGBoost) for optically complex waters under high SZAs. The algorithm presented in this research has been developed using pairs of Geostationary Ocean Colour Imager (GOCI) high-quality noontime remote-sensing reflectance (Rrs) and the Rayleigh-corrected reflectance (ρrc) derived from the Ocean Colour–Simultaneous Marine and Aerosol Retrieval Tool (OC-SMART) in the morning (08:55 LT) and at dusk (15:55 LT). The algorithm was further examined using the daily GOCI images acquired in the morning and at dusk, and the hourly (total suspended sediment) TSS concentration was also obtained based on the atmospherically corrected GOCI data. The results showed that: (i) the model produced an accurate fitting performance (R2 ≥ 0.90, RMSD ≤ 0.0034 sr−1); (ii) the model had a high validation accuracy with an independent dataset (R2 = 0.92–0.97, MAPD = 8.2–26.81% and quality assurance (QA) score = 0.9–1); and (iii) the model successfully retrieved more valid Rrs for GOCI images under high SZAs and enhanced the accuracy and coverage of TSS mapping. This algorithm has great potential to be applied to AC for optically complex waters under high SZAs, thus increasing the frequency of available observations in a day. Full article
(This article belongs to the Special Issue GIS and Remote Sensing in Ocean and Coastal Ecology)
Show Figures

Figure 1

Figure 1
<p>GOCI hourly RGB image (<b>a</b>), <span class="html-italic">ρ<sub>rc</sub></span> (555 nm) (<b>b</b>), and the corresponding Rrs (555 nm) (<b>c</b>) provided by the KOSC on 15:55 LT, 13 January 2021.</p>
Full article ">Figure 2
<p>Frequency distribution of solar zenith angle in the matchup dataset.</p>
Full article ">Figure 3
<p>Scatterplots of Rrs (λ) retrieved by the XGBAC model vs. the reference values in the training dataset at each GOCI band. The numerical values along the colour bar correspond to the pixel density on a logarithmic scale.</p>
Full article ">Figure 4
<p>Scatterplots of Rrs from the XGBAC model and the reference Rrs values of the validation dataset at each GOCI band. The numerical values along the colour bar correspond to the pixel density on a logarithmic scale.</p>
Full article ">Figure 5
<p>Scatterplots of Rrs (490 nm) retrieved from the XGBAC model vs. the values of the evaluation dataset for the different ranges of SZA. The numerical values along the colour bar correspond to the pixel density on a logarithmic scale.</p>
Full article ">Figure 6
<p>Variations in absolute percentage error of Rrs (490 nm) retrieved from the XGBAC and OC-SMART algorithms with the SZA.</p>
Full article ">Figure 7
<p>Frequency distribution of QA scores for the Rrs retrievals obtained using XGBAC and OC-SMART.</p>
Full article ">Figure 8
<p>Rrs maps at 443 nm, 555 nm and 680 nm derived by XGBAC from the GOCI data sensed on 13 January 2021 at 08:55 LT (<b>a</b>,<b>c</b>,<b>e</b>) and 15:55 LT (<b>b</b>,<b>d</b>,<b>f</b>); SZA in the 08:55 LT (<b>g</b>) and 15:55 LT (<b>h</b>).</p>
Full article ">Figure 9
<p>RPD between Rrs (555 nm) values in morning/afternoon hours and noontime observations at 12:55 LT using the XGBAC algorithm (<b>a</b>,<b>c</b>) and OC-SMART algorithm (<b>b</b>,<b>d</b>).</p>
Full article ">Figure 10
<p>The temporal CV of the pixel values examined using multiple noontime (10:55 LT–13:55 LT) Rrs (555 nm) values within a day.</p>
Full article ">Figure 11
<p>The hourly TSS maps in the YRE and HZB retrieved by Rrs product applying the XGBAC (<b>a</b>–<b>h</b>) and KOSC standard TSS products (<b>i</b>,<b>j</b>) on 13 January 2021.</p>
Full article ">
16 pages, 12318 KiB  
Article
Red Tide Detection Method Based on Improved U-Net Model-Taking GOCI Data in East China Sea as an Example
by Yanling Han, Tianhong Ding, Pengxia Cui, Xiaotong Wang, Bowen Zheng, Xiaojing Shen, Zhenling Ma, Yun Zhang, Haiyan Pan and Shuhu Yang
Sensors 2023, 23(22), 9195; https://doi.org/10.3390/s23229195 - 15 Nov 2023
Viewed by 1079
Abstract
In the coastal areas of China, the eutrophication of seawater leads to the continuous occurrence of red tide, which has caused great damage to Marine fisheries and aquatic resources. Therefore, the detection and prediction of red tide have important research significance. The rapid [...] Read more.
