Next Article in Journal
Multi-Branch Attention Fusion Network for Cloud and Cloud Shadow Segmentation
Previous Article in Journal
Retrospective Analysis of Glacial Lake Outburst Flood (GLOF) Using AI Earth InSAR and Optical Images: A Case Study of South Lhonak Lake, Sikkim
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evaluation of Rayleigh-Corrected Reflectance on Remote Detection of Algal Blooms in Optically Complex Coasts of East China Sea

1
States Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, Ministry of Natural Resource, Hangzhou 310012, China
2
Institute of Oceanograhic Instrumentation, Qilu University of Technology (Shandong Academy of Sciences), Qingdao 266100, China
3
Marine College, Shandong University (Weihai), Weihai 264200, China
4
Wenzhou Marine Center, Ministry of Natural Resource, Wenzhou 325000, China
5
Ocean College, Zhejiang University, Zhoushan 316021, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2024, 16(13), 2304; https://doi.org/10.3390/rs16132304
Submission received: 6 April 2024 / Revised: 7 June 2024 / Accepted: 15 June 2024 / Published: 24 June 2024

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), 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.

1. Introduction

Harmful algal blooms (HABs) are a well-known ecological anomaly characterized by rapid proliferation and microplankton aggregation, causing increasingly serious ecological problems along coastal areas worldwide [1,2]. Over the past decade, substantial progress has been achieved in enhancing the spatial and temporal resolution of ocean color satellite remote sensing, enabling the synoptic detection and characterization of the location and extent of HABs from space [3,4,5,6,7,8,9,10]. Several approaches have been developed to detect algal blooms, such as the chlorophyll a concentration (Chla)-based approach [7,8,11,12] and the reflectance-based indices method [13,14]. Furthermore, over the past decade, studies have also advanced in identifying the types or causative species of phytoplankton blooms [4,15,16,17,18,19,20,21,22,23]. For example, Cannizzaro et al. [4] utilized the satellite-derived backscattering/Chla ratio to distinguish Karenia brevis blooms from other blooms in the coastal waters of Florida. Kim et al. [24] explored the potential for optically distinguishing Cochlodinium polykrikoides blooms from non-dinoflagellate blooms based on two band ratios in Korean coastal areas. Notably, all these outcomes rely on the fundamental product of remote sensing reflectance (Rrs) derived from ocean color sensors.
However, the reliable retrieving Rrs over optically complex coastal or estuarine regions still remains challenging [25,26,27,28]. For example, despite the near-daily coverage of moderate resolution imaging spectroradiometer (MODIS) measurements and a ~25–30% cloud free probability over global oceans [29], the probability of obtaining high-quality Rrs or Chla retrievals at the pixel scale is only ~5% [30]. This limited spatial and temporal coverage of MODIS observations provides timely information on phytoplankton bloom dynamics, but other ocean color sensors also face similar challenges. The main causes of such problems are significant uncertainties that occurred in atmospheric correction schemes, including incorrect presumptions such as the assumption of a negligible water signal in the near-infrared and the relationship between different spectral bands used in the iterative atmospheric correction.
Instead of using Rrs data, the Rayleigh-corrected reflectance (Rrc), which is obtained by correcting the atmospheric path radiance received by a sensor at the top-of-atmosphere (TOA) to remove the effects of ozone absorption and molecular (Rayleigh) scattering, proved to be useful in the monitoring of harmful algal blooms (HABs) [29,31,32,33,34,35]. The benefits of using Rrc include (1), the Rrc can be accurately derived since the Rayleigh component can be accurately computed, even when considering polarization effects [25] and (2), Rrc is less susceptible to atmospheric correction failures so that it can then significantly improve data coverage without the need for complex aerosol removal steps [36,37]. Additionally, the coverage of Rrc can even extend to the severe sun glint regions, allowing for the successful derivation of color patterns [33]. Due to the advantages of Rrc, Hu (2010) [32] introduced a MODIS floating algae index (FAI), defined as the difference between Rrc in the near-infrared (NIR) and a baseline formed by the red and short-wave infrared (SWIR) bands, to effectively detect phytoplankton in the ocean. Additionally, Hu (2011) also proposed another Rrc-based color index (CI) to successfully derive the color patterns even under severe sun glint conditions, and recently, this index has been converted into a high spatial coverage Chla product through machine learning [33]. Another example is the maximum peak height (MPH) algorithm, which quantifies the phytoplankton biomass by measuring the height above a spectral baseline established linearly between Rrc at 664 nm and 885 nm, specifically where the maximal Rrc occurs at 681, 709, or 753 nm [38,39]. Recently, Rrc has also shown some potential in the classification of algae species, especially for some macroalgal blooms such as Sargassum and Ulva [40,41]. Cannizzaro et al. (2019) [42] demonstrated that Cyanobacteria index and a spectral shape around 488 nm (SS488) derived from Rrc showed strong correlations with those derived from Rrs, and the estimated uncertainty of SS488 (Rrc) will minimally impact the accuracies of classifying the picocyanobacteria blooms from diatom ones in an optically shallow lagoonal estuary of Florida Bay. However, there is a lack comprehensive analysis on the ability of these methods to meet the needs of operational algal bloom monitoring in the ECS.
Similarly, in the ECS, microalgal blooms of diatoms or dinoflagellate [43,44,45,46] occur frequently every year. Til present Rrs data which still remain pivotal in the microalgal bloom detection. Shang et al. (2014) [44] initially proposed a novel method using the Bloom Index (BI), FLH, and a(443) to distinguish blooms. Tao et al. (2015) [47] employed two MODIS land bands to devise two indices: Prorocentrum donghaiense (P. donghaiense) index (PDI) and diatom index (DI), aimed at discriminating between the two species. Although a modified algal bloom index (RI) derived from the Rrc data, as proposed by Lou and Hu (2014) [43], can effectively delineate the P. donghaiense bloom in sediment-rich waters, other indices (e.g., fluorescence line height based on Rrc and CIE color represented by Rrc) were also proved to be effective [35,48,49]. Nevertheless, only a few studies on the comprehensive evaluation of Rrc data in identifying various species-dominated algal bloom can be found, as optical properties among harmful bloom species can vary, limiting the applicability of most bio-optical methods to specific species [50].
To address the lack of evaluation of Rrc in identifying microalgal blooms in the ECS, this study aims to (1) evaluate the spectral features of Rrc for various optical water types, (2) develop a synthetical Rrc-based baseline algorithm for identifying microalgal blooms in the coastal areas of the ECS, and (3) validate different Rrc baseline indices in bloom detections. By comparing Rrc and Rrs in identifying algal blooms in the ECS, this study comprehensively examines the strengths and weaknesses of the Rrc-based approach, ultimately aiming to recommend the most effective Rrc baseline index for accurately identifying microalgal blooms in the ECS.

2. Data and Methods

2.1. Study Area

This study centers around the ECS, specifically within the coordinates of 26–29°N and 119.5–122.5°E, encompassing most of the coastlines of Zhejiang Province as depicted in Figure 1. Given that the majority of the seabed in this area is less than 50 m deep, the region is greatly influenced by dynamic and intricate physical and biogeochemical oceanographic processes [51,52]. The ECS is biologically productive, with frequent annual occurrences of algal blooms, encompassing both harmful and non-harmful micro- and macro-algae [45,53,54,55]. As a result of its optically complex waters, atmospheric correction failures often leads to substantial data gaps. Improving data coverage by using Rrc is therefore important in assessing long-term bloom monitoring.

2.2. GOCI-II Data

The Geostationary Ocean Color Imager-II (GOCI-II), a maritime sensor located on the geostationary satellite Geo-Kompsat-2B (GK-2B), replaces the previous GOCI and provides comprehensive monitoring coverage of the entire sector, including Northeast Asia. GOCI-II captures images at a spatial resolution of 250 m, capturing up to 10 images daily with hourly intervals from 23 UTC to 8 UTC. Its local modes encompass 12 slots, collectively covering an extensive area of 2500 km × 2500 km across Northeast Asia, among which slot 9 encompassing our study area as depicted in Figure 1. Equipped with 12 spectral bands, GOCI-II facilitates with 1 near-ultraviolet (UV) band (380 nm), 8 VIS bands (412, 443, 490, 510, 555, 620, 660, and 680 nm), and 3 NIR bands (709, 745, and 865 nm) for ocean environmental monitoring. Particularly, GOCI-II features three new bands centered at 510 nm, 620 nm, and 709 nm, providing enhanced wavelength information to discern the optical properties difference among various types of blooms [46,50,56,57]. For the purpose of this study, the GOCI-II Level-2 data, obtained from the Korea Hydrographic and Oceanographic Agency (KHOA), were employed in bloom detection. These data include both Rayleigh-corrected reflectance and fully atmospherically corrected remote sensing reflectance data across 12 spectral bands from 380 to 865 nm. Briefly, Rrc was derived as Equation (1):
R r c , λ = π L T O A , λ C o r r F 0 , λ cos θ s R r , λ
where L T O A , λ C o r r represents the calibrated top-of-atmosphere radiance at band λ after accounting for gaseous absorption and whitecap correction, F0,λ is the extraterrestrial solar irradiance at the data acquisition time, θs is the solar zenith angle, and Rr is the reflectance due to Rayleigh (molecular) scattering. GOCI-II Rrc and Rrs data were generated using the GOCI-II atmospheric correction algorithm (G2AC) [58,59].
For cloud mask, an empirical method was employed to flag cloud pixels by utilizing the relationship between neighboring spectral bands [33]. A pixel is classified as cloud if:
Rrc,865 > 0.1, or
Rrc,865 > 0.06 and R(Rrc,745, Rrc,865) < 1.15, or
Rrc,865 > 0.027 and R(Rrc,745, Rrc,865) < 1.15 and S(Rrc,745, Rrc,865) > 0.01.
where R(Rrc,745, Rrc,865) = Rrc,745/Rrc,865, and S(Rrc,745, Rrc,865) = Rrc,745Rrc,865. Pixels around a cloud pixel marked by this method are also flagged as cloud to minimize the influence caused by cloud shadows.

2.3. Field Data

The in situ Rrs data were derived from two distinct sources. Firstly, a stationary oceanographic platform named Dongou, situated at 27.675°N and 121.358°E, approximately 80 km southeast of Wenzhou, China, served as a reliable source for collecting long-term radiometric measurements. These measurements encompassed remote sensing reflectance (Rrs) and normalized water-leaving radiance (Lwn). The shallow water depth around the platform minimized the impact of bottom effects on Lwn measurements. Since 2017, the platform has been equipped with a Sea-Viewing Wide Field-of-View
Sensor Photometer Revision for Incident Surface Measurements (SeaPRISM, CE318H1590TV12, Cimel Electronique, Paris, Frence) autonomous radiometer system (depicted in Figure 2), which collects sea-viewing data every 30 min, within a four-hour window centered at 12:00 PM local time. This system is capable of performing ocean color measurements across 11 wavelengths ranging from 400 to 1020 nm. The SeaPRISM data used in this study, collected between April 2021 and October 2022, were manually inspected to ensure data integrity and the absence of corrupted spectra. For the purpose of match-up comparison analysis, Lwn and Rrs data were selected at specific center wavelengths: 412, 442, 490, 510, 560, 620, 667, 779, and 865 nm. To leverage the high-frequency measurements provided by SeaPRISM, the in situ data used for the quantitative match-up comparison analysis were carefully chosen, ensuring they corresponded to the measurements made within a narrow (±0.5 h) time window surrounding the satellite overpass time at the Dongou site [60]. This approach minimized potential biases introduced by the rapid dynamic changes observed in coastal waters.
The second source of Rrs data originated from the underway observations conducted during the LORCE cruise in the ECS in August 2021. For this cruise, a shipboard fully automated hyperspectral radiometric measurement system named CrusieAOP (CAOP-H2301, Anhua Ocean Intelligent Equipment Ltd., Guangzhou, China, depicted in Figure 2d) was utilized. This system simultaneously measured the total radiance leaving the sea surface, sky radiance, and surface incident irradiance across wavelengths ranging from 350 to 900 nm with a 1 nm interval. Consequently, the in situ Rrs values were derived from these three radiometric parameters, adhering to the NASA Ocean Optics Protocols.
The records of algal blooms, encompassing the occurring time, location, and causative species, were retrieved from the WMC. These records were collected during field measurements conducted on a series of cruise in the ECS, covering the period from 2021 to 2023. At each visited bloom station during these cruises, we collected 1-L or 500-mL water samples from the surface (at a depth of 2–3 m) using Niskin bottles and preserved them in 2% glutaraldehyde [61]. Additionally, surface water samples were collected in 500 mL bottles and stored with 2% formalin. Later, the phytoplankton taxa were identified and enumerated through sedimentation on a scaled slide, employing a light microscope like the Leica DM2500. For each sample, a minimum of 300 units were counted, adhering to the morphological classification criteria [61,62]. The detailed species identified as the dominant algal species in the bloom are listed in Table 1.

2.4. Baseline Subtraction Algorithm

Rrc-based bloom detection algorithms typically employ a baseline subtraction method to discern spectral signatures exhibited by blooms across various wavebands. This technique has also found widespread application in diverse ocean color analyses. [31,32,38]. The popularity of the algorithms stems from the their lower sensitivity, compared to band-ratio approaches widely used in algorithms based on remote-sensing reflectance, towards various errors caused by instrument noise and imperfect atmospheric correction, including sun glint and whitecap corrections [33]. More importantly, the band ratio of Rrc cannot be directly compared between images captured at different time and under different atmosphere conditions due to the large variation in the aerosol scattering contributions.
A series of spectral indices have been developed using the baseline subtraction algorithm. These indices are generally used to quantify the height of reflectance at the specific waveband (λ) in comparison to a baseline that is linearly established between two neighboring wavebands (λ+ and λ) as follows:
S S ( λ ) = R ( λ ) R ( λ ) { R ( λ + ) R ( λ ) } × λ λ λ + λ
where R was either satellite-derived Rrc(λ) (dimensionless) or π∗Rrs(λ). In this study, five spectral indices—SS490, CI, DI, FLH, and MCI—are applied for algal bloom detection. The specification of λ, λ, and λ+ for the five indices are listed in Table 2.

2.5. Accuracy Assessment

In this paper, we use the confusion matrix to calculate the algal bloom detection accuracy of different BLIs. The measured stations are categorized into algal bloom (b) stations and non-bloom stations (nb) using Chinese Marine Monitoring Specification [63]. Similarly, the remote sensing detection results can be categorized into algal bloom (B) and non-bloom areas (NB), and scenarios will appear after matching ((A) b-B, (B) b-NB, (C) nb-B, (D) nb-NB).
The overall performance of the algorithm was assessed using the F-measure coefficient (FM), which represents the harmonic mean between precision and sensitivity, and was computed as follows [42]:
F M = ( β 2 + 1 ) × p r e c i s i o n × s e n s i t i v i t y β 2 × p r e c i s i o n + s e n s i t i v i t y p r e c i s i o n = A / ( A   +   C ) s e n s i t i v i t y = A / ( A   +   B )
when β is set to 1, precision and sensitivity are weighted equally. However, for this study, a value of β = 0.5 was chosen to prioritize precision over sensitivity.