In the coastal areas of China, the eutrophication of seawater leads to the continuous occurrence of red tide, which has caused great damage to Marine fisheries and aquatic resources. Therefore, the detection and prediction of red tide have important research significance. The rapid development of optical remote sensing technology and deep-learning technology provides technical means for realizing large-scale and high-precision red tide detection. However, the difficulty of the accurate detection of red tide edges with complex boundaries limits the further improvement of red tide detection accuracy. In view of the above problems, this paper takes GOCI data in the East China Sea as an example and proposes an improved U-Net red tide detection method. In the improved U-Net method, NDVI was introduced to enhance the characteristic information of the red tide to improve the separability between the red tide and seawater. At the same time, the ECA channel attention mechanism was introduced to give different weights according to the influence of different bands on red tide detection, and the spectral characteristics of different channels were fully mined to further extract red tide characteristics. A shallow feature extraction module based on Atrous Spatial Pyramid Convolution (ASPC) was designed to improve the U-Net model. The red tide feature information in a multi-scale context was fused under multiple sampling rates to enhance the model’s ability to extract features at different scales. The problem of limited accuracy improvement in red tide edge detection with complex boundaries is solved via the fusion of deep and shallow features and multi-scale spatial features. Compared with other methods, the method proposed in this paper achieves better results and can detect red tide edges with complex boundaries, and the accuracy, precision, recall, and F1-score are 95.90%, 97.15%, 91.53%, and 0.94, respectively. In addition, the red tide detection experiments in other regions with relatively concentrated distribution also prove that the method has good applicability. Full article
(This article belongs to the Special Issue Feature Papers in Remote Sensors 2023)
Show Figures

Figure 1

Figure 1
<p>True color map of GOCI satellite coverage range.</p>
Full article ">Figure 2
<p>Visual results of chlorophyll concentration products from GOCI data. <span class="html-fig-inline" id="sensors-23-09195-i001"><img alt="Sensors 23 09195 i001" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i001.png"/></span> Red tide; <span class="html-fig-inline" id="sensors-23-09195-i002"><img alt="Sensors 23 09195 i002" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i002.png"/></span> Seawater; <span class="html-fig-inline" id="sensors-23-09195-i003"><img alt="Sensors 23 09195 i003" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i003.png"/></span> Land; <span class="html-fig-inline" id="sensors-23-09195-i004"><img alt="Sensors 23 09195 i004" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i004.png"/></span> Coastline; <span class="html-fig-inline" id="sensors-23-09195-i005"><img alt="Sensors 23 09195 i005" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i005.png"/></span> Unstudied region.</p>
Full article ">Figure 3
<p>Overall flow diagram of the improved U-Net model.</p>
Full article ">Figure 4
<p>ECA module.</p>
Full article ">Figure 5
<p>ASPP module.</p>
Full article ">Figure 6
<p>Improved ASPP module.</p>
Full article ">Figure 7
<p>Improved U-Net model structure.</p>
Full article ">Figure 8
<p>Region 1 red tide detection results based on different improved methods. (<b>a</b>) Actual red tide occurrence area, (<b>b</b>) Detection result with original six-band data, (<b>c</b>) Detection result with NDVI feature enhancement, (<b>d</b>) Detection result with the basic U-Net + ECA + ASPC. <span class="html-fig-inline" id="sensors-23-09195-i006"><img alt="Sensors 23 09195 i006" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i006.png"/></span> Red tide; <span class="html-fig-inline" id="sensors-23-09195-i007"><img alt="Sensors 23 09195 i007" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i007.png"/></span> Seawater; <span class="html-fig-inline" id="sensors-23-09195-i008"><img alt="Sensors 23 09195 i008" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i008.png"/></span> Land; <span class="html-fig-inline" id="sensors-23-09195-i009"><img alt="Sensors 23 09195 i009" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i009.png"/></span> Coastline; <span class="html-fig-inline" id="sensors-23-09195-i010"><img alt="Sensors 23 09195 i010" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i010.png"/></span> Unstudied region.</p>
Full article ">Figure 9
<p>Region 2 red tide detection results based on different methods. (<b>a</b>) Actual red tide occurrence area, (<b>b</b>) Detection result with original six-band data, (<b>c</b>) Detection result with enhanced features, (<b>d</b>) Detection result with the improved U-Net. <span class="html-fig-inline" id="sensors-23-09195-i011"><img alt="Sensors 23 09195 i011" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i011.png"/></span> Red tide; <span class="html-fig-inline" id="sensors-23-09195-i012"><img alt="Sensors 23 09195 i012" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i012.png"/></span> Seawater; <span class="html-fig-inline" id="sensors-23-09195-i013"><img alt="Sensors 23 09195 i013" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i013.png"/></span> Land; <span class="html-fig-inline" id="sensors-23-09195-i014"><img alt="Sensors 23 09195 i014" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i014.png"/></span> Coastline; <span class="html-fig-inline" id="sensors-23-09195-i015"><img alt="Sensors 23 09195 i015" src="/sensors/sensors-23-09195/article_deploy/html/images/sensors-23-09195-i015.png"/></span> Unstudied region.</p>
Full article ">
18 pages, 9893 KiB  
Article
Quantitative Retrieval of Chlorophyll-a Concentrations in the Bohai–Yellow Sea Using GOCI Surface Reflectance Products
by Jiru Wang, Jiakui Tang, Wuhua Wang, Yanjiao Wang and Zhao Wang
Remote Sens. 2023, 15(22), 5285; https://doi.org/10.3390/rs15225285 - 8 Nov 2023
Cited by 7 | Viewed by 1861
Abstract
As an environmental parameter, the chlorophyll-a concentration (Chl-a) is essential for monitoring water quality and managing the marine ecosystem. However, current mainstream Chl-a inversion algorithms have limited accuracy and poor spatial and temporal generalization in Case II waters. In this study, we constructed [...] Read more.