3. Results and Discussion

3.1. Spectral Characteristics of Rrc

Figure 3 illustrates typical Rrc spectra of clear, medium turbid, turbid and algal bloom waters, as well as the normalized Rrc spectra using the maximum and minimum values of spectral reflectance. The spectral characteristics of algal bloom water differ from various water types. The clear water spectra show peak reflectance at blue bands, while those of algal bloom indicate a negative spectral shape around 490 nm (positive SS490). Although both are medium turbid and algal bloom water peak at 555 nm (positive CI), the bloom water showed a relatively lower reflectance and peaks up at 680 nm (positive FLH) or 709 nm bands (positive MCI) owing to the integrated effect of chlorophyll fluorescence and absorption at red band. Turbid water Rrc spectra exhibit significantly high values compared to other water types, and their reflectance peaks also shift to the red band around 620 nm (DI). These results suggest that Rrc and its baseline indices can effectively characterize the spectral differences between algal bloom and non-bloom water.
To demonstrate the stability of the Rrc BLIs under different aerosol conditions, we calculated the baseline indices using the in situ measured Rrs from SeaPRISM on the Dongou platform, and compared them with those calculated from GOCI-II Rrc. On top of that, we conducted more direct comparisons between GOCI-II Rrc and Rrs indices. Specifically, CI derived from in situ Rrs and CI derived from GOCI-II Rrc showed strong correlations (R2 > 0.91, N = 204), and correlations for SS490 and DI are still high (Figure 4a–c). Most importantly, the indices derived from GOCI-II Rrs and Rrc appear to be stable and consistent (Figure 4d–f). Since the SeaPRISM does not contain the 680 and 709 nm bands, we also used the hyperspectral Rrs measurement from the shipboard CrusieAOP system to calculate the baseline indices of FLH and MCI to evaluate the synchronous GOCI-II Rrc results, along with the other three indices. Figure 5 depicted a section of underway Rrs measurements that traverse the coastal area, encompassing a range of water bodies from turbid to clean, including turbid, moderately turbid, bloom, and clean water. To minimize the uncertainties caused by different observation times of GOCI-II and shipboard measurements, we used three GOCI-II image data from the different time slots on the same day to match the different shipboard trajectories according to the measurement time of the in situ Rrs. The GOCI-II calculated baseline indices show good agreement with those from the underway Rrs measurements. Notably, SS488 and MCI appear to perform better on identifying algal blooms water, and these two indices exhibit significant increases in the algal bloom water and can be distinguished from other water types.
Additionally, we further evaluate the time series BLIs (including FLH and MCI) from both GOCI-II Rrc and Rrs at the Dongou platform in 2021 (Figure 6). The results of all indices from Rrc and Rrs have shown good consistency. More importantly, the effective data volume of Rrc is approximately 20% much greater than that of Rrs. It can be observed that the points representing only Rrc data mainly occur in January and February. During this season, sediment-dominated waters disturbed near the Dongou Platform, in which the atmospheric aerosol correction was usually poor and thus Rrs data could not be successfully retrieved. Similarly, during algal blooms in May, bloom water with extremely high chlorophyll concentrations also achieves good atmospheric aerosol correction result. Nevertheless, despite these scenarios, the strong stability of Rrc indices is effective in characterizing the spectral features of various optical water types.

3.2. Comparison of the Effectiveness of Different BLIs for Algal Bloom Detection

To compare the algal bloom detection effects of different BLIs, GOCI-II remote sensing images of several typical algal bloom and non-bloom events were selected and matched with the in situ measurements as shown in Figure 7. Each column shows, from left to right, the RGB true-color composite map and the distribution of the corresponding BLIs (SS490, CI, DI, FLH, MCI). The yellow circles indicate the measured bloom stations, with the corresponding algal cell abundance values labeled nearby (the reference concentration of algal bloom is set as 5 × 105 cell/L [63]). There were significant differences in the validation results of different BLIs. The first and second rows showed the remote sensing images of 30 April 2021 and 1 May 2023, both depicting algal blooms caused by P. donghaiense and occurring in similar periods of the year. Among the five BLIs, SS490 showed the best validation for the P. donhaiense blooms, with the measured bloom stations corresponding to areas of high SS490 values. In contrast, CI had errors in indicating algal bloom area, with bloom stations corresponding to the boundary between the high and low-value areas. The stability of DI was poor, with matches in 2021 and 2023 completely reversed, as the measured bloom stations in 2021 fell in the high-value area of the DI, while those in 2023 were situated in the low-value area. The validation effect of FLH was also unstable. Although the measured points fell in areas of relatively high value in both cases, the differentiation between bloom water and non-bloom water was obviously reduced (Figure 7k). The validation effect of MCI was much better than that of CI, DI, and FLH, as the measured bloom points were located in the area of high value.
For diatom bloom (Figure 7m–r), SS490 and MCI also performed well, with both measured bloom sites located in the high-value area, while DI, CI, and FLH correspond to the boundary between low and high value areas or low-value area. The matching of DI, CI, and FLH with the remote sensing imagery of A. sanguinea bloom was still poor. MCI was slightly better, but the southern measured points were matched to the low-value area. Only SS490 had the best validation performance, with all measured bloom stations located in the high- value area. Additionally, based on the winter typical non-bloom images on 22 December 2022, it can be observed that, when blooms did not occur, either type of BLI may also show high values in turbid waters. This result suggests that any single index cannot accurately identify algal blooms, and further analysis for the differences between turbid and bloom waters is needed.

3.3. Algorithm Development of HABs Detection

Affected by turbid waters, no single BLI can accurately identify algal blooms. Therefore, we constructed a Turbidity Index TI (TI = Rrc,660Rrc,745) using the spectral properties in the red to near-infrared bands to indicate water turbidity. Figure 2 illustrates the scatter distribution between different BLIs and TI for the boxed area in Figure 7. When plotted with TI, each BLI showed certain distribution characteristics.
For SS490, the scattering points in the region of the two P. donhaiense blooms (Figure 8a,d) always present a shape resembling the letter “S”, corresponding to the mixing of clean water, algal bloom water, medium turbid water and turbid water in the region. Different water types were labeled in Figure 8 with blue, red, green, and yellow circles, respectively. Clean waters close to the open sea had low values of SS490 and TI, which gradually increased during mixing with algal bloom water. Then, bloom water mixed with medium turbid water with increasing TI but decreasing SS490. Finally, SS490 and TI increased together again during the transition from medium turbid to turbid waters. As a result, the scatterplot shows a clear “S” shape, indicating the mixing trends.
A similar phenomenon was observed in the scatterplots of DI, FLH, and MCI with TI. However, there were some changes in the shape compared to SS490. These three BLIs were much higher in the turbid waters, resulting in the elongation of the upper part of the “S”, reducing the ability to distinguish between clear water and algal bloom water. Also, this mixing demonstrated by the scatter of DI, FLH and MCI was not stable. In Figure 8h–j, the scatter showed almost consistent mixing trends, making it difficult to distinguish algal bloom waters from the other types of waters. The mixing of medium turbid and high turbid waters was less pronounced in the area surrounding the diatom bloom event (Figure 8k–o), where there was a mixing of algal bloom water and high turbid water. Fortunately, all BLIs but CI were able to differentiate the algal bloom scatters clearly. In the scatterplot of the A. sanguinea bloom (Figure 8p–t), the difference was more pronounced. SS490 scatter displayed a distinct “S” shape, demonstrating excellent capability in distinguishing the algal bloom water, yet the scatters of DI, FLH, and MCI showed very consistent scatter variation characteristics as in the second row. Additionally, a typical non-bloom dataset from 22 December 2022 was selected and compared with the scattering points during algal bloom events. When there was no algal bloom, the mixing occurred only between clean, medium turbid, and turbid waters, so BLIs increased as TI increased and showed an upward trend and this trend was similar in the scatter plots of all five BLIs but CI. Overall, among the four algal blooms, SS490 had the best stability and responsiveness, and basically identified the algal bloom points in each bloom event, while DI, FLH, and MCI identified only one P. donghaiense bloom and one diatom bloom, and CI was almost incapable of detecting algal blooms.
Based on the comparisons conducted above, an algorithm for detecting algal blooms was developed, focusing on the correction of turbid water effects (Figure 9a). To demonstrate the efficacy of each step of the algorithm, we take synthetical SS490 as an example, during a P. donghaiense bloom on 31 May 2023 (Figure 9b–f).
First, the distribution of SS490, calculated directly from Rrc data obtained from GOCI-II, is shown in Figure 9c. Subsequently, cloud pixels are identified based on the spectral features of the Rrc data (as described in Section 2.2), resulting in the distribution of the SS490 after cloud masking, as shown in Figure 9d. Meanwhile, it was also observed that, while SS490 captures different algal blooms, it also exhibits similar characteristics in turbid water, so it is imperative to differentiate between various water types. In this case, we empirically set the TI determination threshold for turbid water bodies at 0.02. Since algal blooms usually do not occur in turbid water in the sediment-rich ECS, this study does not conduct further processing. Water with TI values between 0.012 and 0.02, pointing out the medium turbid water or the mixed turbid and algal bloom water, requires identification from true algal bloom area due to the similar high SS490 with bloom area. To address this issue, the cubic deflation of SS490 was carried out for this type of water, so that the reduced SS490 values will be compared to those in the algal bloom area, shown as in Figure 9e. For non-turbid water, SS490 effectively distinguishes algal bloom water from clean water, obviating the need for further processing. Lastly, a synthetical SS490 is obtained by combining SS490 values for non-turbid and corrected SS490 values for turbid waters. A threshold of 0.002, determined from field measurements, is applied to detect areas above this threshold as algal bloom areas (Figure 9f).

3.4. Validation during the 2021–2023 Bloom Events

In order to compare the algal bloom detection accuracy of different BLIs, we collected field-measured data on dominant algal species and the corresponding cell abundance during algal blooms from 2021 to 2023. A total of 75 pairs of data were collected, including the sampling time, latitude, longitude, dominant algal species, and algal abundance. According to the Chinese National Ocean Monitoring Code [63], stations with an algal abundance exceeding 5 × 105 cells/L were determined as algal bloom stations. Furthermore, the field-measured data were matched with the GOCI-II within a 2 km × 2 km area around the sampling points on the same day [64]. After excluding cloudy or partially missing Rrc data, a total of 35 data pairs were successfully matched. In addition to SS490, the other four BLIs were also corrected for turbid water to minimize their influences. The scatter plots of different BLIs against measured algal density are shown in Figure 10, with green dots representing non-algal bloom points and red dots representing algal bloom points recognized by the measured data. Thresholds were also set for each BLI based on previous research and experience, with the points above the threshold determined to be algal bloom points. Clearly, the synthetical SS490 was the best for agal bloom detection, showing a distinct difference between algal bloom and non-bloom scatters. By calculating confusion matrices, the quantitative accuracy of algal bloom detection by different BLIs was obtained, as detailed in Table 3.
According to the results demonstrated in Table 3, SS490 presented the best detection performance, with an overall F-measure coefficient (FM) value of 0.97, surpassing other indices in both algal bloom and non-algal bloom point discrimination accuracy. MCI ranked the second with an FM value of 0.88, indicating relatively high accuracy. However, MCI required data from the 709 nm band, which was missing in some commonly used sensors such as SNPP/VIIRS, limiting its application to some extent. In comparison, DI, CI, and FLH showed lower accuracy in bloom detections. Specifically, DI had a lower accuracy in recognizing non-algal bloom points, limiting the overall accuracy obtained from the confusion matrix calculation. FLH exhibited lower accuracy in identifying algal bloom points, resulting in insufficient sensitivity and reducing the overall FM value.
We also analyzed the results of Rrs of GOCI-II to detect algal bloom using several algorithms based on Rrs, including LHR [46], RAB [47], RI [43], and SS510 [65]. Scatter plots of algal cell abundance and baseline indices are shown in Figure 11, where green dots represent non-algal bloom points, red dots represent algal bloom points, and black dots represent points where Rrc data were valid, but Rrs data were missing. Both Rrc and Rrs data show good potential in terms of detecting algal blooms, but there were signficantly less GOCI-II Rrs data than there were Rrc data. Specifically, the number of data pairs successfully matched with measured data for Rrc data was 35, while the number of data pairs successfully matched with spectral data for Rrs was 20, resulting in a missing rate of 42.86%. value.
Meanwhile, we found that the values of these indices did not show a significant difference between non-bloom points (green) and algal bloom points (red) (Figure 11). Therefore, we marked the positions and corresponding measured algal cell abundance of these three points on RGB images, and plotted their Rrc and pi∗Rrs spectra (as shown in Figure 12, in which Rrs is multiplied by pi to have the same measure as Rrc). Clearly, these three points were all located in turbid waters, lacking algal bloom spectral characteristics, indicating that these points were likely influenced by turbid water and cannot be distinguished as non-algal bloom points. The algorithm proposed for calculating BLIs based on Rrc spectral data includes a turbid correction step, which can effectively eliminate the influence of turbid water on detecting algal blooms. Overall, the accuracy of algal bloom detection using different BLIs calculated based on Rrc spectra varies greatly, with synthetical SS490 achieving the highest detection accuracy and demonstrating good versatility and practicality. Additionally, the accuracy of Rrc-derived BLIs was not weaker than that of Rrs, but there was a significant increase in the amount of valid Rrc data compared to Rrs.

4. Conclusions

In this paper, we developed a synthetical SS490 algorithm to identify algal blooms in the ECS using the Rrc data of GOCI-II and validated the feasibility of Rrc for algal bloom detection. The Rrc-derived BLIs can reflect the water condition well and show a strong agreement (R2 > 0.98) with the Rrs-derived BLIs. Among the listed BLIs, SS490 is optimal for algal bloom detection in the measured bloom events, followed by MCI. However, a single index for algal bloom detection is susceptible to the interference from turbid waters. Thus, we developed an algal bloom detection algorithm based on the synthetical Rrc derived BLI, incorporating a correction for turbid water. The synthetical SS490 performs optimally in the matching validation with the measured algal abundance data (FM = 0.97), and can stably recognize regardless of whether the blooms are P. donghaiense, diatom, or A. sanguinea bloom. We also presented the results of GOCI-II-derived Rrs for algal bloom detection and found that Rrc-derived BLIs were not weaker than the algorithms of Rrs for algal bloom detection, with an obvious increase in valid data for the Rrc-derived BLI. In conclusion, algal bloom detection using Rrc data is feasible, and the synthetical SS490 algorithm analyzed in this study can be widely applied.

Author Contributions

Conceptualization, C.Z. and C.L.; methodology, C.Z. and Y.Z.; validation, C.L.; investigation, C.Z. and L.L.; data curation, C.Z. and L.A.; writing—original draft preparation, C.Z. and Y.L.; writing—review and editing, B.T. and C.L.; visualization, L.L. and H.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key Research and Development Program of China (Grant 2023YFC3107604); National Natural Science Foundation of China (grant 42276200); Scientific Research Fund of the Second Institute of Oceanography, MNR (grant SZ2332). the Project of State Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, MNR (Grant SOEDZZ2103).