As an environmental parameter, the chlorophyll-a concentration (Chl-a) is essential for monitoring water quality and managing the marine ecosystem. However, current mainstream Chl-a inversion algorithms have limited accuracy and poor spatial and temporal generalization in Case II waters. In this study, we constructed a quantitative model for retrieving the spatial and temporal distribution of Chl-a in the Bohai–Yellow Sea area using Geostationary Ocean Color Imager (GOCI) spectral remote sensing reflectance (Rrsλ) products. Firstly, the GOCI Rrsλ correction model based on measured spectral data was proposed and evaluated. Then, the feature variables of the band combinations with the highest correlation with Chl-a were selected. Subsequently, Chl-a inversion models were developed using three empirical ocean color algorithms (OC4, OC5, and YOC) and four machine learning methods: BP neural network (BPNN), random forest (RF), AdaBoost, and support vector regression (SVR). The retrieval results showed that the machine learning methods were much more accurate than the empirical algorithms and that the RF model retrieved Chl-a with the best performance and the highest prediction accuracy, with a determination coefficient R2 of 0.916, a root mean square error (RMSE) of 0.212 mg·m−3, and a mean absolute percentage error (MAPE) of 14.27%. Finally, the Chl-a distribution in the Bohai–Yellow Sea using the selected RF model was derived and analyzed. Spatially, Chl-a was high in the Bohai Sea, including in Laizhou Bay, Bohai Bay, and Liaodong Bay, with a value higher than 4 mg·m−3. Chl-a in the Bohai Strait and northern Yellow Sea was relatively low, with a value of less than 3 mg·m−3. Temporally, the inversion results showed that Chl-a was considerably higher in winter and spring compared to autumn and summer. Diurnal variation retrieval effectively demonstrated GOCI’s potential as a capable tool for monitoring intraday changes in chlorophyll-a concentrations. Full article
(This article belongs to the Special Issue Validation and Evaluation of Global Ocean Satellite Products)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of sample points including AERONET-OC and collected in situ Chl-a. Annotated on the image are the locations of the sampling points: AERONET-OC (which is used for GOCI spectral remote sensing reflectance correction and Chl-a retrieval), other Chl-a points measured in 2012 and 2017 (only used for Chl-a retrieval).</p>
Full article ">Figure 2
<p>The flowchart of Chl-a retrieval modeling.</p>
Full article ">Figure 3
<p>Compariation of the corrected GOCI <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> <mfenced> <mi>λ</mi> </mfenced> </mrow> </semantics></math> products against AERONET-OC field measurements: (<b>a</b>–<b>f</b>) represent scatter plots of the six bands before and after correction, Green squares mean before correction, and blue dots represent after correction (the number of scatter dots in each color N = 360).</p>
Full article ">Figure 4
<p>Boxplot representing the R<sup>2</sup> metric obtained on cross-validation; the mean value is marked with a diamond.</p>
Full article ">Figure 5
<p>Scatter verification of Chl-a values estimated from four models and measured values, the x-coordinate is in situ Chl-a, and the y-coordinate is model estimated Chl-a, both in units of mg<math display="inline"><semantics> <mo>·</mo> </semantics></math>m<sup>−3</sup>, (<b>a</b>) Random forest. (<b>b</b>) SVR. (<b>c</b>) BPNN. (<b>d</b>) AdaBoost.</p>
Full article ">Figure 5 Cont.
<p>Scatter verification of Chl-a values estimated from four models and measured values, the x-coordinate is in situ Chl-a, and the y-coordinate is model estimated Chl-a, both in units of mg<math display="inline"><semantics> <mo>·</mo> </semantics></math>m<sup>−3</sup>, (<b>a</b>) Random forest. (<b>b</b>) SVR. (<b>c</b>) BPNN. (<b>d</b>) AdaBoost.</p>
Full article ">Figure 6
<p>Comparison of in situ and GOCI estimated Chl-a for three AERONET-OC stations: (<b>a</b>) Socheongcho station; (<b>b</b>) Gageocho station; and (<b>c</b>) Ieodo station.</p>
Full article ">Figure 7
<p>January–December 2019 spatial distribution of monthly mean Chl-a estimated from GOCI using the RF model.</p>
Full article ">Figure 8
<p>The histogram distributions of GOCI-estimated Chl-a for different seasons during 2019.</p>
Full article ">Figure 9
<p>Spatial distribution of Chl-a at eight moments estimated from GOCI on 15 April 2019.</p>
Full article ">Figure 10
<p>Diurnal variation curves of chlorophyll concentration and <math display="inline"><semantics> <mrow> <msub> <mi>R</mi> <mrow> <mi>r</mi> <mi>s</mi> </mrow> </msub> <mfenced> <mi>λ</mi> </mfenced> </mrow> </semantics></math> at three typical sites on 15 April 2019. (<b>a</b>–<b>c</b>) represent three randomly selected typical sites, (<b>a1</b>–<b>c1</b>) are the Chl-a diurnal variation for the corresponding sites, and (<b>a2</b>–<b>c2</b>) represent the spectral curves at the corresponding time for each site.</p>
Full article ">
18 pages, 17566 KiB  
Article
Downscaling of Oceanic Chlorophyll-a with a Spatiotemporal Fusion Model: A Case Study on the North Coast of the Yellow Sea
by Qingdian Meng, Jun Song, Yanzhao Fu, Yu Cai, Junru Guo, Ming Liu and Xiaoyi Jiang
Water 2023, 15(20), 3566; https://doi.org/10.3390/w15203566 - 12 Oct 2023
Cited by 3 | Viewed by 1065
Abstract
Chlorophyll-a concentration (Chl-a) is an important indicator of coastal eutrophication. Remote sensing technology provides a global view of it. However, different types of sensors are subject to design constraints and cannot meet the requirements of high temporal and spatial resolution on nearshore engineering [...] Read more.
Chlorophyll-a concentration (Chl-a) is an important indicator of coastal eutrophication. Remote sensing technology provides a global view of it. However, different types of sensors are subject to design constraints and cannot meet the requirements of high temporal and spatial resolution on nearshore engineering simultaneously. To obtain high-spatiotemporal-resolution images, this study examines the performance of the enhanced spatial and temporal adaptive reflectance fusion model (ESTARFM) on GOCI and Landsat Chl-a data fusion. Considering the rapidly changing rate and consistency of oceanic Chl-a, the ESTARFM was modified via segmented fitting and numerical conversion. The results show that both fusion models can fuse multiple data advantages to obtain high-spatiotemporal-resolution Chl-a images. Compared with the ESTARFM, the modified solution has a better performance in terms of the root mean square error and correlation coefficient, and its results have better spatial consistency for coastal Chl-a. In addition, the new solution expands the data utilization range of data fusion by reducing the influence of the time interval of original data and realizes better monitoring of nearshore Chl-a changes. Full article
(This article belongs to the Special Issue Emerging Challenges in Ocean Engineering and Environmental Effects)
Show Figures

Figure 1

Figure 1
<p>Study area: Changhai County and Shicheng Island in the northern part of the North Yellow Sea.</p>
Full article ">Figure 2
<p>Overall workflow in this study.</p>
Full article ">Figure 3
<p>The same spatial index is used to process GOCI and Landsat data.</p>
Full article ">Figure 4
<p>Scatter plot comparing the Landsat data with the GOCI and results of the two models.</p>
Full article ">Figure 5
<p>Comparison of the results derived from the images of 8 October and 11 December.</p>
Full article ">Figure 6
<p>Details of the results of ESTARFM and ESTARFM_p.</p>
Full article ">Figure 7
<p>Fifteen sets of Chl-a results for 22 September derived using the ESTARFM.</p>
Full article ">Figure 8
<p>Fifteen sets of Chl-a results for 22 September derived using the ESTARFM-p model.</p>
Full article ">Figure 9
<p>The statistical coefficients of the results of ESTARFM, ESTARFM_p, and ESTARFM_p with single-date inputs (p_one).</p>
Full article ">Figure 10
<p>Detailed comparison of the areas near Shicheng Island and Wangjia Island.</p>
Full article ">Figure 11
<p>The variation in <span class="html-italic">RMSE</span> and <span class="html-italic">CC</span> of ESTARFM, ESTARFM_p, and ESTARFM_p-one results with time intervals.</p>
Full article ">Figure 12
<p>Scatter plots of the Landsat and results of ESTARFM and ESTARFM_p for green (<b>a</b>), red (<b>b</b>), and NIR-infrared bands (<b>c</b>).</p>
Full article ">Figure 13
<p>Comparison of Landsat actual reflectance image, ESTARFM_p, and ESTARFM results.</p>
Full article ">
21 pages, 11473 KiB  
Article
Retrievals of Chlorophyll-a from GOCI and GOCI-II Data in Optically Complex Lakes
by Yuyu Guo, Xiaoqi Wei, Zehui Huang, Hanhan Li, Ronghua Ma, Zhigang Cao, Ming Shen and Kun Xue
Remote Sens. 2023, 15(19), 4886; https://doi.org/10.3390/rs15194886 - 9 Oct 2023
Cited by 3 | Viewed by 1420
Abstract
The chlorophyll-a (Chla) concentration is a key parameter to evaluate the eutrophication conditions of water, which is very important for monitoring algal blooms. Although Geostationary Ocean Color Imager (GOCI) has been widely used in Chla inversion, the consistency of the [...] Read more.