Data Availability Statement

GOCI-II data were downloaded from https://oceancolor.nasa.gs-fc.gov (accessed on 30 May 2023).

Acknowledgments

We would like to thank Wenzhou Marine Environment Monitoring Center station for providing algal species and cell abundance data of HAB events during the study period and long time series radiometric data from the Dongou Ocean Optical Platform, as well as the Second Institute of Oceanography, Ministry of Natural Resources, LORCE for providing algal species and cell abundance data, Korea Hydrographic and Oceanographic Agency for providing GOCI-II data products. Thank you to the co-authors for their help in this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Anderson, D.M.; Cembella, A.D.; Hallegraeff, G.M. Progress in Understanding Harmful Algal Blooms: Paradigm Shifts and New Technologies for Research, Monitoring, and Management. Annu. Rev. Mar. Sci. 2012, 4, 143–176. [Google Scholar] [CrossRef] [PubMed]
  2. Hallegraeff, G.M.; Anderson, D.M.; Belin, C.; Bottein, M.-Y.D.; Bresnan, E.; Chinain, M.; Enevoldsen, H.; Iwataki, M.; Karlson, B.; McKenzie, C.H.; et al. Perceived Global Increase in Algal Blooms Is Attributable to Intensified Monitoring and Emerging Bloom Impacts. Commun. Earth Environ. 2021, 2, 117. [Google Scholar] [CrossRef] [PubMed]
  3. IOCCG. Observation of Harmful Algal Blooms with Ocean Colour Radiometry; IOCCG Report Series; IOCCG: Dartmouth, NS, Canada, 2021. [Google Scholar] [CrossRef]
  4. Cannizzaro, J.P.; Carder, K.L.; Chen, F.R.; Heil, C.A.; Vargo, G.A. A Novel Technique for Detection of the Toxic Dinoflagellate, Karenia Brevis, in the Gulf of Mexico from Remotely Sensed Ocean Color Data. Cont. Shelf Res. 2008, 28, 137–158. [Google Scholar] [CrossRef]
  5. Carvalho, G.A.; Minnett, P.J.; Fleming, L.E.; Banzon, V.F.; Baringer, W. Satellite Remote Sensing of Harmful Algal Blooms: A New Multi-Algorithm Method for Detecting the Florida Red Tide (Karenia brevis). Harmful Algae 2010, 9, 440–448. [Google Scholar] [CrossRef] [PubMed]
  6. Schofield, O.; Grzymski, J.; Bissett, W.P.; Kirkpatrick, G.J.; Millie, D.F.; Moline, M.; Roesler, C.S. Optical Monitoring and Forecasting Systems for Harmful Algal Blooms: Possibility or Pure Dream? J. Phycol. 1999, 35, 1477–1496. [Google Scholar] [CrossRef]
  7. Stumpf, R.P.; Holderied, K.; Sinclair, M. Determination of Water Depth with High-resolution Satellite Imagery over Variable Bottom Types. Limnol. Oceanogr. 2003, 48, 547–556. [Google Scholar] [CrossRef]
  8. Tomlinson, M.C.; Stumpf, R.P.; Ransibrahmanakul, V.; Truby, E.W.; Kirkpatrick, G.J.; Pederson, B.A.; Vargo, G.A.; Heil, C.A. Evaluation of the Use of SeaWiFS Imagery for Detecting Karenia Brevis Harmful Algal Blooms in the Eastern Gulf of Mexico. Remote Sens. Environ. 2004, 91, 293–303. [Google Scholar] [CrossRef]
  9. Blondeau-Patissier, D.; Gower, J.F.R.; Dekker, A.G.; Phinn, S.R.; Brando, V.E. A Review of Ocean Color Remote Sensing Methods and Statistical Techniques for the Detection, Mapping and Analysis of Phytoplankton Blooms in Coastal and Open Oceans. Prog. Oceanogr. 2014, 123, 123–144. [Google Scholar] [CrossRef]
  10. Mouw, C.B.; Hardman-Mountford, N.J.; Alvain, S.; Bracher, A.; Brewin, R.J.W.; Bricaud, A.; Ciotti, A.M.; Devred, E.; Fujiwara, A.; Hirata, T.; et al. A Consumer’s Guide to Satellite Remote Sensing of Multiple Phytoplankton Groups in the Global Ocean. Front. Mar. Sci. 2017, 4, 41. [Google Scholar] [CrossRef]
  11. Dall’Olmo, G.; Gitelson, A.A. Effect of Bio-Optical Parameter Variability on the Remote Estimation of Chlorophyll-a Concentration in Turbid Productive Waters: Experimental Results—Erratum. Appl. Opt. 2005, 44, 3342. [Google Scholar] [CrossRef]
  12. Gitelson, A.A.; Dall’Olmo, G.; Moses, W.; Rundquist, D.C.; Barrow, T.; Fisher, T.R.; Gurlin, D.; Holz, J. A Simple Semi-Analytical Model for Remote Estimation of Chlorophyll-a in Turbid Waters: Validation. Remote Sens. Environ. 2008, 112, 3582–3593. [Google Scholar] [CrossRef]
  13. Hu, C.; Mullerkarger, F.; Taylor, C.; Carder, K.; Kelble, C.; Johns, E.; Heil, C. Red Tide Detection and Tracing Using MODIS Fluorescence Data: A Regional Example in SW Florida Coastal Waters. Remote Sens. Environ. 2005, 97, 311–321. [Google Scholar] [CrossRef]
  14. Ahn, Y.-H.; Shanmugam, P. Detecting the Red Tide Algal Blooms from Satellite Ocean Color Observations in Optically Complex Northeast-Asia Coastal Waters. Remote Sens. Environ. 2006, 103, 419–437. [Google Scholar] [CrossRef]
  15. Sathyendranath, S.; Watts, L.; Devred, E.; Platt, T.; Caverhill, C.; Maass, H. Discrimination of Diatoms from Other Phytoplankton Using Ocean-Colour Data. Mar. Ecol. Prog. Ser. 2004, 272, 59–68. [Google Scholar] [CrossRef]
  16. Westberry, T.K.; Siegel, D.A.; Subramaniam, A. An Improved Bio-optical Model for the Remote Sensing of Trichodesmium Spp. Blooms. J. Geophys. Res. 2005, 110, 2004JC002517. [Google Scholar] [CrossRef]
  17. Ishizaka, J.; Kitaura, Y.; Touke, Y.; Sasaki, H.; Tanaka, A.; Murakami, H.; Suzuki, T.; Matsuoka, K.; Nakata, H. Satellite Detection of Red Tide in Ariake Sound, 1998–2001. J. Oceanogr. 2006, 62, 37–45. [Google Scholar] [CrossRef]
  18. Lubac, B.; Loisel, H.; Guiselin, N.; Astoreca, R.; Felipe Artigas, L.; Mériaux, X. Hyperspectral and Multispectral Ocean Color Inversions to Detect Phaeocystis globosa Blooms in Coastal Waters. J. Geophys. Res. 2008, 113, 2007JC004451. [Google Scholar] [CrossRef]
  19. Amin, S.A.; Green, D.H.; Hart, M.C.; Küpper, F.C.; Sunda, W.G.; Carrano, C.J. Photolysis of Iron–Siderophore Chelates Promotes Bacterial–Algal Mutualism. Proc. Natl. Acad. Sci. USA 2009, 106, 17071–17076. [Google Scholar] [CrossRef]
  20. Siswanto, E.; Ishizaka, J.; Tripathy, S.C.; Miyamura, K. Detection of Harmful Algal Blooms of Karenia Mikimotoi Using MODIS Measurements: A Case Study of Seto-Inland Sea, Japan. Remote Sens. Environ. 2013, 129, 185–196. [Google Scholar] [CrossRef]
  21. Kurekin, A.A.; Miller, P.I.; Van Der Woerd, H.J. Satellite Discrimination of Karenia Mikimotoi and Phaeocystis Harmful Algal Blooms in European Coastal Waters: Merged Classification of Ocean Colour Data. Harmful Algae 2014, 31, 163–176. [Google Scholar] [CrossRef]
  22. Kim, Y.; Yoo, S.; Son, Y.B. Optical Discrimination of Harmful Cochlodinium Polykrikoides Blooms in Korean Coastal Waters. Opt. Express 2016, 24, A1471. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, C.; Ishizaka, J.; Saitoh, K.; Mine, T.; Yamashita, H. A Novel Method Based on Backscattering for Discriminating Summer Blooms of the Raphidophyte (Chattonella spp.) and the Diatom (Skeletonema spp.) Using MODIS Images in Ariake Sea, Japan. Remote Sens. 2020, 12, 1504. [Google Scholar] [CrossRef]
  24. Kim, Y.-K.; Kim, B.; Kim, V.N. Re-Evaluation of the Roles of DROSHA, Exportin 5, and DICER in microRNA Biogenesis. Proc. Natl. Acad. Sci. USA 2016, 113, E1881–E1889. [Google Scholar] [CrossRef] [PubMed]
  25. Gordon, H.R. Atmospheric Correction of Ocean Color Imagery in the Earth Observing System Era. J. Geophys. Res. 1997, 102, 17081–17106. [Google Scholar] [CrossRef]
  26. Gordon, H.R.; Wang, M. Retrieval of Water-Leaving Radiance and Aerosol Optical Thickness over the Oceans with SeaWiFS: A Preliminary Algorithm. Appl. Opt. 1994, 33, 443. [Google Scholar] [CrossRef] [PubMed]
  27. Hu, C.; Carder, K.L.; Muller-Karger, F.E. Atmospheric Correction of SeaWiFS Imagery: Assessment of the Use of Alternative Bands. Appl. Opt. 2000, 39, 3573. [Google Scholar] [CrossRef] [PubMed]
  28. Ruddick, K.G.; Ovidio, F.; Rijkeboer, M. Atmospheric Correction of SeaWiFS Imagery for Turbid Coastal and Inland Waters. Appl. Opt. 2000, 39, 897. [Google Scholar] [CrossRef] [PubMed]
  29. King, M.D.; Platnick, S.; Menzel, W.P.; Ackerman, S.A.; Hubanks, P.A. Spatial and Temporal Distribution of Clouds Observed by MODIS Onboard the Terra and Aqua Satellites. IEEE Trans. Geosci. Remote Sens. 2013, 51, 3826–3852. [Google Scholar] [CrossRef]
  30. Hu, C.; Feng, L. Modified MODIS Fluorescence Line Height Data Product to Improve Image Interpretation for Red Tide Monitoring in the Eastern Gulf of Mexico. J. Appl. Remote Sens. 2016, 11, 012003. [Google Scholar] [CrossRef]
  31. Gower, J.; King, S.; Borstad, G.; Brown, L. Detection of Intense Plankton Blooms Using the 709 Nm Band of the MERIS Imaging Spectrometer. Int. J. Remote Sens. 2005, 26, 2005–2012. [Google Scholar] [CrossRef]
  32. Hu, C.; Lee, Z.; Ma, R.; Yu, K.; Li, D.; Shang, S. Moderate Resolution Imaging Spectroradiometer (MODIS) Observations of Cyanobacteria Blooms in Taihu Lake, China. J. Geophys. Res. 2010, 115, 2009JC005511. [Google Scholar] [CrossRef]
  33. Hu, C. An Empirical Approach to Derive MODIS Ocean Color Patterns under Severe Sun Glint. Geophys. Res. Lett. 2011, 38, L01603. [Google Scholar] [CrossRef]
  34. Duan, H.; Ma, R.; Zhang, Y.; Loiselle, S.A. Are Algal Blooms Occurring Later in Lake Taihu? Climate Local Effects Outcompete Mitigation Prevention. J. Plankton Res. 2014, 36, 866–871. [Google Scholar] [CrossRef]
  35. Feng, L.; Hou, X.; Li, J.; Zheng, Y. Exploring the Potential of Rayleigh-Corrected Reflectance in Coastal and Inland Water Applications: A Simple Aerosol Correction Method and Its Merits. ISPRS J. Photogramm. Remote Sens. 2018, 146, 52–64. [Google Scholar] [CrossRef]
  36. Feng, L.; Hu, C.; Han, X.; Chen, X.; Qi, L. Long-Term Distribution Patterns of Chlorophyll-a Concentration in China’s Largest Freshwater Lake: MERIS Full-Resolution Observations with a Practical Approach. Remote Sens. 2014, 7, 275–299. [Google Scholar] [CrossRef]
  37. Chen, S.; Hu, C.; Barnes, B.B.; Xie, Y.; Lin, G.; Qiu, Z. Improving Ocean Color Data Coverage through Machine Learning. Remote Sens. Environ. 2019, 222, 286–302. [Google Scholar] [CrossRef]
  38. Matthews, M.W.; Bernard, S.; Robertson, L. An Algorithm for Detecting Trophic Status (Chlorophyll-a), Cyanobacterial-Dominance, Surface Scums and Floating Vegetation in Inland and Coastal Waters. Remote Sens. Environ. 2012, 124, 637–652. [Google Scholar] [CrossRef]
  39. Matthews, M.W.; Odermatt, D. Improved Algorithm for Routine Monitoring of Cyanobacteria and Eutrophication in Inland and Near-Coastal Waters. Remote Sens. Environ. 2015, 156, 374–382. [Google Scholar] [CrossRef]
  40. Hu, C.; Li, D.; Chen, C.; Ge, J.; Muller-Karger, F.E.; Liu, J.; Yu, F.; He, M. On the Recurrent Ulva Prolifera Blooms in the Yellow Sea and East China Sea. J. Geophys. Res. 2010, 115, 2009JC005561. [Google Scholar] [CrossRef]
  41. Qi, L.; Hu, C.; Cannizzaro, J.; Corcoran, A.A.; English, D.; Le, C. VIIRS Observations of a Karenia Brevis Bloom in the Northeastern Gulf of Mexico in the Absence of a Fluorescence Band. IEEE Geosci. Remote Sens. Lett. 2015, 12, 2213–2217. [Google Scholar] [CrossRef]
  42. Cannizzaro, J.P.; Barnes, B.B.; Hu, C.; Corcoran, A.A.; Hubbard, K.A.; Muhlbach, E.; Sharp, W.C.; Brand, L.E.; Kelble, C.R. Remote Detection of Cyanobacteria Blooms in an Optically Shallow Subtropical Lagoonal Estuary Using MODIS Data. Remote Sens. Environ. 2019, 231, 111227. [Google Scholar] [CrossRef]
  43. Lou, X.; Hu, C. Diurnal Changes of a Harmful Algal Bloom in the East China Sea: Observations from GOCI. Remote Sens. Environ. 2014, 140, 562–572. [Google Scholar] [CrossRef]
  44. Shang, S.; Wu, J.; Huang, B.; Lin, G.; Lee, Z.; Liu, J.; Shang, S. A New Approach to Discriminate Dinoflagellate from Diatom Blooms from Space in the East China Sea. J. Geophys. Res. Ocean. 2014, 119, 4653–4668. [Google Scholar] [CrossRef]
  45. Tang, D.; Kawamura, H.; Oh, I.S.; Baker, J. Satellite Evidence of Harmful Algal Blooms and Related Oceanographic Features in the Bohai Sea during Autumn 1998. Adv. Space Res. 2006, 37, 681–689. [Google Scholar] [CrossRef]
  46. Tao, B.; Mao, Z.; Lei, H.; Pan, D.; Bai, Y.; Zhu, Q.; Zhang, Z. A Semianalytical MERIS Green-red Band Algorithm for Identifying Phytoplankton Bloom Types in the E Ast C Hina S Ea. J. Geophys. Res. Ocean. 2017, 122, 1772–1788. [Google Scholar] [CrossRef]
  47. Tao, B.; Mao, Z.; Lei, H.; Pan, D.; Shen, Y.; Bai, Y.; Zhu, Q.; Li, Z. A Novel Method for Discriminating Prorocentrum Donghaiense from Diatom Blooms in the East China Sea Using MODIS Measurements. Remote Sens. Environ. 2015, 158, 267–280. [Google Scholar] [CrossRef]