The chlorophyll-a (Chla) concentration is a key parameter to evaluate the eutrophication conditions of water, which is very important for monitoring algal blooms. Although Geostationary Ocean Color Imager (GOCI) has been widely used in Chla inversion, the consistency of the Rayleigh-corrected reflectance (Rrc) of GOCI and GOCI-II sensors still needs to be further evaluated, and a model suitable for lakes with complex optical properties needs to be constructed. The results show that (1) the derived Chla values of the GOCI and GOCI-II synchronous data were relatively consistent and continuous in three lakes in China. (2) The accuracy of the random forest (RF) model (R2 = 0.84, root mean square error (RMSE) =11.77 μg/L) was higher than that of the empirical model (R2 = 0.79, RMSE = 12.63 μg/L) based on the alternative floating algae index (AFAI). (3) The interannual variation trend fluctuated, with high Chla levels in Lake Chaohu in 2015 and 2019, while those in Lake Hongze were high in 2013, 2015, and 2022, and those in Lake Taihu reached their peak in 2017 and 2019. There were three types of diurnal variation patterns, namely, near-continuous increase (Class 1), near-continuous decrease (Class 2), and first an increase and then a decrease (Class 3), among which Lake Chaohu and Lake Taihu occupied the highest proportion in Class 3. The results analyzed the temporal and spatial variations of Chla in three lakes for 12 years and provided support for the use of GOCI and GOCI-II data and monitoring of Chla in optical complex inland waters. Full article
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Location of Lake Chaohu, Lake Taihu, and Lake Hongze in China. The spatial distributions of samples in (<b>b</b>) Lake Hongze, (<b>c</b>) Lake Chaohu, and (<b>d</b>) Lake Taihu are shown.</p>
Full article ">Figure 2
<p>Spectral response functions of (<b>a</b>) GOCI and (<b>b</b>) GOCI-II in different spectral ranges. Note that the bands marked with * are new bands of GOCI-II.</p>
Full article ">Figure 3
<p>Number of images of Lake Chaohu, Lake Hongze, and Lake Taihu in different months.</p>
Full article ">Figure 4
<p>Consistent verification results for different bands based on GOCI and GOCI-II data.</p>
Full article ">Figure 5
<p>Training and validation of the inversion model for the dominant factor (<b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>), and (<b>m</b>) show the training results for the quadratic polynomial equation with five factors, namely, the AFAI, B7/B5, B7/B6, FLH, and SI, respectively; (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>), and (<b>n</b>) show the training results for the exponential equation; (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>), and (<b>o</b>) show the validation results for the exponential equation.</p>
Full article ">Figure 6
<p>Training and validation of the RF model on Chl<span class="html-italic">a</span> estimation: (<b>a</b>) importance of seven variables; (<b>b</b>) training; and (<b>c</b>) validation.</p>
Full article ">Figure 7
<p>Yearly mean Chl<span class="html-italic">a</span> from 2011 to 2022 in (<b>a</b>) Lake Chaohu, (<b>b</b>) Lake Hongze, and (<b>c</b>) Lake Taihu. Note that the GOCI data for 2011 started in May.</p>
Full article ">Figure 8
<p>Time series of the monthly mean Chl<span class="html-italic">a</span> of GOCI and GOCI-II from 2011 to 2022 in (<b>a</b>) Lake Chaohu, (<b>c</b>) Lake Hongze, and (<b>e</b>) Lake Taihu. Daily mean Chl<span class="html-italic">a</span> of GOCI and GOCI-II from January to March 2021 in (<b>b</b>) Lake Chaohu, (<b>d</b>) Lake Hongze, and (<b>f</b>) Lake Taihu.</p>
Full article ">Figure 9
<p>Monthly mean Chl<span class="html-italic">a</span> in (<b>a</b>) Lake Chaohu, (<b>b</b>) Lake Hongze, and (<b>c</b>) Lake Taihu.</p>
Full article ">Figure 10
<p>Three typical types of Chl<span class="html-italic">a</span> diurnal variation patterns in the three lakes: (<b>a</b>) shows images of Class 1 (near-continuous increase), (<b>b</b>) shows images of Class 2 (near-continuous decrease), (<b>c</b>) shows images of Class 3 (first an increase and then a decrease), and (<b>d</b>–<b>f</b>) show corresponding images for three dates. (<b>a1</b>–<b>f1</b>) show Lake Chaohu, (<b>a2</b>–<b>f2</b>) show Lake Hongze, and (<b>a3</b>–<b>f3</b>) show Lake Taihu.</p>
Full article ">Figure 11
<p>Error of the models with different numbers of samples: (<b>a</b>) R<sup>2</sup>, (<b>b</b>) RMSE, (<b>c</b>) MAPE, and (<b>d</b>) UPD.</p>
Full article ">Figure 12
<p>Location of the samples in (<b>a</b>) Lake Hongze and (<b>b</b>) validation of the models.</p>
Full article ">
18 pages, 4723 KiB  
Article
Spectral and Spatial Dependencies in the Validation of Satellite-Based Aerosol Optical Depth from the Geostationary Ocean Color Imager Using the Aerosol Robotic Network
by Mijeong Kim, Kyunghwa Lee and Myungje Choi
Remote Sens. 2023, 15(14), 3621; https://doi.org/10.3390/rs15143621 - 20 Jul 2023
Viewed by 1271
Abstract
The regional and global scale of aerosols in the atmosphere can be quantified using the aerosol optical depth (AOD) retrieved from satellite observations. To obtain reliable satellite AODs, conducting consistent validations and refining retrieval algorithms are crucial. AODs and Ångström exponents (AEs) measured [...] Read more.
The regional and global scale of aerosols in the atmosphere can be quantified using the aerosol optical depth (AOD) retrieved from satellite observations. To obtain reliable satellite AODs, conducting consistent validations and refining retrieval algorithms are crucial. AODs and Ångström exponents (AEs) measured with the aerosol robotic network (AERONET) are considered as the ground truth for satellite validations. AERONET AEs are used to collocate the wavelength of the AERONET AODs to those of the satellite AODs when there is a discordancy in their wavelengths. However, numerous validation studies have proposed different strategies by applying the AERONET AODs and AEs, and spatiotemporal collocation criteria. This study examined the impact of the wavelength and spatial collocation radius variations by comparing AODs at 550 nm derived from the geostationary ocean color imager (GOCI) with those obtained from the AERONET for the year 2016. The estimated AERONET AODs at 550 nm varied from 5.18% to 11.73% depending on the selection of AOD and AE, and the spatial collocation radii from 0 to 40 km, respectively. The longer the collocation radius and the higher the AODs, the greater the variability observed in the validation results. Overall, the selection of the spatial collocation radius had a stronger impact on the variability in the validation results obtained compared to the selection of the wavelength. The variability was also found in seasonal analysis. Therefore, it is recommended to carefully select the data wavelength and spatial collocation radius, consider seasonal effects, and provide this information when validating satellite AODs using AERONET. Full article
(This article belongs to the Section Atmospheric Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) Annually averaged geostationary ocean color imager (GOCI) aerosol optical depth (AOD) in 2016 and locations of the aerosol robotic network (AERONET) stations (purple circles). The number of data points of (<b>b</b>) the AOD and (<b>c</b>) the Ångström Exponent (AE) in the year of 2016 over the GOCI observation area.</p>
Full article ">Figure 2
<p>Schematic diagram of the AOD and AE pairs of AERONET that were used to calculate the AOD at 550 nm for validating the GOCI AOD at 550 nm. Solid and dashed lines are pairs using AE440-675 and AE440-870, respectively.</p>
Full article ">Figure 3
<p>Statistical results for retrieving AERONET AODs at 550 nm using AE440-675 and AE440-870 depending on various collocation radii (<b>a</b>,<b>c</b>,<b>e</b>) and wavelengths (<b>b</b>,<b>d</b>,<b>f</b>) using the original AERONET AODs to validate the GOCI AODs at 550 nm in 2016.</p>
Full article ">Figure 4
<p>Seasonal analysis of the validation results of GOCI AODs at 550 nm using AERONET AODs at 550 nm based on the collocation radii in (<b>a</b>,<b>e</b>,<b>i</b>) winter (December–January–February), (<b>b</b>,<b>f</b>,<b>j</b>) spring (March–April–May), (<b>c</b>,<b>g</b>,<b>k</b>) summer (June–July–August), and (<b>d</b>,<b>h</b>,<b>l</b>) fall (September–October–November) in the year of 2016.</p>
Full article ">Figure 5
<p>Seasonal analysis of the validation results of GOCI AODs at 550 nm using AERONET AODs at 550 nm based on the AOD wavelengths of AERONET in (<b>a</b>,<b>e</b>,<b>i</b>) winter (December–January–February), (<b>b</b>,<b>f</b>,<b>j</b>) spring (March–April–May), (<b>c</b>,<b>g</b>,<b>k</b>) summer (June–July–August), and (<b>d</b>,<b>h</b>,<b>l</b>) fall (September–October–November) in the year of 2016.</p>
Full article ">Figure 6
<p>Scatter plots of the correlation coefficient (R) and root mean square error (RMSE) between the GOCI and AERONET AODs at 550 nm, colored by collocation radii for (<b>a</b>) annual and (<b>b</b>) seasonal results in the year of 2016.</p>
Full article ">Figure 7
<p>Scatter plots of R against the RMSE between the GOCI and AERONET AODs at 550 nm, colored by the collocation radius for annual (left) and seasonal (right) results for (<b>a</b>) AOD ≥ 0, (<b>b</b>) AOD ≥ 0.5, and (<b>c</b>) AOD ≥ 1.0, respectively.</p>
Full article ">Figure 8
<p>The number of collocated points used for calculating the statistics displayed in <a href="#remotesensing-15-03621-f007" class="html-fig">Figure 7</a> for (<b>a</b>) AOD ≥ 0, (<b>b</b>) AOD ≥ 0.5, and (<b>c</b>) AOD ≥ 1.0, respectively.</p>
Full article ">Figure 9
<p>Color matrices of the averaged and collocated AODs depending on the collocation radii and wavelength pairs for (<b>a</b>) GOCI, (<b>b</b>) AERONET, and (<b>c</b>) the difference rate between them for all AOD values.</p>
Full article ">Figure 10
<p>Statistical matrices of the correlation coefficient (R) (<b>a</b>,<b>d</b>,<b>g</b>), bias (<b>b</b>,<b>e</b>,<b>h</b>), and root mean square error (RMSE) (<b>c</b>,<b>f</b>,<b>i</b>) for AOD interpolation, arranged depending on the collocation radii, spectral combinations, and the range of AOD values.</p>
Full article ">
22 pages, 6214 KiB  
Article
Retrieval of Chlorophyll a Concentration Using GOCI Data in Sediment-Laden Turbid Waters of Hangzhou Bay and Adjacent Coastal Waters
by Yixin Yang, Shuangyan He, Yanzhen Gu, Chengyue Zhu, Longhua Wang, Xiao Ma and Peiliang Li
J. Mar. Sci. Eng. 2023, 11(6), 1098; https://doi.org/10.3390/jmse11061098 - 23 May 2023
Cited by 1 | Viewed by 1617
Abstract
The Geostationary Ocean Color Imager (GOCI) provided images at hourly intervals up to 8 times per day with a spatial resolution of 500 m from 2011 to 2021. However, in the typical sediment-laden turbid water of Hangzhou Bay, valid ocean color parameters in [...] Read more.
The Geostationary Ocean Color Imager (GOCI) provided images at hourly intervals up to 8 times per day with a spatial resolution of 500 m from 2011 to 2021. However, in the typical sediment-laden turbid water of Hangzhou Bay, valid ocean color parameters in operational data products have been extensively missing due to failures in atmospheric correction (AC) and bio-optical retrieval procedures. In this study, the seasonal variations in chlorophyll a (Chl-a) concentrations in Hangzhou Bay derived using GOCI data in 2020 were presented. First, valid remote sensing reflectance data were obtained by transferring neighboring aerosol properties of less to more turbid water pixels. Then, we improved a regionally empirical Chl-a retrieval algorithm in extremely turbid waters using GOCI-derived surface reflectance and field Chl-a measurements and proposed a combined Chl-a retrieval scheme for both moderately and extremely turbid water in Hangzhou Bay. Finally, the seasonal variation in Chl-a was obtained by the GOCI, which was better than operational products and in good agreement with the buoy data. The method in this study can be effectively applied to the inversion of Chl-a concentration in Hangzhou Bay and adjacent sea areas. We also presented its seasonal variations, offering insight into the spatial and seasonal variation of Chl-a in Hangzhou Bay using the GOCI. Full article
(This article belongs to the Section Marine Environmental Science)
Show Figures

Figure 1

Figure 1
<p>Location of the study area. Hangzhou Bay is indicated in a green rectangle. The two red stars (A and B) indicate the locations of the two buoys used in this study. The 26 triangles (in yellow and red) indicate site locations of cruises (6 match-up sites in red were used in this study). The background color represents the water depth.</p>
Full article ">Figure 2
<p>(<b>a</b>) RGB image composited from GOCI Rayleigh-corrected reflectance at 680, 555, and 443 nm at 05:16 (UTC) on 10 February 2020. (<b>b</b>) The corresponding chlorophyll a concentration retrieved by the OC3 algorithm using GDPS 2.0, and the white color represents no valid data. Hangzhou Bay is indicated by a green solid line rectangle, and the adjacent coastal waters are indicated by a green dashed rectangle.</p>
Full article ">Figure 3
<p>Comparison of GOCI spectral <span class="html-italic">R<sub>rs</sub></span> data (at 555, 660, 680 nm) obtained using three windows of 1 × 1, 3 × 3 and 5 × 5 pixels. (<b>a</b>) <span class="html-italic">R<sub>rs</sub></span> values of the 1 × 1 vs. 3 × 3 pixel window. (<b>b</b>) <span class="html-italic">R<sub>rs</sub></span> values of the 1 × 1 vs. 5 × 5 pixel window. (<b>c</b>) <span class="html-italic">R<sub>rs</sub></span> values of the 3 × 3 vs. 5 × 5 pixel window.</p>
Full article ">Figure 4
<p>(<b>a</b>) Relationship between the <span class="html-italic">C<sub>Sediments</sub></span> and <span class="html-italic">R<sub>rs</sub> </span>(745 nm)/<span class="html-italic">R<sub>rs</sub> </span>(490 nm) band ratio. (<b>b</b>) The match-up GOCI-derived <span class="html-italic">R<sub>rs</sub></span> spectrum (in blue) in moderately turbid waters (<span class="html-italic">C<sub>Sediments</sub></span> from 16.51 to 39.02 mg/L). (<b>c</b>) The match-up GOCI-derived <span class="html-italic">R<sub>rs</sub></span> spectrum (in pink) for the extremely turbid waters (<span class="html-italic">C<sub>Sediments</sub></span> from 40.05 to 1761.32 mg/L) in Hangzhou Bay.</p>
Full article ">Figure 5
<p>Comparison of valid <span class="html-italic">R<sub>rs</sub></span> data pixels from the KOSC’s GOCI level 2 products and this study.</p>
Full article ">Figure 6
<p>Comparison between GOCI <span class="html-italic">R<sub>rs</sub></span> level 2 products (blue triangles) and GOCI <span class="html-italic">R<sub>rs</sub></span> from this study (red dots). The results in less turbid water at 02:16 UTC on 1 February 2020 (<b>a</b>,<b>b</b>) and at 06:16 UTC on 21 December 2020 (<b>c</b>,<b>d</b>), in turbid water at 02:16 UTC on 19 June 2020 (<b>e</b>,<b>f</b>) and at 03:16 UTC on 12 October 2020 (<b>g</b>,<b>h</b>). The red crosses (a,c,e,g) indicate locations where <span class="html-italic">R<sub>rs</sub></span> spectrums (b,d,f,h) were obtained.</p>
Full article ">Figure 7
<p>Scatter plot between field Chl-a and the retrieval results from GOCI data. (<b>a</b>) Calibration of 94 match-ups (in pink) in extremely turbid waters, (<b>b</b>) validation of 41 match-ups (in blue) in extremely turbid waters, (<b>c</b>) scatter plots of Chl-a between field data and results derived using OC3, and (<b>d</b>) scatter plots of Chl-a between field data and results derived using the improved method in this study. The retrievals obtained in extremely turbid waters and moderately turbid waters are indicated in red and grey, respectively.</p>
Full article ">Figure 8
<p>Seasonal averaged maps of Chl-a in Hangzhou Bay and adjacent coastal waters. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) represent the Chl-a seasonal variation derived by the GOCI level 2 data products of the KOSC, and (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) are the Chl-a seasonal variation derived after atmospheric correction and the improved combined regional Chl-a retrieval algorithm. Hangzhou Bay is indicated in a red rectangle.</p>
Full article ">Figure 8 Cont.
<p>Seasonal averaged maps of Chl-a in Hangzhou Bay and adjacent coastal waters. (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>) represent the Chl-a seasonal variation derived by the GOCI level 2 data products of the KOSC, and (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>) are the Chl-a seasonal variation derived after atmospheric correction and the improved combined regional Chl-a retrieval algorithm. Hangzhou Bay is indicated in a red rectangle.</p>
Full article ">Figure 9
<p>Seasonal mean values of Chl-a from buoys and GOCI data in Hangzhou Bay (HZB).</p>
Full article ">Figure 10
<p>The Chl-a seasonal variation in the study area retrieved by the GOCI. (<b>a</b>) The location of the selected representative transect indicated by A to E in the study area. (<b>b</b>) The seasonal variations in Chl-a at the representative transect (the dashed line represents the GOCI Chl-a of the KOSC, and the solid line is the Chl-a of this study).</p>
Full article ">Figure 11
<p>Different Chl-a seasonal cycle types in 2020 in the study area: (<b>a</b>) distribution of four Chl-a seasonal cycle types: Hangzhou Bay and its adjacent areas (P area, in pink), nearshore area (Q area, in green), offshore area (M area, in red), and open sea (N area, in blue); (<b>b</b>) seasonal Chl-a variations for different types (the pink, green, red and blue lines represent the P, Q, M and N areas, respectively).</p>
Full article ">Figure 12
<p>Chl-a comparison among the 8 h multitemporal (in blue), dual-temporal (in yellow), and single-time (in pink) observations and the 24 h daily (in black) averages. (<b>a</b>) Comparison of the time series in 2020. (<b>b</b>) Comparison of the time series from 30 September to 7 October 2020. (<b>c</b>) The mean average percentage errors of three temporal observations (8 h multitemporal in blue, dual-temporal in yellow, single-time in pink) in four seasons compared with the 24 h daily averages.</p>
Full article ">
Back to TopTop