  48. Hou, X.; Feng, L.; Dai, Y.; Hu, C.; Gibson, L.; Tang, J.; Lee, Z.; Wang, Y.; Cai, X.; Liu, J.; et al. Global Mapping Reveals Increase in Lacustrine Algal Blooms over the Past Decade. Nat. Geosci. 2022, 15, 130–134. [Google Scholar] [CrossRef]
  49. Dai, Y.; Yang, S.; Zhao, D.; Hu, C.; Xu, W.; Anderson, D.M.; Li, Y.; Song, X.-P.; Boyce, D.G.; Gibson, L.; et al. Coastal Phytoplankton Blooms Expand and Intensify in the 21st Century. Nature 2023, 615, 280–284. [Google Scholar] [CrossRef] [PubMed]
  50. Feng, L.; Dai, Y.; Hou, X.; Xu, Y.; Liu, J.; Zheng, C. Concerns about Phytoplankton Bloom Trends in Global Lakes. Nature 2021, 590, E35–E47. [Google Scholar] [CrossRef]
  51. Zhou, M.; Shen, Z.; Yu, R. Responses of a Coastal Phytoplankton Community to Increased Nutrient Input from the Changjiang (Yangtze) River. Cont. Shelf Res. 2008, 28, 1483–1489. [Google Scholar] [CrossRef]
  52. Chen, C.-T.A. Chemical and Physical Fronts in the Bohai, Yellow and East China Seas. J. Mar. Syst. 2009, 78, 394–410. [Google Scholar] [CrossRef]
  53. Liu, X.; Xiao, W.; Landry, M.R.; Chiang, K.-P.; Wang, L.; Huang, B. Responses of Phytoplankton Communities to Environmental Variability in the East China Sea. Ecosystems 2016, 19, 832–849. [Google Scholar] [CrossRef]
  54. Qi, L.; Hu, C.; Xing, Q.; Shang, S. Long-Term Trend of Ulva Prolifera Blooms in the Western Yellow Sea. Harmful Algae 2016, 58, 35–44. [Google Scholar] [CrossRef] [PubMed]
  55. Qi, L.; Hu, C.; Wang, M.; Shang, S.; Wilson, C. Floating Algae Blooms in the East China Sea. Geophys. Res. Lett. 2017, 44, 11501–11509. [Google Scholar] [CrossRef]
  56. Wynne, T.T.; Stumpf, R.P.; Briggs, T.O. Comparing MODIS and MERIS Spectral Shapes for Cyanobacterial Bloom Detection. Int. J. Remote Sens. 2013, 34, 6668–6678. [Google Scholar] [CrossRef]
  57. Shen, F.; Tang, R.; Sun, X.; Liu, D. Simple Methods for Satellite Identification of Algal Blooms and Species Using 10-Year Time Series Data from the East China Sea. Remote Sens. Environ. 2019, 235, 111484. [Google Scholar] [CrossRef]
  58. Ahn, J.-H.; Park, Y.-J.; Kim, W.; Lee, B. Simple Aerosol Correction Technique Based on the Spectral Relationships of the Aerosol Multiple-Scattering Reflectances for Atmospheric Correction over the Oceans. Opt. Express 2016, 24, 29659. [Google Scholar] [CrossRef] [PubMed]
  59. Ahn, J.-H.; Park, Y.-J. Estimating Water Reflectance at Near-Infrared Wavelengths for Turbid Water Atmospheric Correction: A Preliminary Study for GOCI-II. Remote Sens. 2020, 12, 3791. [Google Scholar] [CrossRef]
  60. Li, C.; Tao, B.; Liu, Y.; Zhang, S.; Zhang, Z.; Song, Q.; Jiang, Z.; He, S.; Huang, H.; Mao, Z. Assessment of VIIRS on the Identification of Harmful Algal Bloom Types in the Coasts of the East China Sea. Remote Sens. 2022, 14, 2089. [Google Scholar] [CrossRef]
  61. Jiang, Z.; Chen, J.; Gao, Y.; Zhai, H.; Jin, H.; Zhou, F.; Shou, L.; Yan, X.; Chen, Q. Regulation of Spatial Changes in Phytoplankton Community by Water Column Stability and Nutrients in the Southern Yellow Sea. J. Geophys. Res. Biogeosci. 2019, 124, 2610–2627. [Google Scholar] [CrossRef]
  62. Jiang, Z.; Liao, Y.; Liu, J.; Shou, L.; Chen, Q.; Yan, X.; Zhu, G.; Zeng, J. Effects of Fish Farming on Phytoplankton Community under the Thermal Stress Caused by a Power Plant in a Eutrophic, Semi-Enclosed Bay: Induce Toxic Dinoflagellate (Prorocentrum minimum) Blooms in Cold Seasons. Mar. Pollut. Bull. 2013, 76, 315–324. [Google Scholar] [CrossRef] [PubMed]
  63. GB 17378.7–2007; Marine Monitoring Specification Seventh Part: Coastal Pollution Ecological Investigation and Biological Monitoring. China Standard Press: Beijing, China, 2007.
  64. Barnes, B.B.; Cannizzaro, J.P.; English, D.C.; Hu, C. Validation of VIIRS and MODIS Reflectance Data in Coastal and Oceanic Waters: An Assessment of Methods. Remote Sens. Environ. 2019, 220, 110–123. [Google Scholar] [CrossRef]
  65. Wynne, T.T.; Stumpf, R.P.; Tomlinson, M.C.; Warner, R.A.; Tester, P.A.; Dyble, J.; Fahnenstiel, G.L. Relating Spectral Shape to Cyanobacterial Blooms in the Laurentian Great Lakes. Int. J. Remote Sens. 2008, 29, 3665–3672. [Google Scholar] [CrossRef]
Figure 1. 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.
Figure 1. 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.
Remotesensing 16 02304 g001
Figure 2. (a) Dongou oceanographic platform, (b) SeaPRISM radiometer, (c) Runjiang No. 1 experimental ship, (d) Cruise AOP observation system.
Figure 2. (a) Dongou oceanographic platform, (b) SeaPRISM radiometer, (c) Runjiang No. 1 experimental ship, (d) Cruise AOP observation system.
Remotesensing 16 02304 g002
Figure 3. Rrc spectra of typical clean water, algal bloom water, medium turbid water, and turbid water, (a) original Rrc spectra and (b) normalized Rrc spectra.
Figure 3. Rrc spectra of typical clean water, algal bloom water, medium turbid water, and turbid water, (a) original Rrc spectra and (b) normalized Rrc spectra.
Remotesensing 16 02304 g003
Figure 4. Scatter plots showing BLIs derived from (ac) in situ measured or (df) GOCI-II Rrs with BLIs derived from GOCI-II Rrc on the Dongou platform in 2021.
Figure 4. Scatter plots showing BLIs derived from (ac) in situ measured or (df) GOCI-II Rrs with BLIs derived from GOCI-II Rrc on the Dongou platform in 2021.
Remotesensing 16 02304 g004
Figure 5. (a) Schematic diagram illustrated the underway observation section of the LORCE on 19 August 2021, and the yellow line represented the track. (bf) BLIs comparison between the underway measured Rrs and GOCI-II-derived Rrc. Three GOCI-II images from the closest times to the measured Rrs 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.
Figure 5. (a) Schematic diagram illustrated the underway observation section of the LORCE on 19 August 2021, and the yellow line represented the track. (bf) BLIs comparison between the underway measured Rrs and GOCI-II-derived Rrc. Three GOCI-II images from the closest times to the measured Rrs 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.
Remotesensing 16 02304 g005
Figure 6. Comparison between long-term Rrc and Rrs calculated BLIs obtained by GOCI-II at Dongou platform location in 2021, (ae) showing comparison results of SS490, CI, DI, FLH, and MCI, respectively.
Figure 6. Comparison between long-term Rrc and Rrs calculated BLIs obtained by GOCI-II at Dongou platform location in 2021, (ae) showing comparison results of SS490, CI, DI, FLH, and MCI, respectively.
Remotesensing 16 02304 g006
Figure 7. 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 ax) show the distribution of different baseline indices for two P. donhaiense blooms, one diatom bloom, and one A. sanguinea bloom, and the fifth row (labeled yD) 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 Figure 8).
Figure 7. 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 ax) show the distribution of different baseline indices for two P. donhaiense blooms, one diatom bloom, and one A. sanguinea bloom, and the fifth row (labeled yD) 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 Figure 8).
Remotesensing 16 02304 g007
Figure 8. Scatterplots showing the distributions of different BLIs with TI, with each subplot corresponding to the selected regions boxed in Figure 7, (ay) correspond to Figure 7 (bl,nr,tx,zD), 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.
Figure 8. Scatterplots showing the distributions of different BLIs with TI, with each subplot corresponding to the selected regions boxed in Figure 7, (ay) correspond to Figure 7 (bl,nr,tx,zD), 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.
Remotesensing 16 02304 g008
Figure 9. (a) Schematic procedure of GOCI-II synthetical SS490 method for detecting algal blooms. Take a P. donghaiense bloom on 1 May 2023 as an example, (b) RGB true-color composite map, (c) original SS490 map, (d) SS490 distribution after cloud mask, (e) SS490 distribution after turbid correction, and (f) algal bloom area determined by synthetical SS490.
Figure 9. (a) Schematic procedure of GOCI-II synthetical SS490 method for detecting algal blooms. Take a P. donghaiense bloom on 1 May 2023 as an example, (b) RGB true-color composite map, (c) original SS490 map, (d) SS490 distribution after cloud mask, (e) SS490 distribution after turbid correction, and (f) algal bloom area determined by synthetical SS490.
Remotesensing 16 02304 g009
Figure 10. Scatterplots showing measured algal cell abundance versus different Rrc-derived (a) SS490, (b) CI, (c) DI, (d) FLH and (e) MCI, respectively.
Figure 10. Scatterplots showing measured algal cell abundance versus different Rrc-derived (a) SS490, (b) CI, (c) DI, (d) FLH and (e) MCI, respectively.
Remotesensing 16 02304 g010
Figure 11. Comparative validation plot of algal bloom detection accuracy based on different Rrc-derived (a) LHR, (b) RAB, (c) RI and (d) SS510, respectively.
Figure 11. Comparative validation plot of algal bloom detection accuracy based on different Rrc-derived (a) LHR, (b) RAB, (c) RI and (d) SS510, respectively.
Remotesensing 16 02304 g011
Figure 12. (ac) The RGB image and (df) Rrc, Rrs spectra corresponding to the green points with detection errors in Figure 11.
Figure 12. (ac) The RGB image and (df) Rrc, Rrs spectra corresponding to the green points with detection errors in Figure 11.
Remotesensing 16 02304 g012
Table 1. Information of algal blooms in the ECS during 2021–2023.
Table 1. Information of algal blooms in the ECS during 2021–2023.
Algal Bloom TypeDateCenter
Longitude
Center
Latitude
Cell Abundance
(×106 cell/L)
P. donghaiense27 April–17 May 2021121.086628.02460.01–41.6
4–7 June 2021120.804527.36793.43–4.59
7–26 May 2022120.910527.43020.15–28.4
2–13 June 2022120.743827.35790.02–74.6
18 August 2022121.358327.67500.72
29 April–28 May 2023120.937827.43630.07–9.85
30 August 2023120.524127.27640.95
A. sanguinea17–28 September 2021121.157327.81360.55–6.10
Diatom1–8 September 2021121.214527.96515.02–6.71
24 August 2022120.935827.96754.00
3 July 2023120.543527.30490.08
27 August 2023120.524126.427611.48
6 July 2023121.081227.79820.08
Table 2. Five BLIs used to detect algal blooms and their formulas.
Table 2. Five BLIs used to detect algal blooms and their formulas.
IndexAlgorithmGOCI-IIReference
SS490 R r c ( λ ) + R r c ( λ + ) R r c ( λ ) ( λ + λ ) ( λ λ ) R r c ( λ ) λ = 443, λ = 490, λ+ = 555Cannizzaro et al., 2019 [42]
CI R r c ( λ ) R r c ( λ ) + R r c ( λ + ) R r c ( λ ) ( λ + λ ) ( λ λ ) λ = 490, λ = 555, λ+ = 620Hu et al., 2011 [33]
DIλ = 555, λ = 620, λ+ = 660Tao et al., 2015 [47]
FLHλ = 660, λ = 680, λ+ = 745Hu et al., 2005 [13]
MCIλ = 660, λ = 709, λ+ = 745Gower et al., 2005 [31]
Table 3. Pixel-based statistics determined by employing confusion matrixes for comparing various algal bloom detection approaches.
Table 3. Pixel-based statistics determined by employing confusion matrixes for comparing various algal bloom detection approaches.
AlgorithmsThreshold(s)A
(b-B)
B
(b-NB)
C
(nb-B)
D
(nb-NB)
Sensitivity
A/(A + B)
Precision
A/(A + C)
False Neg. %
B/(A + B)
False Pos. %
C/(C + D)
FM
SS4900.002233090.8810.1200.97
CI0.005188540.690.780.310.560.76
DI0.000188720.690.720.310.780.71
FLH0.0011016360.380.770.620.330.64
MCI0.000206270.770.910.230.220.88
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, C.; Tao, B.; Li, Y.; Ai, L.; Zhu, Y.; Liang, L.; Huang, H.; Li, C. Evaluation of Rayleigh-Corrected Reflectance on Remote Detection of Algal Blooms in Optically Complex Coasts of East China Sea. Remote Sens. 2024, 16, 2304. https://doi.org/10.3390/rs16132304

AMA Style

Zhang C, Tao B, Li Y, Ai L, Zhu Y, Liang L, Huang H, Li C. Evaluation of Rayleigh-Corrected Reflectance on Remote Detection of Algal Blooms in Optically Complex Coasts of East China Sea. Remote Sensing. 2024; 16(13):2304. https://doi.org/10.3390/rs16132304

Chicago/Turabian Style

Zhang, Chengxin, Bangyi Tao, Yunzhou Li, Libo Ai, Yixian Zhu, Liansong Liang, Haiqing Huang, and Changpeng Li. 2024. "Evaluation of Rayleigh-Corrected Reflectance on Remote Detection of Algal Blooms in Optically Complex Coasts of East China Sea" Remote Sensing 16, no. 13: 2304. https://doi.org/10.3390/rs16132304

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop