[go: up one dir, main page]

 
 
remotesensing-logo

Journal Browser

Journal Browser

Remote Sensing of Ocean Colour

A special issue of Remote Sensing (ISSN 2072-4292). This special issue belongs to the section "Ocean Remote Sensing".

Deadline for manuscript submissions: closed (31 July 2018) | Viewed by 204245

Special Issue Editors


E-Mail Website
Guest Editor
Plymouth Marine Laboratory, Plymouth PL1 3DH, UK
Interests: satellite and biological oceanography; marine remote sensing; ocean colour; phytoplankton; climate change; tropical ecosystems

E-Mail Website
Guest Editor
Centre for Geography and Environmental Science, College of Life and Environmental Sciences, University of Exeter (Penryn Campus), Peter Lanyon Building, Cornwall, Penryn TR10 9EZ, UK
Interests: detection of phytoplankton size structure from satellite data; bio-opical model development and validation; unravelling the interaction between phytoplankton; physical forcing at large temporal and spatial scales using satellite observations

E-Mail Website
Guest Editor
Plymouth Marine Laboratory, Plymouth PL1 3DH, UK
Interests: marine ecosystem dynamics; climate change impacts, risks, opportunities and trade-offs; ocean-colour remote sensing; EO applications for aquatic-system health-risk assessment; ecology of microbial pathogens
Special Issues, Collections and Topics in MDPI journals

E-Mail Website
Guest Editor
IRD/OCEANS/LOPS, IUEM Technopole Brest Iroise Batiment D Rue Dumont D'Urville, Fr-29280 Plouzané, France
Interests: ocean colour remote sensing; physical-biological interactions combining satellite information and model-derived data. Physical-biological interactions combining satellite; observations and model-derived data from seasonal to decadal variability in the global ocean

E-Mail Website
Guest Editor
Plymouth Marine Laboratory, Prospect Place, The Hoe, Plymouth, UK
Interests: ocean colour modelling; spectral characteristics of light penetration underwater; bio-optical properties of phytoplankton; modelling primary production; bio-geochemical cycles in the sea; climate change; biological–physical interactions in the marine system; ecological provinces in the sea; ecological indicators and phytoplankton functional types

E-Mail Website
Guest Editor
Climate Change Center, King Abdullah University of Science and Technology (KAUST), Thuwal, Saudi Arabia
Interests: ocean modelling; data assimilation; remote sensing; climate change
Special Issues, Collections and Topics in MDPI journals

Special Issue Information

Dear Colleagues,

Understanding the functioning of marine ecosystems, their response to global pressures (climate change, pollution, overharvesting), and forecasting their fate, requires investigations of their past and present states. However, to date, large-scale biological dynamics remain poorly understood in many regions of the global oceans, often due to limited availability of adequate in-water measurements. To improve our knowledge on the functioning of marine ecosystems, an inter-disciplinary approach is necessary, taking advantage of complementary biophysical observations. Sensors on-board satellite platforms sample the Earth at synoptic temporal and spatial scales, offering a cost-effective approach to study biophysical fields and their interactions. In some regions, satellite sensors provide the only available spatially comprehensive biological datasets, covering the last two decades. These are the phytoplankton variables (including chlorophyll, primary production, phytoplankton phenology, phytoplankton functional types or PFTs, including harmful algal species, etc.) derived from satellite measurements of ocean colour. Ocean colour is also used to map other biotic and abiotic products, including suspended sediment load and light absorption by coloured dissolved organic matter.

In this Special Issue, we encourage submissions focusing on ocean colour applications, including, but not limited to:

  • Changes/trends/shifts in ocean colour observations
  • Interactions between ocean colour observations and higher trophic levels, including zooplankton and fisheries
  • Biophysical and climate interactions
  • Ocean colour algorithm development, validation and calibration
  • Remotely sensed PFTs including Harmful Algal Blooms (HABs)
  • Assimilation of ocean colour and other applications of ocean-colour products in modelling

We particularly encourage submissions of multidisciplinary approach (merging remotely-sensed ocean colour observations with in situ and modelled datasets) addressing ecological issues.

Dr. Dionysios Raitsos
Dr. Robert Brewin
Dr. Marie-Fanny Racault
Dr. Elodie Martinez
Prof. Shubha Sathyendranath
Prof. Ibrahim Hoteit
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Remote Sensing is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2700 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Keywords

  • Ocean Colour
  • Phytoplankton Functional Types - PFTs
  • Harmful Algal Blooms – HABs
  • Development, validation and calibration of Ocean colour algorithms
  • Assimilation and modelling of Ocean Colour

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (34 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Research

Jump to: Other

25 pages, 8854 KiB  
Article
Modelling the Vertical Distribution of Phytoplankton Biomass in the Mediterranean Sea from Satellite Data: A Neural Network Approach
by Michela Sammartino, Salvatore Marullo, Rosalia Santoleri and Michele Scardi
Remote Sens. 2018, 10(10), 1666; https://doi.org/10.3390/rs10101666 - 21 Oct 2018
Cited by 28 | Viewed by 5499
Abstract
Knowledge of the vertical structure of the bio-chemical properties of the ocean is crucial for the estimation of primary production, phytoplankton distribution, and biological modelling. The vertical profiles of chlorophyll-a (Chla) are available via in situ measurements that are usually quite rare [...] Read more.
Knowledge of the vertical structure of the bio-chemical properties of the ocean is crucial for the estimation of primary production, phytoplankton distribution, and biological modelling. The vertical profiles of chlorophyll-a (Chla) are available via in situ measurements that are usually quite rare and not uniformly distributed in space and time. Therefore, obtaining estimates of the vertical profile of the Chla field from surface observations is a new challenge. In this study, we employed an Artificial Neural Network (ANN) to reconstruct the 3-Dimensional (3D) Chla field in the Mediterranean Sea from surface satellite estimates. This technique is able to reproduce the highly nonlinear nature of the relationship between different input variables. A large in situ dataset of temperature and Chla calibrated fluorescence profiles, covering almost all Mediterranean Sea seasonal conditions, was used for the training and test of the network. To separate sources of errors due to surface Chla and temperature satellite estimates, from errors due to the ANN itself, the method was first applied using in situ surface data and then using satellite data. In both cases, the validation against in situ observations shows comparable statistical results with respect to the training, highlighting the feasibility of applying an ANN to infer the vertical Chla field from surface in situ and satellite estimates. We also analyzed the usefulness of our approach to resolve the Chla prediction at small temporal scales (e.g., day) by comparing it with the most widely used Mediterranean climatology (MEDATLAS). The results demonstrated that, generally, our method is able to reproduce the most reliable profile of Chla from synoptical satellite observations, thus resolving finer spatial and temporal scales with respect to climatology, which can be crucial for several marine applications. We demonstrated that our 3D reconstructed Chla field could represent a valid alternative to overcome the absence or discontinuity of in situ sampling. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Spatial distribution of the stations used for the training set (black crosses in the image, about 70% of the total stations) and test set (red crosses in the image, corresponding to 30% of the total stations). The map grid (1° × 1°) is used to select the training and test set stations uniformly in the spatial domain.</p>
Full article ">Figure 2
<p>BPN structure and configuration (7-10-1), with 7 input variables, 10 hidden nodes, and 1 output.</p>
Full article ">Figure 3
<p>Training phase (<b>left panel</b>) results and network’s performance evaluation on test set (<b>right panel</b>). Density plot of the observed Chla concentration (on x axis) vs. the BPN estimated Chla values (y axis). The Chla concentrations are reported in log-scale. The statistics for log-transformed values are r<sup>2</sup> = 0.72, RMSE = 0.23, and bias = 0.01 for the training and r<sup>2</sup> = 0.72, RMSE = 0.23, and bias = 0.02 for the test set, respectively. In black, the 1:1 bisector.</p>
Full article ">Figure 4
<p>Comparison of the mean profiles of observed values (black line) and BPN predictions (red line) inferred from <span class="html-italic">in situ</span> surface data and grouped according to different surface Chla concentration classes (in bold on top of each panel; <b>a</b>–<b>i</b>). For each panel (<b>a</b>–<b>i</b>) the black line represents the mean profile computed from the average of all observed profiles falling within a specific surface Chla class and the red line is the mean profile of all correspondent BPN-predicted profiles. The shaded area identifies the standard deviation of the two mean profiles, respectively, the original (grey area) and network-estimated ones (red area). The classes for which there were no profiles to be averaged are not represented. These profiles are related to the test set.</p>
Full article ">Figure 5
<p>Comparison of the observed Chla concentrations vs. BPN modelled Chla inferred from satellite data in the matchup stations. Density plot of the observed Chla concentration (on x axis) vs. the BPN estimated Chla values (y axis). The Chla concentrations are reported in log-scale. For log-transformed values, the statistics are r<sup>2</sup> = 0.69, RMSE = 0.24, and bias = −0.01. In black, the 1:1 bisector.</p>
Full article ">Figure 6
<p>Comparison of the mean profiles of observed values (black line) and BPN predictions (red line) inferred from satellite data and grouped according to different surface Chla concentration classes (in bold on top of each panel; <b>a</b>–<b>i</b>). For each panel (<b>a</b>–<b>i</b>) the black line represents the mean profile computed from the average of all observed profiles falling within a specific surface Chla class and the red line is the mean profile of all correspondent BPN-predicted profiles. The shaded area refers to the standard deviation of the two mean profiles, respectively: the original (grey area) and network-estimated ones (red area). The classes for which there were no profiles to be averaged are not represented. These profiles are related to the assessment of the BPN performance to infer vertical profiles from satellite data.</p>
Full article ">Figure 7
<p>An example of the BPN application on a specified transect in the NwMed (Latitude: 39.12°–42.8° N, Longitude: 4.89° E). Contour plots of the vertical Chla field estimated by the BPN using monthly satellite data as inputs, for the year 2009.</p>
Full article ">Figure 8
<p>Satellite monthly maps of Chla concentration for the year 2009. The red line refers to the transect of vertical reconstruction in <a href="#remotesensing-10-01666-f007" class="html-fig">Figure 7</a>.</p>
Full article ">Figure 9
<p>Contour plot of the chosen transect extracted from NB02 oceanographic cruise. The data are sampled in the Gulf of Lion from 18–12–2001 to 20–12–2001. Panels show the vertical section of (<b>a</b>) observed Chla field, (<b>b</b>) reconstructed field from <span class="html-italic">in situ</span> surface Chla, (<b>c</b>) reconstructed field from satellite data, and (<b>d</b>) extracted from the fall MEDTALAS climatology. The black lines in (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) identify the station points. On the bottom, the three plots show, for each depth, the mean absolute error of computed differences between the observed and modelled data as (<b>e</b>) predicted from <span class="html-italic">in situ</span> surface Chla, (<b>f</b>) from satellite observations, and (<b>g</b>) extracted from fall MEDTALAS climatology.</p>
Full article ">Figure 10
<p>Upper panel shows the scatterplots relative to the comparison on matchup database (for details see <a href="#sec2dot2-remotesensing-10-01666" class="html-sec">Section 2.2</a>) of observed vs. BPN modelled data and climatology. In blue, the comparison of observed values (x axis) against predictions from superficial <span class="html-italic">in situ</span> data (y axis); in red, comparison against the Chla values predicted from satellite data (y axis); and finally, in green, the comparison against annual climatology (y axis). N is the number of available data. Bottom panel shows (<b>a</b>) the MBE and (<b>b</b>) the mean absolute error computed for each depth, for the three cases.</p>
Full article ">
17 pages, 12579 KiB  
Article
Improved Detection of Tiny Macroalgae Patches in Korea Bay and Gyeonggi Bay by Modification of Floating Algae Index
by Ahmed Harun-Al-Rashid and Chan-Su Yang
Remote Sens. 2018, 10(9), 1478; https://doi.org/10.3390/rs10091478 - 16 Sep 2018
Cited by 5 | Viewed by 4833
Abstract
This work focuses on the detection of tiny macroalgae patches in the eastern parts of the Yellow Sea (YS) using high-resolution Landsat-8 images from 2014 to 2017. In the comparison between floating algae index (FAI) and normalized difference vegetation index (NDVI) better detection [...] Read more.
This work focuses on the detection of tiny macroalgae patches in the eastern parts of the Yellow Sea (YS) using high-resolution Landsat-8 images from 2014 to 2017. In the comparison between floating algae index (FAI) and normalized difference vegetation index (NDVI) better detection by FAI was observed, but many tiny patches still remained undetected. By applying a modification on the FAI around 12% to 27% increased and correct detection of macroalgae is achieved from 35 images compared to the original. Through this method many scattered tiny patches were detected in June or July in Korea Bay and Gyeonggi Bay. Though it was a small-scale phenomenon they occurred in the similar period of macroalgal bloom occurrence in the YS. Thus, by using this modified method we could detect macroalgae in the study areas around one month earlier than the previously used Geostationary Ocean Color Imager NDVI-based detection. Later, more macroalgae patches including smaller ones occupying increased areas were detected. Thus, it seems that those macroalgae started growing locally from tiny patches rather than being transported from the western parts of the YS. Therefore, this modified FAI could be used for the precise detection of macroalgae. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>Study area map displaying Landsat-8 path/row coverage by some natural colour composite images (RGB 4, 3 and 2 respectively for 655, 561 and 482 nm) on 2 August 2014 (116/34), 6 June 2014 (117/33 and 117/34), 5 August 2016 (118/32 and 118/33), 25 June 2016 (119/34), and 25 July 2015 (119/35). Three large white rectangles indicate the study areas, and the three small red rectangles and the large black rectangle indicate the ROI for showing the comparison between detection results by different vegetation indices: NDVI, FAI, and mod-FAI.</p>
Full article ">Figure 2
<p>Flow diagram for the modified FAI-based macroalgae detection.</p>
Full article ">Figure 3
<p>Comparison of macroalgal bloom detection results (dark red) between (<b>a</b>) Geostationary Ocean Color Imager (GOCI) normalized difference vegetation index (NDVI), (<b>b</b>) Landsat-8 FAI, and (<b>c</b>) Landsat-8 modified FAI. The GOCI image was acquired on 25 July 2015 03:31 UTC, and the Landsat-8 based images represent for ROI area inside path/row 119/35 image (large black rectangle in <a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>) at 02:30 UTC on the same date. The small ROI area inside each image is marked by a black rectangle.</p>
Full article ">Figure 4
<p>Comparison of macroalgae patches between Landsat-8 based (<b>a</b>,<b>b</b>) are true colour RGB (bands 4, 3, 2) images after applying different contrast stretches for better visibility of the macroalgae patches at different locations; (<b>c</b>) FAI and (<b>d</b>) modified FAI for the smaller ROI area (red rectangle in <a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>) in path/row 119/35 on 25 July 2015. The colourmaps of FAI and modified FAI are shown as (<b>e</b>) and (<b>f</b>), respectively.</p>
Full article ">Figure 5
<p>Landsat-8 based macroalgae detection results for the ROI area in <a href="#remotesensing-10-01478-f004" class="html-fig">Figure 4</a> by (<b>a</b>) only FAI, (<b>b</b>) only modified FAI, and (<b>c</b>) combined detection results. For FAI and modified FAI threshold values of 0 and 0.001, respectively, were used. Respective true colour composite RGB image is shown as (<b>d</b>) after applying different colour stretches to different portions of the image.</p>
Full article ">Figure 6
<p>Comparison of Landsat-8 based vegetation indices colourmaps of (<b>a</b>) NDVI, (<b>c</b>) FAI, and (<b>e</b>) modified FAI with the thresholding-based detected macroalgae in the corresponding vegetation indices (<b>b</b>,<b>d</b>, and <b>f</b>, respectively) for the small ROI area (<a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>) of path/row 116/34 on 2 August 2014 at 02:11 UTC. The respective (<b>g</b>) common as well as only FAI and modified FAI-based detection, and (<b>h</b>) the natural colour composite image (Landsat bands 4, 3, and 2) are also compared.</p>
Full article ">Figure 7
<p>Comparison of Landsat-8 based vegetation indices colourmaps of (<b>a</b>) NDVI, (<b>c</b>) FAI, and (<b>e</b>) modified FAI with the thresholding-based detected macroalgae in the corresponding vegetation indices (<b>b</b>,<b>d</b>, and <b>f</b>, respectively) for the small ROI area (<a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>) of path/row 119/34 on 25 June 2016 at 02:29 UTC. The respective (<b>g</b>) common as well as only FAI and modified FAI-based detection, and (<b>h</b>) the natural colour composite image (Landsat bands 4, 3, and 2) are also compared.</p>
Full article ">Figure 8
<p>Changes in the distribution of macroalgae and sea pixels due to the application of modification to the FAI. Results of (<b>b</b>) to (<b>f</b>) are shown for the whole image areas of respective Landsat-8 images except (<b>a</b>) showing the results for the area in <a href="#remotesensing-10-01478-f004" class="html-fig">Figure 4</a>e,f. Vertical hatch lines are shown along the modified FAI threshold values (mod-FAI).</p>
Full article ">Figure 9
<p>Detection results of macroalgae from Landsat-8 modified FAI images for path/row 116/34. Detected macroalgae pixels are exaggerated in size for better visibility. For the map area refer to the dashed rectangle in <a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 10
<p>Detection results of macroalgae from Landsat-8 modified FAI images for path/row 117/33 and 117/34. Detected macroalgae pixels are exaggerated in size for better visibility. For the map area refer to the largest dashed rectangle in <a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 11
<p>Detection results of macroalgae from Landsat-8 modified FAI images for path/row 118/32 and 118/33. Detected macroalgae pixels are exaggerated in size for better visibility. For the map area refer to the smaller dashed rectangle <a href="#remotesensing-10-01478-f001" class="html-fig">Figure 1</a>.</p>
Full article ">
21 pages, 1302 KiB  
Article
A 55-Year Time Series Station for Primary Production in the Adriatic Sea: Data Correction, Extraction of Photosynthesis Parameters and Regime Shifts
by Žarko Kovač, Trevor Platt, Živana Ninčević Gladan, Mira Morović, Shubha Sathyendranath, Dionysios E. Raitsos, Branka Grbec, Frano Matić and Jere Veža
Remote Sens. 2018, 10(9), 1460; https://doi.org/10.3390/rs10091460 - 12 Sep 2018
Cited by 20 | Viewed by 4469
Abstract
In 1962, a series of in situ primary production measurements began in the Adriatic Sea, at a station near the island of Vis. To this day, over 55 years of monthly measurements through the photic zone have been accumulated, including close to 3000 [...] Read more.
In 1962, a series of in situ primary production measurements began in the Adriatic Sea, at a station near the island of Vis. To this day, over 55 years of monthly measurements through the photic zone have been accumulated, including close to 3000 production measurements at different depths. The measurements are conducted over a six-hour period around noon, and the average production rate extrapolated linearly over day length to calculate daily production. Here, a non-linear primary production model is used to correct these estimates for potential overestimation of daily production due to linear extrapolation. The assimilation numbers are recovered from the measured production profiles and subsequently used to model production at depth. Using the recovered parameters, the model explained 87% of variability in measured normalized production at depth. The model is then used to calculate daily production at depth, and it is observed to give on average 20% lower daily production at depth than the estimates based on linear extrapolation. Subsequently, water column production is calculated, and here, the model predicted on average 26% lower water column production. With the recovered parameters and the known magnitude of the overestimation, the time-series of water column production is then re-established with the non-linearly-corrected data. During this 55-year period, distinct regimes were observed, which were classified with a regime shift detection method. It is then demonstrated how the recovered parameters can be used in a remote sensing application. A seasonal cycle of the recovered assimilation number is constructed along with the seasonal cycle of remotely-sensed chlorophyll. The two are then used to model the seasonal cycle of water column production. An upper and a lower bound on the seasonal cycle of water column production based on remotely-sensed chlorophyll data are then presented. Measured water column production was found to be well within the range of remotely-sensed estimates. With this work, the utility of in situ measurements as a means of providing information on the assimilation number is presented and its application as a reference for remote sensing models highlighted. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Percentage difference in daily production between the estimate based on linear extrapolation of average production during a six-hour interval around noon (<a href="#FD5-remotesensing-10-01460" class="html-disp-formula">5</a>) and the analytical solution for daily production (<a href="#FD6-remotesensing-10-01460" class="html-disp-formula">6</a>). The orange curve corresponds to the Julian day with the longest day length <math display="inline"><semantics> <msub> <mi>D</mi> <mi>max</mi> </msub> </semantics></math> and the blue with the shortest day length <math display="inline"><semantics> <msub> <mi>D</mi> <mi>min</mi> </msub> </semantics></math>, at the latitude of the Stončica station.</p>
Full article ">Figure 2
<p>Histogram of estimated values of the assimilation number <math display="inline"><semantics> <msubsup> <mi>P</mi> <mrow> <mi>m</mi> </mrow> <mi>B</mi> </msubsup> </semantics></math> obtained from 185 cruises in the Adriatic Sea. The abscissa corresponds to parameter value, and the ordinate gives the percentage of cruises that fall into a certain interval of parameter values.</p>
Full article ">Figure 3
<p>Measured versus modelled normalized production at depth. Production at depth is measured during a 6-h incubation interval and normalized to biomass measurement at the corresponding depth. Modelled production is given by numerical integration of (<a href="#FD3-remotesensing-10-01460" class="html-disp-formula">3</a>) with the recovered photosynthesis parameters. The <math display="inline"><semantics> <msup> <mi>r</mi> <mn>2</mn> </msup> </semantics></math> between the measured and the modelled normalized production at depth is 0.87. In total, there are 1040 points.</p>
Full article ">Figure 4
<p>Comparison of water column production <math display="inline"><semantics> <msub> <mover accent="true"> <mi>P</mi> <mo>˜</mo> </mover> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math> (abscissa) calculated by (<a href="#FD9-remotesensing-10-01460" class="html-disp-formula">9</a>) from extrapolated daily production at depth and water column production <math display="inline"><semantics> <msub> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math> (ordinate) calculated by (<a href="#FD10-remotesensing-10-01460" class="html-disp-formula">10</a>) from the analytical solution for daily production, with parameters estimated from the measured production profiles. The dashed grey line is the 1:1 line, and the thick grey line is the linear regression between <math display="inline"><semantics> <msub> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math> and <math display="inline"><semantics> <msub> <mover accent="true"> <mi>P</mi> <mo>˜</mo> </mover> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math>.</p>
Full article ">Figure 5
<p>Annual running means of water column production from April 1962–July 2017. The original time-series, obtained by application of (<a href="#FD9-remotesensing-10-01460" class="html-disp-formula">9</a>), is given in grey. The corrected time-series, obtained by application of (<a href="#FD10-remotesensing-10-01460" class="html-disp-formula">10</a>), is given in thick blue. Start of each regime is indicated by an orange circle with the corresponding year above it. The red line at the beginning of the sampling era highlights the period without chlorophyll measurements.</p>
Full article ">Figure 6
<p>Time-series of measured (orange) and remotely-sensed chlorophyll (grey). Measured chlorophyll corresponds to the vertically-averaged chlorophyll concentration for each profile. The grey curve is the mean remotely-sensed chlorophyll concentration, and the shaded area corresponds to the min-max range. Data are from September 1997–December 2016.</p>
Full article ">Figure 7
<p>(<b>a</b>) Seasonal cycles of remotely-sensed (green points) and fitted chlorophyll (green curve), alongside with the seasonal cycle of the assimilation number <math display="inline"><semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics></math> (thin blue curve) and the fitted <math display="inline"><semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics></math> (thick blue curve). (<b>b</b>) Estimated water column production <math display="inline"><semantics> <msub> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math> using remotely-sensed chlorophyll and <math display="inline"><semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics></math> from above, with the climatological (light grey) and one half the climatological (dark grey) values for noon irradiance on each Julian day, alongside <math display="inline"><semantics> <msub> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics></math> obtained by using (<a href="#FD10-remotesensing-10-01460" class="html-disp-formula">10</a>) (thin orange curve) and the fitted seasonal cycle (thick orange curve).</p>
Full article ">
21 pages, 8258 KiB  
Article
Applications of DINEOF to Satellite-Derived Chlorophyll-a from a Productive Coastal Region
by Andrea Hilborn and Maycira Costa
Remote Sens. 2018, 10(9), 1449; https://doi.org/10.3390/rs10091449 - 11 Sep 2018
Cited by 47 | Viewed by 7564
Abstract
A major limitation for remote sensing analyses of oceanographic variables is loss of spatial data. The Data INterpolating Empirical Orthogonal Functions (DINEOF) method has demonstrated effectiveness for filling spatial gaps in remote sensing datasets, making them more easily implemented in further applications. However, [...] Read more.
A major limitation for remote sensing analyses of oceanographic variables is loss of spatial data. The Data INterpolating Empirical Orthogonal Functions (DINEOF) method has demonstrated effectiveness for filling spatial gaps in remote sensing datasets, making them more easily implemented in further applications. However, the spatial and temporal coverage of the input image dataset can heavily impact the outcomes of using this method and, thus, further metrics derived from these datasets, such as phytoplankton bloom phenology. In this study, we used a three-year time series of MODIS-Aqua chlorophyll-a to evaluate the DINEOF reconstruction output accuracy corresponding to variation in the form of the input data used (i.e., daily or week composite scenes) and time series length (annual or three consecutive years) for a dynamic region, the Salish Sea, Canada. The accuracy of the output data was assessed considering the original chla pixels. Daily input time series produced higher accuracy reconstructing chla (95.08–97.08% explained variance, RMSExval 1.49–1.65 mg m−3) than did all week composite counterparts (68.99–76.88% explained variance, RMSExval 1.87–2.07 mg m−3), with longer time series producing better relationships to original chla pixel concentrations. Daily images were assessed relative to extracted in situ chla measurements, with original satellite chla achieving a better relationship to in situ matchups than DINEOF gap-filled chla, and with annual DINEOF-processed data performing better than the multiyear. These results contribute to the ongoing body of work encouraging production of ocean color datasets with consistent processing for global purposes such as climate change studies. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Salish Sea, oceanic and geographic features, and population centers. The region includes the Juan de Fuca Strait (JFS), Strait of Georgia (SoG), Puget Sound (PS), and Queen Charlotte Strait (QCS). QCS is included in this study considering its use in salmon migration research [<a href="#B26-remotesensing-10-01449" class="html-bibr">26</a>]. Locations of in situ <span class="html-italic">chla</span> matchups (<a href="#sec2dot4dot2-remotesensing-10-01449" class="html-sec">Section 2.4.2</a>) are indicated by blue (DINEOF-reconstructed <span class="html-italic">chla</span>) and blue-ringed circles (satellite and DINEOF-reconstructed <span class="html-italic">chla</span>).</p>
Full article ">Figure 2
<p>Temporal coverage displayed as (<b>a</b>) number of images per month and (<b>b</b>) percent spatial coverage of the study region per month. Presence of a given pixel is shown for (<b>c</b>) the daily time series and (<b>d</b>) week composite.</p>
Full article ">Figure 3
<p>D1 (<b>a</b>), W1 (<b>b</b>), D3 (<b>c</b>), and W3 (<b>d</b>) linear correlation results. The 40.00 mg m<sup>−3</sup> threshold (<a href="#sec2dot2dot1-remotesensing-10-01449" class="html-sec">Section 2.2.1</a>) is evident as a cutoff feature in all plots.</p>
Full article ">Figure 4
<p>Per-pixel R<sup>2</sup> of DINEOF results for D1 (<b>a</b>), W1 (<b>b</b>), D3 (<b>c</b>), and W3 (<b>d</b>).</p>
Full article ">Figure 5
<p>Daily reconstruction of February 28, 2014, shown as the original <span class="html-italic">chla<sub>sat</sub></span> (<b>a</b>), D1 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>b</b>), and D3 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>c</b>); similarly, the week composite <span class="html-italic">chla<sub>sat</sub></span> (<b>d</b>), W1 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>e</b>), and W3 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>f</b>). Salish Sea thalweg is shown in (<b>a</b>), with a gap excluding the region of no data in Johnstone Strait.</p>
Full article ">Figure 6
<p>Daily image time series shown as Hovmöller plot along Salish Sea thalweg (<span class="html-italic">y</span> axis, shown in <a href="#remotesensing-10-01449-f005" class="html-fig">Figure 5</a>a), contrasting <span class="html-italic">chla<sub>sat</sub></span> (<b>a</b>), D1 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>b</b>), and D3 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>c</b>) for 2014–2016. The dashed line represents a spatial gap in Johnstone Strait due to the inability of MODISA to resolve data in the narrow passages.</p>
Full article ">Figure 7
<p>Week composite time series extracted along the Salish Sea thalweg (<span class="html-italic">y</span> axis, <a href="#remotesensing-10-01449-f005" class="html-fig">Figure 5</a>a) for <span class="html-italic">chla<sub>sat</sub></span> (<b>a</b>)<span class="html-italic">,</span> W1 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>b</b>), and W3 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>c</b>) for 2014–2016.</p>
Full article ">Figure 8
<p>Statistical results for <span class="html-italic">chla<sub>insitu</sub></span> between <span class="html-italic">chla<sub>sat</sub></span> (<b>a</b>), D1 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>b</b>), and D3 <span class="html-italic">chla<sub>sat+rec</sub></span> (<b>c</b>). All <span class="html-italic">p</span>-values are &lt;0.05.</p>
Full article ">Figure 9
<p>Relationship between original and reconstructed pixel time series for a D3 example pixel located in the Fraser River plume (<b>a</b>) and in central JFS (<b>b</b>), and for the W3 reconstruction in (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">Figure A1
<p>Spatial median and shaded ±1 standard deviation for <span class="html-italic">chla<sub>sat+rec</sub></span> of D1/D3 (<b>a</b>), divided by year for legibility, and 2014–2016 for W1/W3 (<b>b</b>). Corresponding per-scene median <span class="html-italic">chla<sub>sat</sub></span> shown as black dots with ±1 standard deviation.</p>
Full article ">
17 pages, 5504 KiB  
Article
Evaluation of Semi-Analytical Algorithms to Retrieve Particulate and Dissolved Absorption Coefficients in Gulf of California Optically Complex Waters
by Stella Patricia Betancur-Turizo, Adriana González-Silvera, Eduardo Santamaría-del-Ángel, Jing Tan and Robert Frouin
Remote Sens. 2018, 10(9), 1443; https://doi.org/10.3390/rs10091443 - 10 Sep 2018
Cited by 5 | Viewed by 3764
Abstract
Two semi-analytical algorithms, Generalized Inherent Optical Property (GIOP) and Garver-Siegel-Maritorena (GSM), were evaluated in terms of how well they reproduced the absorption coefficient of phytoplankton (aph(λ)) and dissolved and detrital organic matter (adg(λ)) [...] Read more.
Two semi-analytical algorithms, Generalized Inherent Optical Property (GIOP) and Garver-Siegel-Maritorena (GSM), were evaluated in terms of how well they reproduced the absorption coefficient of phytoplankton (aph(λ)) and dissolved and detrital organic matter (adg(λ)) at three wavelengths (λ of 412, 443, and 488 nm) in a zone with optically complex waters, the Upper Gulf of California (UGC) and the Northern Gulf of California (NGC). In the UGC, detritus determines most of the total light absorption, whereas, in the NGC, chromophoric dissolved organic material (CDOM) and phytoplankton dominate. Upon comparing the results of each model with a database assembled from four cruises done from spring to summer (March through September) between 2011 and 2013, it was found that GIOP is a better estimator for aph(λ) than GSM, independently of the region. However, both algorithms underestimate in situ values in the NGC, whereas they overestimate them in the UGC. Errors are associated with the following: (a) the constant a*ph(λ) value used by GSM and GIOP (0.055 m2 mgChla−1) is higher than the most frequent value observed in this study’s data (0.03 m2 mgChla−1), and (b) satellite-derived chlorophyll a concentration (Chla) is biased high compared with in situ Chla. GIOP gave also better results for the adg(λ) estimation than GSM, especially in the NGC. The spectral slope Sdg was identified as an important parameter for estimating adg(λ), and this study’s results indicated that the use of a fixed input value in models was not adequate. The evaluation confirms the lack of generality of algorithms like GIOP and GSM, whose reflectance model is too simplified to capture expected variability. Finally, a greater monitoring effort is suggested in the study area regarding the collection of in situ reflectance data, which would allow explaining the effects that detritus and CDOM may have on the semi-analytical reflectance inversions, as well as isolating the possible influence of the atmosphere on the satellite-derived water reflectance and Chla. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Study area map. (<b>b</b>) Transition zone between the Upper Gulf of California (UGC) and Northern Gulf of California (NGC) bio-optical regions, indicated by the dotted line. (<b>c</b>) Station location for each cruise.</p>
Full article ">Figure 2
<p>Data of the six cruises that had in situ information on the coefficients (<b>a</b>) <span class="html-italic">a<sub>ph</sub></span>(<span class="html-italic">λ</span>) and (<b>b</b>) <span class="html-italic">a<sub>dg</sub></span>(<span class="html-italic">λ</span>) in the study zone. Black circles represent stations paired with the Generalized Inherent Optical Property (GIOP) algorithm, and red points represent stations paired with the Garver-Siegel-Maritorena (GSM) algorithm. The dotted black line represents the intermediate position of the transitional zone that separates the bio-optical regions UGC and NGC [<a href="#B29-remotesensing-10-01443" class="html-bibr">29</a>].</p>
Full article ">Figure 3
<p>Comparative analysis between in situ and satellite <span class="html-italic">a<sub>ph</sub></span>(412, 443, and 488 nm) for GIOP and GSM models, with the statistics Root Mean Square Error (RMSE), bias, <span class="html-italic">r<sub>p</sub></span>, and <span class="html-italic">χ</span><sup>2</sup>; the 1:1 line is indicated for reference. The green and blue colors correspond to the UGC and NGC regions, respectively. In the first column (<b>a</b>, <b>d</b>, <b>g</b>, <b>j</b>, <b>m</b>, <b>p</b>) the entire database was used, in the second (<b>b</b>, <b>e</b>, <b>h</b>, <b>k</b>, <b>n</b>, <b>q</b>) only data from UGC, and in the third (<b>c</b>, <b>f</b>, <b>i</b>, <b>l</b>, <b>o</b>, <b>r</b>) only data from NGC.</p>
Full article ">Figure 4
<p>Taylor diagram illustrating the relative performance of the GIOP and GSM algorithms upon reproducing the absorption coefficient <span class="html-italic">a<sub>ph</sub></span>(412, 443, and 488 nm). Diagrams represent (<b>a</b>) the entire dataset and data collected in the (<b>b</b>) UGC and (<b>c</b>) NGC. The red line represents the critical value of Pearson’s correlation coefficient and indicates the best-model performance.</p>
Full article ">Figure 5
<p>(<b>a</b>) In situ <span class="html-italic">a*<sub>ph</sub></span> (m<sup>2</sup> mgChla<span class="html-italic"><sup>−1</sup></span>) variability for all cruises analyzed in this study, including the value used in GIOP and GSM (<span class="html-italic">a*<sub>ph</sub></span> = 0.055 m<sup>2</sup> mgChl<span class="html-italic">a<sup>−1</sup></span>, dotted line). Also indicated are the frequency histograms for (<b>b</b>) UGC and (<b>c</b>) NGC.</p>
Full article ">Figure 6
<p>Relationship between in situ chlorophyll data (Chl<span class="html-italic">a</span>) and satellite chlorophyll data for the June 2008, June 2010, March 2011, August 2012, and June 2013 cruises, plotted on a logarithmic scale. The dashed line is the one-to-one line. RMSE is computed on log10-transformed data, and bias on original data.</p>
Full article ">Figure 7
<p>Comparative analysis between in situ and satellite <span class="html-italic">a<sub>dg</sub></span>(412, 443, and 448) for the GIOP and GSM algorithms, with statistics Root Mean Square Error (RMSE), bias, <span class="html-italic">r<sub>p</sub></span>, and <span class="html-italic">χ</span><sup>2</sup>. The 1:1 is indicated for reference. The green and blue colors correspond to the UGC and NGC regions, respectively. In the first column (<b>a</b>, <b>d</b>, <b>g</b>, <b>j</b>, <b>m</b>, <b>p</b>) the entire database was used, in the second (<b>b</b>, <b>e</b>, <b>h</b>, <b>k</b>, <b>n</b>, <b>q</b>) only data from UGC, and in the third (<b>c</b>, <b>f</b>, <b>i</b>, <b>l</b>, <b>o</b>, <b>r</b>) only data from NGC.</p>
Full article ">Figure 8
<p>Taylor diagram illustrating the relative performance of the algorithms GIOP and GSM in reproducing the absorption coefficient <span class="html-italic">a<sub>dg</sub></span>(412, 443, and 488 nm). Each diagram represents the analysis applied to (<b>a</b>) all of the data and by bio-optical regions (<b>b</b>) UGC and (<b>c</b>) NGC. The dotted red line represents the critical value of Pearson’s correlation coefficient.</p>
Full article ">Figure 9
<p>Average spectra of the absorption coefficient of dissolved and detrital matter (<span class="html-italic">a<sub>dg</sub></span>(<span class="html-italic">λ</span>)) calculated for each cruise in the UGC (<b>a</b>–<b>d</b>) and NGC (<b>e</b>–<b>g</b>). The pie chart inside diagrams represents the percentage contribution of detritus (<span class="html-italic">a<sub>d</sub></span>) and chromophoric dissolved organic material (CDOM) (<span class="html-italic">a<sub>g</sub></span>) to <span class="html-italic">a<sub>dg</sub></span>(<span class="html-italic">λ</span>). Open circles represent GSM <span class="html-italic">a<sub>dg</sub></span>(<span class="html-italic">λ</span>) values and black crosses represent GIOP <span class="html-italic">a<sub>dg</sub></span>(<span class="html-italic">λ</span>) values. Note: The August 2012 cruise was represented by a single station and was not included in the figure.</p>
Full article ">Figure 10
<p>(<b>a</b>) In situ <span class="html-italic">S<sub>dg</sub></span> variability for all cruises analyzed in this study, including the value used in GIOP (<span class="html-italic">S<sub>dg</sub></span> = 0.018, black dotted line) and GSM (<span class="html-italic">S<sub>dg</sub></span> = 0.02061, gray dotted line). Also indicated are the frequency histograms for (<b>b</b>) UGC and (<b>c</b>) NGC.</p>
Full article ">
16 pages, 1271 KiB  
Article
Can Multispectral Information Improve Remotely Sensed Estimates of Total Suspended Solids? A Statistical Study in Chesapeake Bay
by Nicole M. DeLuca, Benjamin F. Zaitchik and Frank C. Curriero
Remote Sens. 2018, 10(9), 1393; https://doi.org/10.3390/rs10091393 - 1 Sep 2018
Cited by 30 | Viewed by 4990
Abstract
Total suspended solids (TSS) is an important environmental parameter to monitor in the Chesapeake Bay due to its effects on submerged aquatic vegetation, pathogen abundance, and habitat damage for other aquatic life. Chesapeake Bay is home to an extensive and continuous network of [...] Read more.
Total suspended solids (TSS) is an important environmental parameter to monitor in the Chesapeake Bay due to its effects on submerged aquatic vegetation, pathogen abundance, and habitat damage for other aquatic life. Chesapeake Bay is home to an extensive and continuous network of in situ water quality monitoring stations that include TSS measurements. Satellite remote sensing can address the limited spatial and temporal extent of in situ sampling and has proven to be a valuable tool for monitoring water quality in estuarine systems. Most algorithms that derive TSS concentration in estuarine environments from satellite ocean color sensors utilize only the red and near-infrared bands due to the observed correlation with TSS concentration. In this study, we investigate whether utilizing additional wavelengths from the Moderate Resolution Imaging Spectroradiometer (MODIS) as inputs to various statistical and machine learning models can improve satellite-derived TSS estimates in the Chesapeake Bay. After optimizing the best performing multispectral model, a Random Forest regression, we compare its results to those from a widely used single-band algorithm for the Chesapeake Bay. We find that the Random Forest model modestly outperforms the single-band algorithm on a holdout cross-validation dataset and offers particular advantages under high TSS conditions. We also find that both methods are similarly generalizable throughout various partitions of space and time. The multispectral Random Forest model is, however, more data intensive than the single band algorithm, so the objectives of the application will ultimately determine which method is more appropriate. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of Chesapeake Bay estuary showing the 86 Chesapeake Bay Program measurement stations (black dots) used in the satellite-in situ matchup dataset in this study.</p>
Full article ">Figure 2
<p>Partial dependence plots for the 11 MODIS bands used as predictors in the Random Forest model before pruning.</p>
Full article ">Figure 3
<p>Log-log plots showing one-to-one regressions of CBP in situ measured versus satellite-derived TSS from (<b>A</b>) the pruned Random Forest model and (<b>B</b>) the O-2012 algorithm. Solid black line is 1:1 line, dotted line is the linear regression and dashed line shows the cutoff for higher TSS range analyses.</p>
Full article ">Figure 4
<p>Mapped comparisons of daily remotely sensed TSS (mg/L) derived from O-2012 (<b>A</b>–<b>C</b>) and RF model (<b>D</b>–<b>F</b>) for 2017 dates not included in model training or holdout datasets. In situ measurement values shown in color-filled black circles.</p>
Full article ">
22 pages, 5545 KiB  
Article
Estimation of Size-Fractionated Primary Production from Satellite Ocean Colour in UK Shelf Seas
by Kieran Curran, Robert J. W. Brewin, Gavin H. Tilstone, Heather A. Bouman and Anna Hickman
Remote Sens. 2018, 10(9), 1389; https://doi.org/10.3390/rs10091389 - 31 Aug 2018
Cited by 12 | Viewed by 5225
Abstract
Satellite ocean-colour based models of size-fractionated primary production (PP) have been developed for the oceans on a global level. Uncertainties exist as to whether these models are accurate for temperate Shelf seas. In this paper, an existing ocean-colour based PP model is tuned [...] Read more.
Satellite ocean-colour based models of size-fractionated primary production (PP) have been developed for the oceans on a global level. Uncertainties exist as to whether these models are accurate for temperate Shelf seas. In this paper, an existing ocean-colour based PP model is tuned using a large in situ database of size-fractionated measurements from the Celtic Sea and Western English Channel of chlorophyll-a (Chl a) and the photosynthetic parameters, the maximum photosynthetic rate ( P m B ) and light limited slope ( α B ). Estimates of size fractionated PP over an annual cycle in the UK shelf seas are compared with the original model that was parameterised using in situ data from the open ocean and a climatology of in situ PP from 2009 to 2015. The Shelf Sea model captured the seasonal patterns in size-fractionated PP for micro- and picophytoplankton, and generally performed better than the original open ocean model, except for nanophytoplankton PP which was over-estimated. The overestimation in PP is in part due to errors in the parameterisation of the biomass profile during summer, stratified conditions. Compared to the climatology of in situ data, the shelf sea model performed better when phytoplankton biomass was high, but overestimated PP at low Chl a. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sampling sites of Western English Channel and Celtic Sea. Green points are autumn campaign; Red points are summer campaign. Dotted box shows transect between Central Celtic Sea (CCS) stations (49–51.5° N) and off-shelf (48–49° N). L4 and E1 are the Western English Channel time series stations.</p>
Full article ">Figure 2
<p>Relationship of <math display="inline"><semantics> <mrow> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <msup> <mi>α</mi> <mi>B</mi> </msup> </mrow> </semantics></math> with respect to ζ for each size class using the retuned shelf model (black lines) with the models of Uitz et al. [<a href="#B35-remotesensing-10-01389" class="html-bibr">35</a>] (dotted lines) and Brewin et al. [<a href="#B68-remotesensing-10-01389" class="html-bibr">68</a>] (dashed lines). Values for light saturation parameter <math display="inline"><semantics> <mrow> <msub> <mi>E</mi> <mi>k</mi> </msub> </mrow> </semantics></math> of retuned model in bottom right panel.</p>
Full article ">Figure 3
<p>In situ measurements of surface Chl <span class="html-italic">a</span> and <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mrow> <mi>e</mi> <mi>u</mi> </mrow> </msub> </mrow> </semantics></math> measured in the Western English Channel and Celtic Sea during four separate cruises (DY026, DY029, JR98, CD173) [<a href="#B72-remotesensing-10-01389" class="html-bibr">72</a>,<a href="#B73-remotesensing-10-01389" class="html-bibr">73</a>]. Equation (3) from the model of Brewin et al. [<a href="#B68-remotesensing-10-01389" class="html-bibr">68</a>] is overlaid (black line) to indicate differences with the measured shelf sea data. r is the correlation coefficient between <math display="inline"><semantics> <mrow> <msub> <mi>Z</mi> <mrow> <mi>e</mi> <mi>u</mi> </mrow> </msub> <mtext> </mtext> </mrow> </semantics></math> estimates from the Brewin et al. [<a href="#B68-remotesensing-10-01389" class="html-bibr">68</a>] model and the data, and RMSE is the root mean square difference in meters, between the model and data.</p>
Full article ">Figure 4
<p>Modelled Chl <span class="html-italic">a</span> calculated using Equation (3) assuming a stratified water column vs. profiles of in situ measurements of chlorophyll from UK shelf seas. Statistical tests were performed in log<sub>10</sub> space.</p>
Full article ">Figure 5
<p>In (<b>a</b>) it shows the relative biomass <math display="inline"><semantics> <mrow> <mrow> <mo>(</mo> <mrow> <msup> <mi>B</mi> <mrow> <mi>B</mi> <mi>s</mi> </mrow> </msup> </mrow> <mo>)</mo> </mrow> </mrow> </semantics></math> with respect to dimensionless depth (ζ) and surface chlorophyll concentration <math display="inline"><semantics> <mrow> <mo stretchy="false">(</mo> <msub> <mi>B</mi> <mi>s</mi> </msub> <mo stretchy="false">)</mo> </mrow> </semantics></math>, and (<b>b</b>) shows the absolute concentration of chlorophyll-<span class="html-italic">a</span> at specific depths across a range of surface chlorophyll values. Note that the x-axis is normalised to surface Chl <span class="html-italic">a</span>.</p>
Full article ">Figure 6
<p>Estimates of total and size-fractionated primary production (PP) using the retuned shelf model run with the monthly composites of European Space Agency (ESA) Ocean Colour Climate Change Intiative (OC-CCI) Chl <span class="html-italic">a</span> product data and NASA MODIS-Aqua photosynthetically available radiation (PAR).</p>
Full article ">Figure 7
<p>Differences in total and &lt;2 µm PP using the Shelf Sat and Open Sat models.</p>
Full article ">Figure 8
<p>Estimates of monthly micro + nanophytoplankton PP using the Shelf Sat model (<b>left panel</b>) and the difference in micro + nanophytoplankton PP estimated by the Shelf Sat and Open Sat models (<b>right panel</b>).</p>
Full article ">Figure 9
<p>Climatological means of measured total and size-fractionated PP for station L4 2009–2015 with standard error and the Shelf Sat model for 2016.</p>
Full article ">Figure 10
<p>Climatological means of picophytoplankton and micro + nanophytoplankton PP for station L4 2009–2015 with standard error plotted with estimates by the Shelf Sat and Open Sat models using OC-CCI v3.1 chlorophyll imagery for 2016.</p>
Full article ">
18 pages, 7129 KiB  
Article
Variations in Remotely-Sensed Phytoplankton Size Structure of a Cyclonic Eddy in the Southwest Indian Ocean
by Tarron Lamont, Raymond G. Barlow and Robert J. W. Brewin
Remote Sens. 2018, 10(7), 1143; https://doi.org/10.3390/rs10071143 - 19 Jul 2018
Cited by 4 | Viewed by 5937
Abstract
Phytoplankton size classes were derived from weekly-averaged MODIS Aqua chlorophyll a data over the southwest Indian Ocean in order to assess changes in surface phytoplankton community structure within a cyclonic eddy as it propagated across the Mozambique Basin in 2013. Satellite altimetry was [...] Read more.
Phytoplankton size classes were derived from weekly-averaged MODIS Aqua chlorophyll a data over the southwest Indian Ocean in order to assess changes in surface phytoplankton community structure within a cyclonic eddy as it propagated across the Mozambique Basin in 2013. Satellite altimetry was used to identify and track the southwesterly movement of the eddy from its origin off Madagascar in mid-June until mid-October, when it eventually merged with the Agulhas Current along the east coast of South Africa. Nano- and picophytoplankton comprised most of the community in the early phase of the eddy development in June, but nanophytoplankton then dominated in austral winter (July and August). Microphytoplankton was entrained into the eddy by horizontal advection from the southern Madagascar shelf, increasing the proportion of microphytoplankton to 23% when the chlorophyll a levels reached a peak of 0.36 mg·m−3 in the third week of July. Chlorophyll a levels declined to <0.2 mg·m−3 in austral spring (September and October) as the eddy propagated further to the southwest. Picophytoplankton dominated the community during the spring period, accounting for >50% of the population. As far as is known, this is the first study to investigate temporal changes in chlorophyll a and community structure in a cyclonic eddy propagating across an ocean basin in the southwest Indian Ocean. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>Main oceanographic features in the Mozambique Channel and Mozambique Basin. The southern branch of the East Madagascar Current (SEMC), the Agulhas Current, Mozambique Channel eddies, as well as dipoles stemming from the SEMC are indicated. Anticlockwise (clockwise) circulation features indicate anticyclonic (cyclonic) eddies. Black contours indicate the 1000 m, 2000 m, 3000 m, 4000 m, and 5000 m bathymetric contours.</p>
Full article ">Figure 2
<p>Verification of the use of the three-component model for studying phytoplankton size structure within the MB eddy. (<b>a</b>–<b>c</b>) show the fractions of micro-, nano-, and picophytoplankton, respectively, as a function of chlorophyll <span class="html-italic">a</span> for in situ measurements collected during the passage of the MB eddy with the three-component model overlain. For comparison, the data from the Lamont et al. [<a href="#B19-remotesensing-10-01143" class="html-bibr">19</a>] (L18) study is also shown, and the grey shading represents uncertainty in the fractions based on the validation in the L18 study (see their <a href="#remotesensing-10-01143-f003" class="html-fig">Figure 3</a>). MAD is the median absolute difference between the model and the in situ size fractions from the MB eddy. (<b>d</b>–<b>g</b>) show the in situ chlorophyll <span class="html-italic">a</span> and size fractions overlain onto MODIS-Aqua estimates from 20 and 24 July 2013, merged (averaged) into a single image. The in situ samples are coloured on the same scale as the satellite images.</p>
Full article ">Figure 3
<p>Sea Surface Height (colour contours) and geostrophic velocity (black arrows) over the Mozambique Basin on 17 June 2013. The black box highlights the location of the cyclonic eddy.</p>
Full article ">Figure 4
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 21 June to 22 July 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 18 June to 27 July 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 5
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 18 June to 27 July 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 6
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 31 July to 24 August 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 28 July to 28 August 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 7
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 28 July to 28 August 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 8
<p>(<b>a</b>–<b>d</b>) Daily Sea Surface Height (colour contours) and geostrophic velocity (black arrows) on selected days for 1 September to 11 October 2013; and (<b>e</b>–<b>h</b>) 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> composites for 29 August to 15 October 2013, over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 9
<p>Fractional contribution of (<b>a</b>–<b>d</b>) micro-, (<b>e</b>–<b>h</b>) nano-, and (<b>i</b>–<b>l</b>) picophytoplankton to MODIS Aqua chlorophyll <span class="html-italic">a</span> for 29 August to 15 October 2013 over the Mozambique Basin. Black boxes highlight the location of the cyclonic eddy and black dots indicate the centre of the eddy. White areas indicate missing data due to cloud cover.</p>
Full article ">Figure 10
<p>Temporal variation in (<b>a</b>) daily Sea Surface Height (SSH) (black line) and 8-day MODIS Aqua chlorophyll <span class="html-italic">a</span> (green line and dots), and (<b>b</b>) the fractional contribution of micro-, nano-, and picophytoplankton at the centre of the cyclonic eddy as it propagated across the Mozambique Basin. Vertical bars indicate the standard deviation of chlorophyll <span class="html-italic">a</span> and the fractional contributions of micro- (green dots and line), nano- (blue dots and line), and picophytoplankton (red dots and line) for the 3 × 3 pixel window at the centre of the eddy.</p>
Full article ">
24 pages, 4032 KiB  
Article
Accuracy Assessment of Primary Production Models with and without Photoinhibition Using Ocean-Colour Climate Change Initiative Data in the North East Atlantic Ocean
by Polina Lobanova, Gavin H. Tilstone, Igor Bashmachnikov and Vanda Brotas
Remote Sens. 2018, 10(7), 1116; https://doi.org/10.3390/rs10071116 - 12 Jul 2018
Cited by 12 | Viewed by 5802
Abstract
The accuracy of three satellite models of primary production (PP) of varying complexity was assessed against 95 in situ 14C uptake measurements from the North East Atlantic Ocean (NEA). The models were run using the European Space Agency (ESA), Ocean Colour Climate [...] Read more.
The accuracy of three satellite models of primary production (PP) of varying complexity was assessed against 95 in situ 14C uptake measurements from the North East Atlantic Ocean (NEA). The models were run using the European Space Agency (ESA), Ocean Colour Climate Change Initiative (OC-CCI) version 3.0 data. The objectives of the study were to determine which is the most accurate PP model for the region in different provinces and seasons, what is the accuracy of the models using both high (daily) and low (eight day) temporal resolution OC-CCI data, and whether the performance of the models is improved by implementing a photoinhibition function? The Platt-Sathyendranath primary production model (PPPSM) was the most accurate over all NEA provinces and, specifically, in the Atlantic Arctic province (ARCT) and North Atlantic Drift (NADR) provinces. The implementation of a photoinhibition function in the PPPSM reduced its accuracy, especially at lower range PP. The Vertical Generalized Production Model-VGPM (PPVGPM) tended to over-estimate PP, especially in summer and in the NADR. The accuracy of PPVGPM improved with the implementation of a photoinhibition function in summer. The absorption model of primary production (PPAph), with and without photoinhibition, was the least accurate model for the NEA. Mapped images of each model showed that the PPVGPM was 150% higher in the NADR compared to PPPSM. In the North Atlantic Subtropical Gyre (NAST) province, PPAph was 355% higher than PPPSM, whereas PPVGPM was 215% higher. A sensitivity analysis indicated that chlorophyll-a (Chl a), or the absorption of phytoplankton, at 443 nm (aph (443)) caused the largest error in the estimation of PP, followed by the photosynthetic rate terms and then the irradiance functions used for each model. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Stations sampled for the determination of in situ integrated daily water column primary production (<span class="html-italic">PP<sub>eu</sub></span>) in the North East Atlantic. Filled circles are stations sampled in summer; open circles are stations sampled in autumn.</p>
Full article ">Figure 2
<p>Photosynthesis-irradiance functions of the models used in the study: Solid lines—without photoinhibition, dotted lines—with photoinhibition.</p>
Full article ">Figure 3
<p>Scatter plots of log satellite and in situ <span class="html-italic">PP<sub>eu</sub></span> [mg C m<sup>−2</sup> day<sup>−1</sup>] in the NEA using daily OC-CCI data (<span class="html-italic">N</span> = 46) for (<b>a</b>) PP<sub>VGPM</sub>, (<b>b</b>) PP<sub>PSM</sub>, and (<b>c</b>) PP<sub>Aph</sub>. Solid line is the 1:1 line; dotted line is the linear regression.</p>
Full article ">Figure 4
<p>Scatter plots of log satellite and in situ <span class="html-italic">PP<sub>eu</sub></span> [mg C m<sup>−2</sup> day<sup>−1</sup>] in the NEA using eight day OC-CCI composites (<span class="html-italic">N</span> = 95); (<b>a</b>) PP<sub>VGPM</sub> with photoinhibition, (<b>b</b>) PP<sub>PSM</sub> with photoinhibition, (<b>c</b>) PP<sub>Aph</sub> with photoinhibition, (<b>d</b>) PP<sub>VGPM</sub> with no photoinhibition, (<b>e</b>) PP<sub>PSM</sub> with no photoinhibition, and (<b>f</b>) PP<sub>Aph</sub> with no photoinhibition. Filled circles are data collected during summer; open squares are autumn data. Solid line is the 1:1 line; dotted line is the linear regression.</p>
Full article ">Figure 5
<p>Taylor diagram of log satellite and in situ <span class="html-italic">PP<sub>eu</sub></span> [mg C m<sup>−2</sup> day<sup>−1</sup>] for the NEA using eight day OC-CCI composites (<span class="html-italic">N</span> = 95). SD—solid arc, centre-pattern RMSE—dotted arc, <span class="html-italic">r</span>—dashed lines with a dot.</p>
Full article ">Figure 6
<p>Spatial distribution of satellite <span class="html-italic">PP<sub>eu</sub></span> [mg C m<sup>−2</sup> day<sup>−1</sup>] using OC-CCI climatology for 1998–2011 for; (<b>a</b>) PP<sub>VGPM</sub> in summer (June–August), (<b>b</b>) PP<sub>PSM</sub> in summer, (<b>c</b>) PP<sub>Aph</sub> in summer, (<b>d</b>) PP<sub>VGPM</sub> in autumn (September–November), (<b>e</b>) PP<sub>PSM</sub> in autumn, and (<b>f</b>) PP<sub>Aph</sub> in autumn.</p>
Full article ">Figure 7
<p>Spatial transects of satellite PP<sub>VGPM</sub>, PP<sub>PSM</sub>, and PP<sub>Aph</sub> [mg C m<sup>−2</sup> day<sup>−1</sup>] along 20° W from 20 to 60° N for (<b>a</b>) summer (June–August) and (<b>b</b>) autumn (September–November) using OC-CCI climatology for 1998–2011.</p>
Full article ">Figure 8
<p>Box-whisker plots for the sensitivity analysis on (<b>a</b>) PP<sub>VGPM</sub>, (<b>b</b>) PP<sub>PSM</sub>, and (<b>c</b>) PP<sub>Aph</sub>: For the left hand y-axis, PP is from 0–4400 mg C m<sup>−2</sup> day<sup>−1</sup>; right y-axis PP is from 0–1200 mg C m<sup>−2</sup> day<sup>−1</sup>. <span class="html-italic">Chl</span><sub>90</sub> is the average Chl <span class="html-italic">a</span> concentration over the first optical depth, <span class="html-italic">k<sub>d</sub></span> is the downwelling diffuse attenuation coefficient, <span class="html-italic">I<sub>0</sub></span> is the daily surface PAR, <span class="html-italic">P<sup>B</sup><sub>opt</sub></span> is the biomass-specific optimum rate of photosynthesis, <span class="html-italic">DL</span> is the day length, <span class="html-italic">P<sup>B</sup><sub>m</sub></span> is the biomass-specific maximum rate of photosynthesis, <span class="html-italic">α<sup>B</sup></span> is the initial slope of the <span class="html-italic">P</span>-<span class="html-italic">I</span> curve, <span class="html-italic">a<sub>ph</sub></span> (443) is the coefficient of light absorption by phytoplankton at 443 nm, <span class="html-italic">φ<sub>m</sub></span> is the maximum quantum yield of photosynthesis, and <span class="html-italic">K<sub>φ</sub></span> is the half-saturation constant of the <span class="html-italic">φ</span>-<span class="html-italic">I</span> curve. The rectangular boundaries are the first and third quartiles (the 25th and 75th percentiles) of the modelled <span class="html-italic">PP<sub>eu</sub></span> obtained by changing each parameter in turn. The line in the rectangle is the median (50th percentile) of the sample, the edges of the “whiskers” are the size of the sample (minimum and maximum of the sample), and the symbol, “o”, represents extreme values. The larger the size of the box and whiskers, the greater the contribution of the parameter to the variability in the modelled <span class="html-italic">PP<sub>eu</sub></span>.</p>
Full article ">Figure 9
<p>Box-and-whisker diagram from the sensitivity analysis for PP<sub>VGPM</sub>, PP<sub>PSM</sub>, and PP<sub>Aph</sub> in the ARCT province. Chl <span class="html-italic">a</span> and <span class="html-italic">a<sub>ph</sub></span> (443) were changed in the models based on the variability in values at stations.</p>
Full article ">Figure 10
<p>Histograms of the coefficient of variation of satellite <span class="html-italic">PP<sub>eu</sub></span> for the area of 12 × 12 km (3 × 3 pixels) for each match-up station analysed using eight day composite data (<span class="html-italic">N</span> = 95) for (<b>a</b>) PP<sub>VGPM</sub>; (<b>b</b>) PP<sub>PSM</sub>; and (<b>c</b>) PP<sub>Aph</sub> and using daily data (<span class="html-italic">N</span> = 46) with (<b>d</b>) PP<sub>VGPM</sub>; (<b>e</b>) PP<sub>PSM</sub>; and (<b>f</b>) PP<sub>Aph</sub>.</p>
Full article ">
30 pages, 7406 KiB  
Article
Bio-Optical Characterization and Ocean Colour Inversion in the Eastern Lagoon of New Caledonia, South Tropical Pacific
by Luciane Rafaele Favareto, Natália Rudorff, Milton Kampel, Robert Frouin, Rüdiger Röttgers, David Doxaran, Hiroshi Murakami and Cécile Dupouy
Remote Sens. 2018, 10(7), 1043; https://doi.org/10.3390/rs10071043 - 2 Jul 2018
Cited by 12 | Viewed by 5579
Abstract
The Eastern Lagoon of New Caledonia (ELNC) is a semi-enclosed system surrounded by an extensive coral reef barrier. The system has been suffering impacts from climate variability and anthropogenic activities, including mining exploitation. Satellite monitoring is thus an essential tool to detect such [...] Read more.
The Eastern Lagoon of New Caledonia (ELNC) is a semi-enclosed system surrounded by an extensive coral reef barrier. The system has been suffering impacts from climate variability and anthropogenic activities, including mining exploitation. Satellite monitoring is thus an essential tool to detect such changes. The present study aimed to assess the bio-optical variability of the ELNC and examine the applicability of ocean colour algorithms, using in situ bio-optical and radiometric data, collected during the March 2014 CALIOPE 2 cruise. The chlorophyll a concentration (Chla) varied from 0.13–0.72 mg·m−3, and the coastal stations were spectrally dominated by non-algal particles (NAP) and coloured dissolved organic matter (CDOM) (>80% of the total non-water absorption at 443 nm), due to the contribution of allochthonous sources. The phytoplankton specific absorption was generally lower (mean, 0.049 m2·mg Chla−1) than typical values observed for the corresponding Chla range, as well as the spectral slopes of the absorption of CDOM plus NAP (adg) (mean, 0.016 nm−1) and of the particle backscattering coefficient (bbp) (mean, 0.07 nm−1). The remote sensing reflectance obtained using two in-water approaches and modelled from Inherent Optical Properties (IOPs) showed less than 20% relative percent differences (RPD). Chla estimates were highly biased for the empirical (OC4 and OC3) and semi-analytical (GSM, QAA, GIOP, LMI) algorithms, especially at the coastal stations. Excluding these stations, the GSM01 yielded the best retrievals with 35–40% RPD. adg(443) was well retrieved by all algorithms with ~18% RPD, and bbp(443) with ~40% RPD. Turbidity algorithms also performed reasonably well (30% RPD), showing the capacity and usefulness of the derived products to monitor the water quality of the ELNC, provided accurate atmospheric correction of the satellite data. Regionally tuned algorithms may potentially improve the Chla retrievals, but better parameterization schemes that consider the spatiotemporal variability of the specific IOPs are still needed. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>Map of the study area indicating the location of the sampling stations from the CALIOPE 2 conducted from 8 to 21 of March 2014 in the Eastern Lagoon of New Caledonia (ELNC). Data from bathymetry, hydrography, villages, and mining sites were downloaded from <a href="http://ftp://ftp.gouv.nc/sig/" target="_blank">ftp://ftp.gouv.nc/sig/</a> accessed on 5 February 2014.</p>
Full article ">Figure 2
<p>Photographs of the Satlantic Hyperpro-II free-falling radiometer (<b>A</b>,<b>B</b>) and the TriOS-RAMSES in-water radiometer attached to a floating polyvinyl chloride (PVC) frame (<b>C</b>,<b>D</b>).</p>
Full article ">Figure 3
<p>Surface distribution maps of the surface water salinity (<b>A</b>) and turbidity (<b>B</b>) in the ELNC, during CALIOPE 2 (March 2014) cruises.</p>
Full article ">Figure 4
<p>Phytoplankton absorption (<span class="html-italic">a</span><sub>phy</sub>) spectra (<b>A</b>) and specific phytoplankton absorption coefficient (<span class="html-italic">a</span><sub>phy</sub>*) at 443 nm versus the TChl<span class="html-italic">a</span>, with the power law regression fit (solid line) and the Bricaud et al. [<a href="#B23-remotesensing-10-01043" class="html-bibr">23</a>] fit for reference (dashed line) (<b>B</b>) (Number of samples = 52).</p>
Full article ">Figure 5
<p>Coloured dissolved organic matter (CDOM) absorption (<span class="html-italic">a</span><sub>cdom</sub>) spectra (<b>A</b>) and non-algal particles (NAP) absorption (<span class="html-italic">a</span><sub>nap</sub>) spectra for each station, mean and standard deviation for all stations (<b>B</b>) (Number of samples = 52).</p>
Full article ">Figure 6
<p>Mean (solid lines) and standard deviation (dashed lines) of phytoplankton absorption (<span class="html-italic">a</span><sub>phy</sub>), Coloured dissolved organic matter (CDOM) absorption (<span class="html-italic">a</span><sub>cdom</sub>) and non-algal particles (NAP) absorption (<span class="html-italic">a</span><sub>nap</sub>) (<b>A</b>). Ternary diagram with the proportions of <span class="html-italic">a</span><sub>phy</sub>(443), <span class="html-italic">a</span><sub>cdom</sub>(443) and <span class="html-italic">a</span><sub>nap</sub>(443) for each station (Number of samples = 51) colour coded by the bottom depth (m) (<b>B</b>).</p>
Full article ">Figure 7
<p>Particulate backscattering coefficient (<span class="html-italic">b</span><sub>bp</sub>) spectra for each station, mean and standard deviation (Number of samples = 52) (<b>A</b>). The <span class="html-italic">b</span><sub>bp</sub> at 555 nm versus phytoplankton absorption (<span class="html-italic">a</span><sub>phy</sub>) at 443 nm, Coloured dissolved organic matter (CDOM) absorption (<span class="html-italic">a</span><sub>cdom</sub>) at 443 nm and non-algal particles (NAP) absorption (<span class="html-italic">a</span><sub>nap</sub>) at 443 nm with their mean square error, coefficient of determination (R<sup>2</sup>) and the power law fit for the 51 stations sampled in ELNC (<b>B</b>) (axes in log10).</p>
Full article ">Figure 8
<p>In situ remote sensing reflectance (<span class="html-italic">R</span><sub>rs</sub>) spectra obtained with the TriOS sensor (<b>A</b>) and the surface distribution map of <span class="html-italic">R</span><sub>rs</sub>(555) (<b>B</b>) (Number of samples = 48).</p>
Full article ">Figure 9
<p>Comparison of Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) vs. Remote sensing reflectance from Satlantic radiometer (<span class="html-italic">R</span><sub>rs</sub>S) for the ocean colour bands and the 443:555 ratio (Number of samples = 34).</p>
Full article ">Figure 10
<p>Comparison of the modelled Remote sensing reflectance (<span class="html-italic">R</span><sub>rs</sub>) of the radiative-transfer equation (RTE) following Park and Ruddick [<a href="#B66-remotesensing-10-01043" class="html-bibr">66</a>] and measured Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) for each band and for the 443:555 band ratio (Number of samples = 48).</p>
Full article ">Figure 11
<p>Comparisons between the in situ measured and modelled chlorophyll <span class="html-italic">a</span> concentration (Chl<span class="html-italic">a</span>) using Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) (TriOS Chl<span class="html-italic">a</span>), determined with the: OC3M (<b>A</b>), OC4 (<b>B</b>), Garver-Siegel-Maritorena (GSM01) (<b>C</b>) and Generalized IOP (GIOP) (<b>D</b>) (Number of samples = 48). Axes in log scale.</p>
Full article ">Figure 12
<p>Comparisons between the in situ measured and modelled phytoplankton absorption (<span class="html-italic">a</span><sub>phy</sub>) at 443 nm using Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) (TriOS <span class="html-italic">a</span><sub>phy</sub>) obtained with: Garver-Siegel-Maritorena (GSM01) (<b>A</b>), Quasi-Analytical Algorithm (QAA) (<b>B</b>), Generalized IOP (GIOP) (<b>C</b>) and Linear Matrix Inversion (LMI) (<b>D</b>) (Number of samples = 47). Axes in log scale.</p>
Full article ">Figure 13
<p>Comparisons between the in situ measured and modelled absorption of Coloured dissolved organic matter (CDOM) plus non-algal particles (NAP), the <span class="html-italic">a</span><sub>dg</sub> at 443 nm using Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) (TriOS <span class="html-italic">a</span><sub>dg</sub>) obtained with: Garver-Siegel-Maritorena (GSM01) (<b>A</b>), Quasi-Analytical Algorithm (QAA) (<b>B</b>), Generalized IOP (GIOP) (<b>C</b>) and Linear Matrix Inversion (LMI) (<b>D</b>) (Number of samples = 47). Axes in log scale.</p>
Full article ">Figure 14
<p>Comparisons between the in situ measured and modelled particle backscattering coefficient (<span class="html-italic">b</span><sub>bp</sub>) at 443 nm using Remote sensing reflectance from TriOS radiometer (<span class="html-italic">R</span><sub>rs</sub>T) (TriOS <span class="html-italic">b</span><sub>bp</sub>) obtained with: Garver-Siegel-Maritorena (GSM01) (<b>A</b>), Quasi-Analytical Algorithm (QAA) (<b>B</b>), Generalized IOP (GIOP) (<b>C</b>) and Linear Matrix Inversion (LMI) (<b>D</b>) (Number of samples = 47). Axes in log scale.</p>
Full article ">Figure 15
<p>Turbidity estimated following Ouillon et al. [<a href="#B38-remotesensing-10-01043" class="html-bibr">38</a>] for New Caledonia (<b>A</b>) and “global” tropical coastal waters (<b>B</b>) and Dogliotti et al. [<a href="#B39-remotesensing-10-01043" class="html-bibr">39</a>] (<b>C</b>), compared to turbidity measured in situ (Number of samples = 48).</p>
Full article ">
25 pages, 11100 KiB  
Article
Remote Sensing of Phytoplankton Size Class in Northwest Atlantic from 1998 to 2016: Bio-Optical Algorithms Comparison and Application
by Xiaohan Liu, Emmanuel Devred and Catherine Johnson
Remote Sens. 2018, 10(7), 1028; https://doi.org/10.3390/rs10071028 - 28 Jun 2018
Cited by 10 | Viewed by 4962
Abstract
Phytoplankton community structure and phytoplankton size class (PSC) are linked to ecological and biogeochemical changes in the oceanic environment. Many models developed to obtain the fraction of PSCs from satellite remote sensing have only been evaluated in open oceans, and very limited effort [...] Read more.
Phytoplankton community structure and phytoplankton size class (PSC) are linked to ecological and biogeochemical changes in the oceanic environment. Many models developed to obtain the fraction of PSCs from satellite remote sensing have only been evaluated in open oceans, and very limited effort has been carried out to report on the performance of these PSC models in productive continental shelf waters. In this study, we evaluated the performance of nine PSC models in the coastal Northwest Atlantic (NWA) by comparison of in situ phytoplankton pigment measurements with coincidental satellite data from the Sea-Viewing Wide Field-of-View Sensor (SeaWiFS), Moderate-resolution Imaging Spectroradiometer (MODIS), and the Visible Infrared Imaging Radiometer Suite (VIIRS). Our results show that no PSC model retrieved all three phytoplankton size classes (pico-, nano-, and micro-phytoplankton) with reliable accuracy in the region of interest. In particular, these PSC models showed poor performance for retrieval of the picophytoplankton fraction of total phytoplankton in our study region, which could be related to the under-representation of pico-dominated samples in the productive waters of the NWA. For the accuracy of retrieved microphytoplankton and combined nano–pico phytoplankton fraction, the regional model developed by Devred et al. (2011) yielded the best result, followed by the model of Brewin et al. (2011). The model of Devred et al. (2011) was applied to satellite-derived chlorophyll-a concentration from the Ocean Color Climate Change Initiative (OC-CCI) archive in the NWA from 1998 to 2016. We report solely on the microphytoplankton biomass and fraction given the inverse relationship that exists with the nano–pico class. The multi-decadal trend along with the deseasonalized trend of microphytoplankton fraction was computed and analyzed for six biogeochemical provinces located in the NWA. Over the 19-year time series, there were significant, positive trends for four of the six provinces, with a slope of 0.36%·yr−1 in the Northwest Continental Shelf (NWCS), 0.25%·yr−1 in the Arctic Waters (ARCT), 0.12%·yr−1 in the Slope Waters (SW) and 0.06%·yr−1 in the Gulf Stream (GFST). Strong positive anomalies of microphytoplankton fraction were found in winter months in NWCS between 2009 and 2014, which could be associated with changes in environmental factors. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Location of (<b>a</b>) in situ and satellite matchups of Chl<span class="html-italic">a</span>; (<b>b</b>) in situ and satellite matchups of phytoplankton absorption, and the dynamic assignment of ecological provinces in the NWA [<a href="#B34-remotesensing-10-01028" class="html-bibr">34</a>].</p>
Full article ">Figure 2
<p>Seasonal distribution of in situ and satellite matchups.</p>
Full article ">Figure 3
<p>Frequency distribution of in situ Chl<span class="html-italic">a</span> for matched (<b>a</b>) SeaWiFS; (<b>b</b>) MODIS; and (<b>c</b>) VIIRS, respectively. Frequency distribution of Chl<span class="html-italic">a</span> for NOMAD dataset is also presented for reference.</p>
Full article ">Figure 4
<p>Flow chart of the evaluation method.</p>
Full article ">Figure 5
<p>Performing accuracies of different models for (<b>a</b>) micro-, (<b>b</b>) mixed nano–pico, (<b>c</b>) nano-, and (<b>d</b>) pico-phytoplankton. The error bars represent the 95% confidence intervals.</p>
Full article ">Figure 6
<p>Percentage variation of micro-, mixed nano–pico, nano- and pico- phytoplankton as a function of Chl<span class="html-italic">a</span> (ranging within a range of ±30%) for models (<b>a</b>) D, (<b>b</b>) E and (<b>c</b>) H.</p>
Full article ">Figure 7
<p>Histograms of percentage error of (<b>a</b>) SeaWiFS; (<b>b</b>) MODIS; (<b>c</b>) VIIRS OCx Chla products, compared to in situ measured Chl<span class="html-italic">a</span>.</p>
Full article ">Figure 8
<p>Histograms of percentage error of satellite derived <span class="html-italic">a</span><sub>ph</sub>(λ) (λ = 443, 490, and 555 nm) for SeaWiFS, MODIS and VIIRS compared to measured <span class="html-italic">a</span><sub>ph</sub>. The <span class="html-italic">a</span><sub>ph</sub> (λ) derived by IOP model of EOF_NWA and QAA v5 was shown in black and blue bars, respectively.</p>
Full article ">Figure 9
<p>Time-series of microphytoplankton in ecological provinces of (<b>a</b>) Gulf Stream; (<b>b</b>) Slope Waters; (<b>c</b>) North Atlantic Drift; (<b>d</b>) Northwest Continental Shelf; (<b>e</b>) Arctic Waters; and (<b>f</b>) Polar Boreal Current. The deseasonalized trend and smoothed trend are also shown for each province.</p>
Full article ">Figure 10
<p>Climatological variation of microphytoplankton in the six ecological provinces.</p>
Full article ">Figure 11
<p>Linear regression of “smoothed deseasonalized trend” in the past 19 years (1998–2016, thick black line), 2000s (1998–2010, magenta line) and 2010s (2011–2016, blue line) in ecological provinces of (<b>a</b>) Gulf Stream; (<b>b</b>) Slope Waters; (<b>c</b>) North Atlantic Drift; (<b>d</b>) Northwest Continental Shelf; (<b>e</b>) Arctic Waters; and (<b>f</b>) Polar Boreal Current. Determination coefficient with “**” denotes <span class="html-italic">p</span> &lt; 0.001.</p>
Full article ">Figure 12
<p>Monthly anomaly heatmaps for fractions of microphytoplankton in (<b>a</b>) Gulf Stream; (<b>b</b>) Slope Waters; (<b>c</b>) North Atlantic Drift; (<b>d</b>) Northwest Continental Shelf; (<b>e</b>) Arctic Waters; and (<b>f</b>) Polar Boreal Current. Values in each cell are anomalies from the mean of each month in the whole study period (1998–2016). Cells in red indicate higher than normal levels; Cells in blue indicate lower than normal levels. A blank cell indicates low valid pixel percentage (&lt;60%) due to ice or cloud cover.</p>
Full article ">Figure 13
<p>Comparison of CCI-derived (<b>a</b>) microphytoplankton concentration; and (<b>b</b>) microphytoplankton fraction by model E [<a href="#B25-remotesensing-10-01028" class="html-bibr">25</a>] with in situ measurements.</p>
Full article ">
26 pages, 16312 KiB  
Article
Using 250-M Surface Reflectance MODIS Aqua/Terra Product to Estimate Turbidity in a Macro-Tidal Harbour: Darwin Harbour, Australia
by Gang Yang, Xiaohua Wang, Elizabeth A. Ritchie, Lulu Qiao, Guangxue Li and Zhixin Cheng
Remote Sens. 2018, 10(7), 997; https://doi.org/10.3390/rs10070997 - 22 Jun 2018
Cited by 12 | Viewed by 6594
Abstract
Turbidity is an indicator of the quality of water and usually exhibits variability associated with changing hydrodynamic conditions, which can be reflected in the sediment dynamics in coastal regions. Darwin Harbour is a typical macro-tidal, well mixed, and complex environment influenced by industries, [...] Read more.
Turbidity is an indicator of the quality of water and usually exhibits variability associated with changing hydrodynamic conditions, which can be reflected in the sediment dynamics in coastal regions. Darwin Harbour is a typical macro-tidal, well mixed, and complex environment influenced by industries, human activities, and natural factors—including winds, currents, river discharges, waves, and tides. As a case study, hydrodynamics and sediment dynamics in Darwin Harbour are investigated using moderate resolution imaging spectroradiometer (MODIS) measurements. This study focuses on understanding the variability of turbidity, mechanisms that control the variations of turbidity and analyzing field data to determine the main factors that influence the sediment dynamics in Darwin Harbour. The results of this study illustrate the seasonal turbidity variation is mainly influenced by the wind waves. The dredging campaigns in 2013 and 2014 wet seasons contributed to the rise of turbidity in Darwin Harbour. The action of tidal currents appears to be the dominant factor controlling the turbidity pattern in a spring–neap cycle and the turbidity intra-tidal variation. In addition, the turbidity maximum zone (TMZ) near Charles Point is formed by the tidal current convergence based on the results of current modelling. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of Darwin Harbour and the adjacent coastal region. The red dot shows the location of the Integrated Marine Observing System station (IMOS). Blue dots show the positions of selected stations A and B. The line between P1 and P2 shows the location of transect 1. The red line between Mandorah Point and East Point is the boundary between the outer and inner harbour.</p>
Full article ">Figure 2
<p>(<b>a</b>) The calibrated SASM (Equations (2) and (3)) using regression analysis between in situ turbidity and x; (<b>b</b>) comparisons between measured turbidity at the IMOS station and the MODIS retrieved surface turbidity.</p>
Full article ">Figure 3
<p>MODIS retrieved seasonally averaged turbidity for the low water spring tide in Darwin Harbour for the wet season (<b>left</b>) and dry season (<b>right</b>) during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The dredging offshore disposal area is shown by the red box.</p>
Full article ">Figure 3 Cont.
<p>MODIS retrieved seasonally averaged turbidity for the low water spring tide in Darwin Harbour for the wet season (<b>left</b>) and dry season (<b>right</b>) during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The dredging offshore disposal area is shown by the red box.</p>
Full article ">Figure 4
<p>The turbidity with standard deviation in the wet and dry seasons at (<b>a</b>) A station; and (<b>b</b>) B station.</p>
Full article ">Figure 5
<p>Mean seasonal turbidity along transect 1 (~36 km) in: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The standard deviation is indicated by the error bars.</p>
Full article ">Figure 5 Cont.
<p>Mean seasonal turbidity along transect 1 (~36 km) in: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The standard deviation is indicated by the error bars.</p>
Full article ">Figure 6
<p>Monthly river discharge of Blackmore River (Station G8150098), Berry Creek (Station 8150028), and Elizabeth River (Station G8150018).</p>
Full article ">Figure 7
<p>Wind speed and direction roses (e.g., direction with a value of 315 means wind comes from the northwest) during the wet season (<b>left</b>) and dry season (<b>right</b>) in: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017.</p>
Full article ">Figure 7 Cont.
<p>Wind speed and direction roses (e.g., direction with a value of 315 means wind comes from the northwest) during the wet season (<b>left</b>) and dry season (<b>right</b>) in: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017.</p>
Full article ">Figure 8
<p>The relationship between the MODIS-Aqua retrieved turbidity at the IMOS station during the low water spring tide and corresponding significant waves height and wind speed for cloudless days during the observation period from 2013 to 2017.</p>
Full article ">Figure 9
<p>MODIS retrieved averaged turbidity for low water spring (left) and neap (right) tide in the dry season during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The location of the TMZ formed by eddies is shown by the red box.</p>
Full article ">Figure 9 Cont.
<p>MODIS retrieved averaged turbidity for low water spring (left) and neap (right) tide in the dry season during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The location of the TMZ formed by eddies is shown by the red box.</p>
Full article ">Figure 10
<p>Mean retrieved turbidity with standard deviation along transect 1 at low water spring and neap tide in the dry season during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017.</p>
Full article ">Figure 11
<p>The relationship between retrieved turbidity at low water level and the tidal range at Station A.</p>
Full article ">Figure 12
<p>MODIS-retrieved averaged turbidity during spring tide in the dry season during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017.</p>
Full article ">Figure 13
<p>Time averaged retrieved turbidity for each phase along transect 1 during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017.</p>
Full article ">Figure 14
<p>Turbidity at stations A and B for each phase during the spring tide during: (<b>a</b>) 2013; (<b>b</b>) 2014; (<b>c</b>) 2015; (<b>d</b>) 2016; and (<b>e</b>) 2017. The standard deviation is indicated by the error bars.</p>
Full article ">Figure 15
<p>Location of the TMZ near Charles Point (Area A) and the TMZ in the northwest of Charles Point (Area B) during the four tidal phases (<b>a</b>) ebb tide; (<b>b</b>) low water level; (<b>c</b>) flood tide; and (<b>d</b>) high water level during the 2014 dry season.</p>
Full article ">Figure 15 Cont.
<p>Location of the TMZ near Charles Point (Area A) and the TMZ in the northwest of Charles Point (Area B) during the four tidal phases (<b>a</b>) ebb tide; (<b>b</b>) low water level; (<b>c</b>) flood tide; and (<b>d</b>) high water level during the 2014 dry season.</p>
Full article ">Figure 16
<p>The 2D current model domain and open boundary for the water elevation boundary (red dash line).</p>
Full article ">Figure 17
<p>Tidal current convergence and divergence (−<math display="inline"> <semantics> <mrow> <mo>????</mo> <mo>⋅</mo> <mi>V</mi> <mo>=</mo> <mfrac> <mrow> <mo>∂</mo> <mi>u</mi> </mrow> <mrow> <mo>∂</mo> <mi>x</mi> </mrow> </mfrac> <mo>+</mo> <mfrac> <mrow> <mo>∂</mo> <mi>ν</mi> </mrow> <mrow> <mo>∂</mo> <mi>y</mi> </mrow> </mfrac> </mrow> </semantics> </math>, positive value: convergence, negative value: divergence) at flood and ebb tide (05:30 p.m. on 28 May, 11:30 p.m. on 28 May 2013), respectively.</p>
Full article ">Figure 18
<p>The Delft 3D modelling tidal depth averaged current velocity and direction with current data from IMOS station from 00:00 a.m. 28 May to 10:00 p.m. 30 May 2013.</p>
Full article ">
24 pages, 6646 KiB  
Article
Evaluation of MODIS—Aqua Chlorophyll-a Algorithms in the Basilicata Ionian Coastal Waters
by Teodosio Lacava, Emanuele Ciancia, Carmine Di Polito, Alice Madonia, Simone Pascucci, Nicola Pergola, Viviana Piermattei, Valeria Satriano and Valerio Tramutoli
Remote Sens. 2018, 10(7), 987; https://doi.org/10.3390/rs10070987 - 21 Jun 2018
Cited by 12 | Viewed by 5229
Abstract
Standard chlorophyll-a (chl-a) algorithms, which rely on Moderate Resolution Imaging Spectro-radiometer (MODIS) data aboard the Aqua satellite, usually show different performances depending on the area under consideration. In this paper, we assessed their accuracy in retrieving the chl-a concentration in the Basilicata Ionian [...] Read more.
Standard chlorophyll-a (chl-a) algorithms, which rely on Moderate Resolution Imaging Spectro-radiometer (MODIS) data aboard the Aqua satellite, usually show different performances depending on the area under consideration. In this paper, we assessed their accuracy in retrieving the chl-a concentration in the Basilicata Ionian Coastal waters (Ionian Sea, South of Italy). The outputs of one empirical (Med-OC3) and two semi-analytical algorithms, the Garver–Siegel–Maritorena (GSM) and the Generalized Inherent Optical Properties (GIOP) model, have been compared with ground measurements acquired during three different measurement campaigns. The achieved results prove the poor accuracy (adjusted R2 value of 0.12) of the investigated empirical algorithm and, conversely, the good performance of semi-analytical algorithms (adjusted R2 ranging from 0.74 to 0.79). The co-existence of Coloured Dissolved Organic Matter (CDOM) and Non-Algal Particles (NAP) has likely determined large errors in the reflectance ratios used in the OCx form algorithms. Finally, a local scale assessment of the bio-optical properties, on the basis of the in situ dataset, allowed for the definition of an operational local scale-tuned version of the MODIS chl-a algorithm, which assured increased accuracy (adjusted R2 value of 0.86). Such a tuned algorithm version can provide useful information which can be used by local authorities within regional management systems. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>(<b>a</b>) BICW localization (red box) in the Mediterranean Sea; (<b>b</b>) magnification of the study area within the red box of (a) where the Basilicata region is depicted in orange. The main rivers of the region (blue lines) and the bathymetry from 50 to 500 m (in blue tones and with the black contour lines) are also reported.</p>
Full article ">Figure 2
<p>Sampling locations for each of the IOSMOS measurement campaigns.</p>
Full article ">Figure 3
<p>Daily chl-a maps derived from MODIS-Aqua using (<b>a</b>) Med-OC3, (<b>b</b>) GSM, (<b>c</b>) GIOP for 19 April 2013 (top panel) and 1 July 2014 (bottom panel).</p>
Full article ">Figure 4
<p>Scatter plots of in situ versus MODIS-A chl-a values for Med-OC3, GSM, and GIOP algorithms. In each plot, the dashed line is the 1:1 line.</p>
Full article ">Figure 5
<p>In situ R<sub>rs</sub>(λ) spectra acquired on 26 IOSMOS stations. Within the legend, two numbers characterize each spectrum: the first is the identification of the stations and the second (in parentheses) relates to the measurement campaigns (from 1 to 3).</p>
Full article ">Figure 6
<p>The flowchart of the revised max-classification performed over 26 R<sub>rs</sub>(λ) spectra.</p>
Full article ">Figure 7
<p>(<b>a</b>) The normalized average spectral shapes (R<sub>rs</sub>(λ)/max(R<sub>rs</sub>(λ)) for the classes A and B. (<b>b</b>) Locations of the analysed stations falling within the classes A (red dots) and B (green dots). The continuous lines are bathymetry depths ranging from 50 to 500 m depth.</p>
Full article ">Figure 8
<p>Plot of the averaged a<sub>CDM</sub>(λ) spectrum in the 350–500 nm domain. The dotted red line is the exponential best fit.</p>
Full article ">Figure 9
<p>Plot of the power-law relationship between η and the averaged ratio r<sub>rs</sub>(443)/r<sub>rs</sub>(555). The red square indicates the selected η value.</p>
Full article ">Figure 10
<p>Plot of comparison between the averaged a*<sub>ph</sub>(λ) values computed within the locale scale configuration (hereafter GSM-BICW, red line) and the standard ones related to the default configuration (GSM—standard, blue line). All the a*<sub>ph</sub>(λ) values were normalized at 443 nm to facilitate the comparison.</p>
Full article ">Figure 11
<p>Scatter plots of in situ versus MODIS-A chl-a values for GSM-standard and GSM-BICW algorithms. In each plot, the dashed line is the 1:1 line.</p>
Full article ">Figure 12
<p>Scatter plot between in situ and satellite chl-a data. The blue rhombus are chl-a values obtained by GSM-BICW algorithm, while the red circles are related to the CMEMS-L3 product. The dashed line is the 1:1 line.</p>
Full article ">Figure 13
<p>Comparison between the water type masks (averaged value on a 3 × 3-pixel box centered on the location of the in situ chl-a samples) included in the CMEMS-L3 product and the classes derived by the R<sub>rs</sub>(λ) supervised classification implemented on the chl-a samples selected for the match-up analysis (<a href="#remotesensing-10-00987-f012" class="html-fig">Figure 12</a>). The full circles represent the classes (red and green for classes A and B), while the empty circles the water types (red and green for Case 1 and Case 2 waters). The continuous lines are bathymetry depths ranging from 50 to 500 m depth.</p>
Full article ">
23 pages, 6257 KiB  
Article
Evaluation of the First Year of Operational Sentinel-2A Data for Retrieval of Suspended Solids in Medium- to High-Turbidity Waters
by Isabel Caballero, François Steinmetz and Gabriel Navarro
Remote Sens. 2018, 10(7), 982; https://doi.org/10.3390/rs10070982 - 21 Jun 2018
Cited by 65 | Viewed by 7988
Abstract
In this study, we apply high-resolution Sentinel-2A imagery to assist in the monitoring of the southwestern Spanish coast during its first year of data. The aim is to test suitability of MultiSpectral Imager (MSI) at higher resolution (10 m) for mapping Total Suspended [...] Read more.
In this study, we apply high-resolution Sentinel-2A imagery to assist in the monitoring of the southwestern Spanish coast during its first year of data. The aim is to test suitability of MultiSpectral Imager (MSI) at higher resolution (10 m) for mapping Total Suspended Solids (TSS). Several field campaigns are carried out to collect TSS at three different sites in the Guadalquivir estuary, Cadiz Bay and Conil port. A regional multi-conditional remote sensing algorithm with a switching method that automatically selects the most sensitive TSS vs. water reflectance relationship is developed to estimate TSS concentration while avoiding saturation effects. An existing semi-analytical algorithm is calibrated by means of a cross-validation procedure based on both red 664 nm (r = 0.8, NRMSE of 25.06%) and near-infrared (NIR) 865 nm (r = 0.98, NRMSE of 10.28%) parts of the spectrum, showing the MSI sensor’s great potential to estimate TSS even though it was not designed for aquatic remote sensing. The first year of data reveals improved monitoring along the coastal region at unprecedented resolution with accuracy to detect the Estuarine Turbidity Maximum (ETM). ACOLITE and POLYMER Atmospheric Correction strategies are applied over this coastal region (no in-situ data on water reflectance). The results confirm that the flexible POLYMER algorithm can address intense sun-glint effects. These findings encourage further research of water quality studies relying on both operational Sentinel-2A and Sentinel-2B, with great implications to improve the understanding of turbid coastal and inland water environments. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location of the study area (SW Iberian Peninsula) showing the Strait of Gibraltar and the Gulf of Cadiz. Blue and red rectangles delimit the ROI (Region of Interest) in the Guadalquivir estuary and Cadiz Bay, respectively; (<b>b</b>) In-situ sampling data set corresponding to match-ups in each site is indicated for the Guadalquivir estuary (blue dots, 16 June and 13 December 2016), Cadiz Bay (red dots, 6 June 2016) and Conil port (pink dots, 27 May 2016). Blue (Guadalquivir estuary) and red (Cadiz Bay) stars correspond to the location of the 5 × 5 pixels used for the Atmospheric Correction comparison ranging from offshore estuary at Box 1 (36.696°N, 6.676°W) to estuarine or turbid waters at Box 2 (36.773°N, 6.438°W) and Box 3 (36.811°N, 6.343°W), and from offshore bay at Box 1 (36.571°N, 6.38°W) to turbid waters at Box 2 (36.55°N, 6.253°W) and Box 3 (36.503°N, 6.203°W).</p>
Full article ">Figure 2
<p>Schematic flow chart of the satellite image processing system describing the Total Suspended Solids (TSS) sampling, the Atmospheric Correction (AC), the multi-conditional algorithm using the red and near-infrared (NIR) bands, and the cross-validation procedure.</p>
Full article ">Figure 3
<p>(<b>a</b>) Scatter plot showing the comparison between in-situ Total Suspended Solids (TSS) concentration (mg/L) and water reflectance (ρw) of Sentinel-2A during the four field campaigns conducted during 2016 in the Guadalquivir estuary (blue) and Cádiz Bay and Conil port (red) for the red band (664 nm), (<b>b</b>) the same for near-infrared (NIR, 865 nm) band.</p>
Full article ">Figure 4
<p>Relationship for Total Suspended Solids (TSS) retrieval in the Guadalquivir estuary based on the recalibrated reflectance model [<a href="#B29-remotesensing-10-00982" class="html-bibr">29</a>] using the red band (B4) at 664 nm (blue line, logarithmic scales). Black dashed line represents the original fit at this wavelength. Linear fit obtained in Cadiz Bay and Conil port with the red band is presented (red line).</p>
Full article ">Figure 5
<p>Relationship for Total Suspended Solids (TSS) retrieval in the Guadalquivir estuary based on the recalibrated reflectance model [<a href="#B29-remotesensing-10-00982" class="html-bibr">29</a>] using the near-infrared (NIR) band at 865 nm (blue line, logarithmic scales). Black dashed line represents the original fit at this wavelength.</p>
Full article ">Figure 6
<p>Scatter plot between water reflectance (ρw) in the red (664 nm) and near-infrared (NIR, 865 nm) bands extracted for four images on 8 March, 27 May, and 6 and 16 June 2016 in the Guadalquivir estuary and Cadiz Bay Region of Interest (ROI) (<a href="#remotesensing-10-00982-f001" class="html-fig">Figure 1</a>). The solid black line corresponds to the logarithmic regression fit. The 95% confidence intervals for the regression are represented as red lines. The switching point S is indicated at the intersection of the tangent (blue dashed line) with the y axis.</p>
Full article ">Figure 7
<p>Mean water reflectance spectra (ρw) extracted from the 5 × 5 pixels in the Guadalquivir estuary and Cadiz Bay after correction with ACOLITE (cyan and magenta lines) and POLYMER (blue and red lines). The location of the boxes is shown in <a href="#remotesensing-10-00982-f001" class="html-fig">Figure 1</a> ranging from offshore waters (Box 1) to estuarine or turbid waters (Box 2 and Box 3). The scenes corresponded to 8 March 2016 (<b>a</b>–<b>f</b>) and 24 September 2016 (<b>g</b>–<b>l</b>). All y-axes are the same except (c).</p>
Full article ">Figure 8
<p>Scatter plots showing the comparison between water reflectance (ρw) derived from the 8 March 2016 and 24 September 2016 images over the Andalusian coastal waters within the Guadalquivir (<b>a</b>–<b>h</b>) and Cadiz Region of Interest (ROI) (<b>i</b>–<b>p</b>) with POLYMER and ACOLITE for the visible (B1–B7, 444–783 nm) and near-infrared (NIR) bands (B8, 865 nm). The blue line is the fitted line. The dashed black line is the 1:1 line.</p>
Full article ">Figure 8 Cont.
<p>Scatter plots showing the comparison between water reflectance (ρw) derived from the 8 March 2016 and 24 September 2016 images over the Andalusian coastal waters within the Guadalquivir (<b>a</b>–<b>h</b>) and Cadiz Region of Interest (ROI) (<b>i</b>–<b>p</b>) with POLYMER and ACOLITE for the visible (B1–B7, 444–783 nm) and near-infrared (NIR) bands (B8, 865 nm). The blue line is the fitted line. The dashed black line is the 1:1 line.</p>
Full article ">Figure 9
<p>Scatter plots showing the comparison between water reflectance (ρw) derived from the 13 December 2016 image over the Andalusian coastal waters within the Guadalquivir (<b>a</b>–<b>h</b>) and Cadiz Region of Interest (ROI) (<b>i</b>–<b>p</b>) with POLYMER and ACOLITE for the visible (B1–B7, 444–783 nm) and near-infrared (NIR) bands (B8, 865 nm). The blue line is the fitted line. The dashed black line is the 1:1 line.</p>
Full article ">Figure 10
<p>Total Suspended Solids (TSS) concentration (mg/L) derived from the multi-conditional algorithm for Sentinel-2 imagery using the red (664 nm) and near-infrared NIR (865 nm) bands after correction with ACOLITE (10 m spatial resolution) in the Guadalquivir estuary on 19 December 2015 (<b>a</b>–<b>d</b>) and on 17 February 2016 (<b>e</b>–<b>h</b>). Note different ranges in the color bar (TSS concentration).</p>
Full article ">Figure 11
<p>Sentinel-2 derived Total Suspended Solids (TSS, mg/L) concentration in the longitudinal transect of the first 47 km of the Guadalquivir estuary (24 September 2016–1 February 2017). All the images were corrected with ACOLITE. Red line on the map indicates the start at the mouth and the end upstream of the longitudinal transect. The meander with a sinuous curve located at Brazo de la Torre is indicated on the map with the red arrow.</p>
Full article ">Figure 12
<p>Total Suspended Solids (TSS) concentration (mg/L) derived from the multi-conditional algorithm for Sentinel-2 imagery using the red (664 nm) and near-infrared (NIR, 865 nm) bands after correction with ACOLITE (10 m spatial resolution) in Cadiz Bay on 19 December 2015 (<b>a</b>–<b>d</b>) and on 17 February 2016 (<b>e</b>–<b>h</b>).</p>
Full article ">
18 pages, 9846 KiB  
Article
Parameterization of Spectral Particulate and Phytoplankton Absorption Coefficients in Sognefjord and Trondheimsfjord, Two Contrasting Norwegian Fjord Ecosystems
by Veloisa J. Mascarenhas and Oliver Zielinski
Remote Sens. 2018, 10(6), 977; https://doi.org/10.3390/rs10060977 - 20 Jun 2018
Cited by 4 | Viewed by 4748
Abstract
We present here parameterizations of particulate and phytoplankton absorption coefficients as functions of pigment concentrations (Tchla) in Sognefjord and Trondheimsfjord along the northwestern coast of Norway. The total particulate and non-algal optical densities were measured via quantitative filter technique (QFT) in a spectrophotometer [...] Read more.
We present here parameterizations of particulate and phytoplankton absorption coefficients as functions of pigment concentrations (Tchla) in Sognefjord and Trondheimsfjord along the northwestern coast of Norway. The total particulate and non-algal optical densities were measured via quantitative filter technique (QFT) in a spectrophotometer with integrating sphere. The spectral parameter coefficients A(λ) and E(λ) of the power law describing variations of particulate and phytoplankton absorption coefficients as a function of Tchla, were not only different from those provided for open ocean case 1 waters, but also exhibited differences in the two fjords under investigation. Considering the influence of glacial meltwater leading to increased inorganic sediment load in Sognefjord we investigate differences in two different parameterizations, developed by excluding and including inner Sognefjord stations. Tchla are modelled to test the parameterizations and validated against data from the same cruise and that from a repeated campaign. Being less influenced by non-algal particles parameterizations performed well in Trondheimsfjord and yielded high coefficients of determination (R2) of modelled vs. measured Tchla. In Sognefjord, the modelled vs. measured Tchla resulted in better R2 with parameter coefficients developed excluding the inner-fjord stations influenced by glacial meltwater influx. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Sognefjord and Trondheimsfjord, along the northwestern coast of Norway. Blue dots indicate the stations sampled in summer 2015, cruise ID HE448.</p>
Full article ">Figure 2
<p>(<b>A</b>) Near surface optically active constituent (OAC) concentrations and absorption coefficients in Sognefjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter (CDOM), (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Sognefjord (<span class="html-italic">N</span> = 16). (<span class="html-italic">N</span>: number of sampled stations). (<b>B</b>) Near surface OAC concentrations and absorption coefficients in Trondheimsfjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter CDOM, (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Trondheimsfjord (<span class="html-italic">N</span> = 9). (<span class="html-italic">N</span>: number of sampled stations).</p>
Full article ">Figure 2 Cont.
<p>(<b>A</b>) Near surface optically active constituent (OAC) concentrations and absorption coefficients in Sognefjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter (CDOM), (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Sognefjord (<span class="html-italic">N</span> = 16). (<span class="html-italic">N</span>: number of sampled stations). (<b>B</b>) Near surface OAC concentrations and absorption coefficients in Trondheimsfjord. (<b>a</b>) total suspended matter concentration (TSM), (<b>b</b>) pigment concentration (Tchla), (<b>c</b>) colored dissolved organic matter CDOM, (<b>e</b>) total particulate a<sub>p</sub>440, (<b>f</b>) phytoplankton a<sub>phy</sub>440, (<b>g</b>) non-algal a<sub>nap</sub>440, (<b>h</b>) pigment-specific phytoplankton a*<sub>phy</sub>440 absorption coefficients at 440 nm. (<b>d</b>) sampling stations in Trondheimsfjord (<span class="html-italic">N</span> = 9). (<span class="html-italic">N</span>: number of sampled stations).</p>
Full article ">Figure 3
<p>Average absorption spectra (solid lines) with plus/minus one standard deviation (dotted lines) for total particulate (<span class="html-italic">a<sub>p</sub></span>), phytoplankton (<span class="html-italic">a<sub>phy</sub></span>), non-algal (<span class="html-italic">a<sub>nap</sub></span>), pigment-specific phytoplankton absorption (<span class="html-italic">a*<sub>phy</sub></span>) and colored dissolved organic matter (<span class="html-italic">a<sub>cdom</sub></span>) absorption in Sognefjord (<b>a</b>–<b>e</b>, <span class="html-italic">N</span> = 16) and Trondheimsfjord (<b>f</b>–<b>j</b>, <span class="html-italic">N</span> = 9).</p>
Full article ">Figure 4
<p>Trilinear graphs illustrating the relative contributions of OACs: phytoplankton (<span class="html-italic">a<sub>phy</sub></span>), non-algal particles (<span class="html-italic">a<sub>nap</sub></span>) and CDOM (<span class="html-italic">a<sub>cdom</sub></span>) to the total absorption coefficient at 440 nm in Sognefjord (<b>a</b>, <span class="html-italic">N</span> = 16) and Trondheimsfjord (<b>b</b>, <span class="html-italic">N</span> = 9). CDOM dominated absorption at 440 nm in both fjords under investigation. (<span class="html-italic">N</span>: number of data points).</p>
Full article ">Figure 5
<p>Variations in (<b>a</b>,<b>c</b>,<b>e</b>) total particulate (<span class="html-italic">a<sub>p</sub></span>) and (<b>b</b>,<b>d</b>,<b>f</b>) phytoplankton (<span class="html-italic">a<sub>phy</sub></span>) absorption coefficients as a function of pigment concentration (Tchla) at the blue (440 nm, blue filled circles) and red (675 nm, red filled circles) Chla absorption peaks in Sognefjord (<b>a</b>–<b>d</b>) and Trondheimsfjord (<b>e</b>,<b>f</b>). (<b>a</b>,<b>b</b>) represent trends in Sognefjord corresponding to analysis excluding the inner-fjord stations and (<b>c</b>,<b>d</b>) the trends including the inner-fjord stations. (<span class="html-italic">N</span>: number of data points; R<sup>2</sup>: coefficient of determination).</p>
Full article ">Figure 6
<p>Spectral parameter coefficients <span class="html-italic">A<sub>p</sub></span>, <span class="html-italic">E<sub>p</sub></span> and <span class="html-italic">A<sub>phy</sub></span>, <span class="html-italic">E<sub>phy</sub></span> in the power functions (Equations (7) and (8)) representing variations of the absorption coefficients <span class="html-italic">a<sub>p</sub></span> (<span class="html-italic">λ</span>) and <span class="html-italic">a<sub>phy</sub></span> (<span class="html-italic">λ</span>) respectively as functions of pigment concentration (Tchla) in Sognefjord and Trondheimsfjord. Spectra represent coefficients developed excluding (<b>a</b>,<b>b</b>,<b>g</b>,<b>h</b>) and including (<b>c</b>,<b>d</b>,<b>i</b>,<b>j</b>) the inner Sognefjord stations and Trondheimsfjord (<b>e</b>,<b>f</b>,<b>k</b>,<b>l</b>). (<b>A</b>) (<b>a</b>,<b>c</b>,<b>e</b>) Spectral coefficients of <span class="html-italic">A<sub>p</sub></span> (blue), <span class="html-italic">A<sub>phy</sub></span> (red) in comparison to <span class="html-italic">A<sub>phy</sub></span>. B95 (black) reported for low and mid latitude waters covering Tchla range of 0.02–25.0 mg m<sup>−3</sup> [<a href="#B5-remotesensing-10-00977" class="html-bibr">5</a>]. (<b>b</b>,<b>d</b>,<b>f</b>) Spectral coefficients of <span class="html-italic">E<sub>p</sub></span> (blue), <span class="html-italic">E<sub>phy</sub></span> (red) in comparison to <span class="html-italic">E<sub>phy</sub></span>. B95 (black) reported for low- and mid-latitude waters covering Tchla range of 0.02–25.0 mg m<sup>−3</sup> [<a href="#B5-remotesensing-10-00977" class="html-bibr">5</a>]. (<b>B</b>) Same as in (<b>A</b>) with 95% confidence intervals (ci). Dotted blue and red spectra correspond to ci of the solid blue and red spectra respectively.</p>
Full article ">Figure 7
<p>Variations in pigment-specific phytoplankton absorption coefficients, <span class="html-italic">a*<sub>phy</sub></span> (<span class="html-italic">λ</span>) as functions of Tchla (over a wavelength range 0.96–2.58 mg m<sup>−3</sup>), at selected wavelengths (<b>a</b>–<b>f</b>; 412, 440, 490, 510, 555, 675 nm) in Sognefjord (<span class="html-italic">N</span> = 15). Decrease in <span class="html-italic">a*<sub>phy</sub></span> with increase in Tchla at the blue-green wavelength bands indicating possible effects of pigment packaging. (<span class="html-italic">N</span>: number of data points).</p>
Full article ">
12 pages, 5460 KiB  
Article
Spatio-Temporal Variability of the Habitat Suitability Index for Chub Mackerel (Scomber Japonicus) in the East/Japan Sea and the South Sea of South Korea
by Dabin Lee, SeungHyun Son, Wonkook Kim, Joo Myun Park, Huitae Joo and Sang Heon Lee
Remote Sens. 2018, 10(6), 938; https://doi.org/10.3390/rs10060938 - 13 Jun 2018
Cited by 39 | Viewed by 6086
Abstract
The climate-induced decrease in fish catches in South Korea has been a big concern over the last decades. The increase in sea surface temperature (SST) due to climate change has led to not only a decline in fishery landings but also a shift [...] Read more.
The climate-induced decrease in fish catches in South Korea has been a big concern over the last decades. The increase in sea surface temperature (SST) due to climate change has led to not only a decline in fishery landings but also a shift in the fishing grounds of several fish species. The habitat suitability index (HSI), a reliable indicator of the capacity of a habitant to support selected species, has been widely used to detect and forecast fishing ground formation. In this study, the catch data of the chub mackerel and satellite-derived environmental factors were used to calculate the HSI for the chub mackerel in the South Sea, South Korea. More than 80% of the total catch was found in areas with an SST of 14.72–25.72 °C, chlorophyll-a of 0.30–0.92 mg m−3, and primary production of 523.7–806.46 mg C m−2 d−1. Based on these results, the estimated climatological monthly HSI from 2002 to 2016 clearly showed that the wintering ground of the chub mackerel generally formed in the South Sea of South Korea, coinciding with the catch distribution during the same period. This outcome implies that our estimated HSI can yield a reliable prediction of the fishing ground for the chub mackerel in the East/Japan Sea and South Sea of South Korea. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Summarized commercial catch data for the chub mackerel from 2010 to 2016.</p>
Full article ">Figure 2
<p>Correlation between logarithmic monthly catches and amount of reported catches (<span class="html-italic">n</span> = 67).</p>
Full article ">Figure 3
<p>Monthly distribution of total fishery landings (M/T) and the number of fishing records.</p>
Full article ">Figure 4
<p>Frequency distributions of (<b>a</b>) Sea surface temperature (SST; °C), (<b>b</b>) Chlorophyll-<span class="html-italic">a</span> (Chl-<span class="html-italic">a</span>; mg m<sup>−3</sup>), and (<b>c</b>) Primary production (PP; mg C m<sup>−2</sup> d<sup>−1</sup>) on the fishing locations for the chub mackerel. Gray squares represent the optimum ranges for each parameter.</p>
Full article ">Figure 5
<p>Least squares fitting results of (<b>a</b>) SST (°C), (<b>b</b>) Chl-<span class="html-italic">a</span> (mg m<sup>−3</sup>), and (<b>c</b>) PP (mg C m<sup>−2</sup> d<sup>−1</sup>) with the number of fishing sets (solid line: habitat suitability index (HSI) model, black dot: in situ fishing data).</p>
Full article ">Figure 6
<p>Climatological monthly distribution of the HSI around South Korea from 2002 to 2016.</p>
Full article ">Figure 7
<p>Distribution of fishing records in each range of HSI (gray cross) and the total fishery landings (black dot). Total fishery landings are summed catches in each range of the HSI value. Black line represents the correlation between total fishery landings and HSIs.</p>
Full article ">Figure 8
<p>Spatial distribution of the 8-day composited HSI and the chub mackerel catches at corresponding periods in the South Sea.</p>
Full article ">Figure 9
<p>The HSI hotspots for the chub mackerel observed in the South Sea and the East/Japan Sea.</p>
Full article ">
23 pages, 1581 KiB  
Article
Extraction of Photosynthesis Parameters from Time Series Measurements of In Situ Production: Bermuda Atlantic Time-Series Study
by Žarko Kovač, Trevor Platt, Shubha Sathyendranath and Michael W. Lomas
Remote Sens. 2018, 10(6), 915; https://doi.org/10.3390/rs10060915 - 9 Jun 2018
Cited by 8 | Viewed by 5000
Abstract
Computing the vertical structure of primary production in ocean ecosystem models requires information about the vertical distribution of available light, chlorophyll concentration and photosynthesis response parameters. Conversely, given information on vertical structure of chlorophyll and light, we can extract photosynthesis parameters from vertical [...] Read more.
Computing the vertical structure of primary production in ocean ecosystem models requires information about the vertical distribution of available light, chlorophyll concentration and photosynthesis response parameters. Conversely, given information on vertical structure of chlorophyll and light, we can extract photosynthesis parameters from vertical profiles of primary production measured at sea, as we illustrate here for the Bermuda Atlantic Time-Series Study. The procedure is based on a model of the production profile, which itself depends on the underwater light field. To model the light field, attenuation coefficients were estimated from measured optical profiles using a simple model of exponential decay of photosynthetically-available irradiance with depth, which accounted for 97% of the variance in the measured optical data. With the underwater light climate known, an analytical solution for the production profile was employed to recover photosynthesis parameters by minimizing the residual model error. The recovered parameters were used to model normalized production profiles and normalized watercolumn production. The model explained 95% of the variance in the measured normalized production at depth and 97% of the variance in measured normalized watercolumn production. A shifted Gaussian function was used to model biomass profiles and accounted for 93% of the variance in measured biomass at depth. An analytical solution for watercolumn production with the shifted Gaussian biomass was also tested. With the recovered photosynthesis parameters, maximum instantaneous growth rates were estimated by using a literature value for the carbon-to-chlorophyll ratio in this region of the Atlantic. An exact relationship between the maximum instantaneous growth rate and the daily growth rate in the ocean was derived. It was shown that calculating the growth rate by dividing the production by the carbon-to-chlorophyll ratio is equivalent to calculating it from the ratio of the final to the initial biomass, even when production is time dependent. Finally, the seasonal cycle of the recovered assimilation number at the Bermuda Station was constructed and analysed. The presented approach enables the estimation of photosynthesis parameters and growth rates from measured production profiles with only a few model assumptions, and increases the utility of in situ primary production measurements. The retrieved parameters have direct applications in satellite-based estimates of primary production from ocean-colour data, of which we give an example. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Histograms of estimated parameter values: (<b>a</b>) distribution of the initial slope, <math display="inline"> <semantics> <msup> <mi>α</mi> <mi>B</mi> </msup> </semantics> </math>, obtained from 87 cruises and (<b>b</b>) distribution of the assimilation number, <math display="inline"> <semantics> <msubsup> <mi>P</mi> <mrow> <mi>m</mi> </mrow> <mi>B</mi> </msubsup> </semantics> </math>, obtained from 138 cruises. The abscissa corresponds to the parameter values and the ordinate gives the percentage of cruises that fell into a certain interval of parameter values.</p>
Full article ">Figure 2
<p>Comparison of the model versus measured production at depth obtained by combining the measured data with the estimated parameter values. The abscissa corresponds to the ratio of daily production at depth <math display="inline"> <semantics> <mrow> <msubsup> <mi>P</mi> <mi>T</mi> <mi>B</mi> </msubsup> <mrow> <mo stretchy="false">(</mo> <mi>z</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> to the maximum possible production, <math display="inline"> <semantics> <mrow> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mi>D</mi> </mrow> </semantics> </math>. The ordinate gives the dimensionless irradiance, <math display="inline"> <semantics> <mrow> <msubsup> <mi>I</mi> <mo>∗</mo> <mi>m</mi> </msubsup> <msup> <mi>e</mi> <mrow> <mo>−</mo> <mi>K</mi> <mi>z</mi> </mrow> </msup> </mrow> </semantics> </math>. The continuous curve is recognized as the <math display="inline"> <semantics> <mrow> <msub> <mi>f</mi> <mi>z</mi> </msub> <mrow> <mo stretchy="false">(</mo> <msubsup> <mi>I</mi> <mo>∗</mo> <mi>m</mi> </msubsup> <msup> <mi>e</mi> <mrow> <mo>−</mo> <mi>K</mi> <mi>z</mi> </mrow> </msup> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> function. The coordinates of each point are <math display="inline"> <semantics> <mrow> <mo stretchy="false">(</mo> <msubsup> <mover accent="true"> <mi>P</mi> <mo>˜</mo> </mover> <mi>T</mi> <mi>B</mi> </msubsup> <mrow> <mo stretchy="false">(</mo> <msub> <mi>z</mi> <mi>n</mi> </msub> <mo stretchy="false">)</mo> </mrow> <mo>/</mo> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mi>D</mi> <mo>,</mo> <msup> <mi>α</mi> <mi>B</mi> </msup> <msubsup> <mover accent="true"> <mi>I</mi> <mo>˜</mo> </mover> <mn>0</mn> <mi>m</mi> </msubsup> <msup> <mi>e</mi> <mrow> <mo>−</mo> <mover accent="true"> <mi>K</mi> <mo>˜</mo> </mover> <msub> <mi>z</mi> <mi>n</mi> </msub> </mrow> </msup> <mo>/</mo> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mo stretchy="false">)</mo> </mrow> </semantics> </math>, where <math display="inline"> <semantics> <msup> <mi>α</mi> <mi>B</mi> </msup> </semantics> </math> and <math display="inline"> <semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics> </math> are the estimated parameters for each profile. The <math display="inline"> <semantics> <msup> <mi>r</mi> <mn>2</mn> </msup> </semantics> </math> value between the measured normalized production and the modeled normalized production is 0.95. In total, there are 1049 points.</p>
Full article ">Figure 3
<p>Comparison of the model and measured normalized daily watercolumn production obtained by combining the measured data with the estimated parameter values. The abscissa is the dimensionless irradiance, <math display="inline"> <semantics> <msubsup> <mi>I</mi> <mo>∗</mo> <mi>m</mi> </msubsup> </semantics> </math>, and the ordinate is the ratio of normalized watercolumn production to <math display="inline"> <semantics> <mrow> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mi>D</mi> <mo>/</mo> <mi>K</mi> </mrow> </semantics> </math>. The continuous curve is the <math display="inline"> <semantics> <mrow> <mi>f</mi> <mo stretchy="false">(</mo> <msubsup> <mi>I</mi> <mo>∗</mo> <mi>m</mi> </msubsup> <mo stretchy="false">)</mo> </mrow> </semantics> </math> function (<a href="#FD10-remotesensing-10-00915" class="html-disp-formula">10</a>). The coordinates of each point are <math display="inline"> <semantics> <mrow> <mo stretchy="false">(</mo> <msubsup> <mover accent="true"> <mi>P</mi> <mo>˜</mo> </mover> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> <mi>B</mi> </msubsup> <mover accent="true"> <mi>K</mi> <mo>˜</mo> </mover> <mo>/</mo> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mi>D</mi> <mo>,</mo> <msup> <mi>α</mi> <mi>B</mi> </msup> <msubsup> <mover accent="true"> <mi>I</mi> <mo>˜</mo> </mover> <mn>0</mn> <mi>m</mi> </msubsup> <mo>/</mo> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mo stretchy="false">)</mo> </mrow> </semantics> </math>, where <math display="inline"> <semantics> <msup> <mi>α</mi> <mi>B</mi> </msup> </semantics> </math> and <math display="inline"> <semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics> </math> are the estimated parameters for each profile. The <math display="inline"> <semantics> <msup> <mi>r</mi> <mn>2</mn> </msup> </semantics> </math> between the measured normalized watercolumn production and the modeled normalized watercolumn production is 0.97.</p>
Full article ">Figure 4
<p>(<b>a</b>) Scatter plot of measured <math display="inline"> <semantics> <mover accent="true"> <mi>B</mi> <mo>˜</mo> </mover> </semantics> </math> and modelled biomass <span class="html-italic">B</span> with the shifted Gaussian function. There are, in total, 1049 points. (<b>b</b>) Scatter plot of measured <math display="inline"> <semantics> <msub> <mover accent="true"> <mi>P</mi> <mo>˜</mo> </mover> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics> </math> and modeled <math display="inline"> <semantics> <msub> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> </msub> </semantics> </math> watercolumn production with the analytical solution for the shifted Gaussian biomass. There are, in total, 138 points. The grey line on both plots represents the 1:1 model versus the measurement ratio.</p>
Full article ">Figure 5
<p>Histogram of the estimated maximum daily growth rate, <math display="inline"> <semantics> <msub> <mi>μ</mi> <mi>m</mi> </msub> </semantics> </math>, obtained from 138 cruises based on Equation (<a href="#FD12-remotesensing-10-00915" class="html-disp-formula">12</a>) and here multiplied by the daylength, <span class="html-italic">D</span>, to convert hourly into daily rates. The value of the carbon-to-chlorophyll ratio <math display="inline"> <semantics> <mi>χ</mi> </semantics> </math> is 146 mg C (mg Chl)<sup>−1</sup>, taken from Maranon (2005).</p>
Full article ">Figure 6
<p>Estimated seasonal cycles for the BATS station. The thin curves represent the monthly averages on each Julian day, which were calculated from the daily values, 15 days prior and 15 days post a given Julian day. The thick curves are the fits of a sum of two sine functions (superimposed onto an annual mean) to the monthly averages. (<b>a</b>) Seasonal cycle of <math display="inline"> <semantics> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> </semantics> </math> (blue) and <math display="inline"> <semantics> <msub> <mi>B</mi> <mi>Z</mi> </msub> </semantics> </math> (orange). (<b>b</b>) Seasonal cycle of <math display="inline"> <semantics> <msubsup> <mi>P</mi> <mrow> <mi>Z</mi> <mo>,</mo> <mi>T</mi> </mrow> <mi>B</mi> </msubsup> </semantics> </math> (red).</p>
Full article ">Figure 7
<p>Comparison of the measured (grey curve) versus the modelled seasonal cycle of watercolumn production based on remotely sensed-chlorophyll with the time-dependent assimilation number, <math display="inline"> <semantics> <mrow> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mrow> <mo stretchy="false">(</mo> <mi>t</mi> <mo stretchy="false">)</mo> </mrow> </mrow> </semantics> </math> (blue curve), and the average assimilation number, <math display="inline"> <semantics> <mrow> <mo>〈</mo> <msubsup> <mi>P</mi> <mi>m</mi> <mi>B</mi> </msubsup> <mo>〉</mo> </mrow> </semantics> </math> (dashed blue curve).</p>
Full article ">
20 pages, 5139 KiB  
Article
Scratching Beneath the Surface: A Model to Predict the Vertical Distribution of Prochlorococcus Using Remote Sensing
by Priscila K. Lange, Robert J. W. Brewin, Giorgio Dall’Olmo, Glen A. Tarran, Shubha Sathyendranath, Mikhail Zubkov and Heather A. Bouman
Remote Sens. 2018, 10(6), 847; https://doi.org/10.3390/rs10060847 - 29 May 2018
Cited by 19 | Viewed by 6128
Abstract
The unicellular cyanobacterium Prochlorococcus is the most dominant resident of the subtropical gyres, which are considered to be the largest biomes on earth. In this study, the spatial and temporal variability in the global distribution of Prochlorococcus was estimated in the Atlantic Ocean [...] Read more.
The unicellular cyanobacterium Prochlorococcus is the most dominant resident of the subtropical gyres, which are considered to be the largest biomes on earth. In this study, the spatial and temporal variability in the global distribution of Prochlorococcus was estimated in the Atlantic Ocean using an empirical model based on data from 13 Atlantic Meridional Transect cruises. Our model uses satellite-derived sea surface temperature (SST), remote-sensing reflectance at 443 and 488 nm, and the water temperature at a depth of 200 m from Argo data. The model divides the population of Prochlorococcus into two groups: ProI, which dominates under high-light conditions associated with the surface, and ProII, which favors low light found near the deep chlorophyll maximum. ProI and ProII are then summed to provide vertical profiles of the concentration of Prochlorococcus cells. This model predicts that Prochlorococcus cells contribute 32 Mt of carbon biomass (7.4 × 1026 cells) to the Atlantic Ocean, concentrated mainly within the subtropical gyres (35%) and areas near the Equatorial Convergence Zone (30%). When projected globally, 3.4 × 1027 Prochlorococcus cells represent 171 Mt of carbon biomass, with 43% of this global biomass allocated to the upper ocean (0–45 m depth). Annual cell standing stocks were relatively stable between the years 2003 and 2014, and the contribution of the gyres varies seasonally as gyres expand and contract, tracking changes in light and temperature, with lowest cell abundances during the boreal and austral winter (1.4 × 1013 cells m−2), when surface cell concentrations were highest (9.8 × 104 cells mL−1), whereas the opposite scenario was observed in spring–summer (2 × 1013 cells m−2). This model provides a three-dimensional view of the abundance of Prochlorococcus cells, revealing that Prochlorococcus contributes significantly to total phytoplankton biomass in the Atlantic Ocean, and can be applied using either in situ measurements at the sea surface (r2 = 0.83) or remote-sensing observables (r2 = 0.58). Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Transects of the 13 AMT cruises (704 stations) used to create the empirical model to predict the abundance of <span class="html-italic">Prochlorococcus</span> cells.</p>
Full article ">Figure 2
<p>Flowchart of computations used to calculate the cell abundance of <span class="html-italic">Prochlorococcus</span>. Variable acronyms and symbols are described in <a href="#remotesensing-10-00847-t001" class="html-table">Table 1</a>.</p>
Full article ">Figure 3
<p>Vertical profiles of observed and estimated <span class="html-italic">Prochlorococcus</span> cell abundances (<b>a</b>) over depth and (<b>b</b>) over the fractional PAR <span class="html-italic">fPAR</span>, with (<b>c</b>) corresponding profiles of temperature and chlorophyll from CTD measurements at a site inside the North Atlantic Gyre (26° N, 50° W). For (<b>a</b>,<b>b</b>), in situ observations are represented by red dots, predicted profiles of <span class="html-italic">ProI</span> by the orange dashed line, predicted profiles of <span class="html-italic">ProII</span> by the blue dashed line, and predicted profiles of total <span class="html-italic">Prochlorococcus</span> abundance by the solid black line. Data from AMT 24 (2014).</p>
Full article ">Figure 4
<p>(<b>a</b>) Comparison between observed versus predicted depth of the deep chlorophyll maximum (<span class="html-italic">Z<sub>DCM</sub></span>) across the Atlantic Ocean (locations displayed in <a href="#remotesensing-10-00847-f001" class="html-fig">Figure 1</a>) using Equation (8) of the present work; (<b>b</b>) Observed (AMT12-24, <span class="html-italic">n</span> = 693 observations) and predicted <span class="html-italic">Z<sub>DCM</sub></span> (<span class="html-italic">n</span> = 449) in the Atlantic Ocean (AMTs 12 to 24, locations displayed in <a href="#remotesensing-10-00847-f001" class="html-fig">Figure 1</a>). For each CTD cast, observed <span class="html-italic">Z<sub>DCM</sub></span> was determined as the depth of the maximum chlorophyll concentration measured using the CTD fluorometer.</p>
Full article ">Figure 5
<p>Vertical profiles of <span class="html-italic">Prochlorococcus</span> cell abundance over depth (<span class="html-italic">Pro<sub>total</sub></span>(<span class="html-italic">z</span>)) across the Atlantic Ocean (locations displayed in <a href="#remotesensing-10-00847-f001" class="html-fig">Figure 1</a>): (<b>a</b>) observed in situ on AMTs 12–24 (<span class="html-italic">Pro<sub>total</sub></span>(<span class="html-italic">z</span>)<math display="inline"><semantics> <mo>′</mo> </semantics></math>); (<b>b</b>) predicted using the partial model with observed inputs of <span class="html-italic">Pro<sub>surf</sub></span> and <span class="html-italic">Z<sub>DCM</sub></span> (<span class="html-italic">Pro<sub>total</sub></span>(<span class="html-italic">z</span>)<span class="html-italic"><sup>1</sup></span> from Equations (1) to (2) and (6) to (7)); and (<b>c</b>) predicted using the full model with remote-sensing inputs (<span class="html-italic">Pro<sub>total</sub></span>(<span class="html-italic">z</span>)<span class="html-italic"><sup>3</sup></span> from Equations (1)–(9).</p>
Full article ">Figure 6
<p>(<b>a</b>) Comparison between observed and predicted <span class="html-italic">Prochlorococcus</span> cell abundance integrated in the water column (<span class="html-italic">Pro<sub>int</sub><sup>1</sup></span>) across the Atlantic Ocean (locations displayed in <a href="#remotesensing-10-00847-f001" class="html-fig">Figure 1</a>) using the partial model where in situ observations of <span class="html-italic">Z<sub>DCM</sub></span>, <span class="html-italic">K<sub>d</sub>PAR</span>, and <span class="html-italic">Pro<sub>surf</sub></span> are used as inputs (i.e., Equations (1), (3)–(5), and (8) are excluded); (<b>b</b>) Observed and predicted <span class="html-italic">Pro<sub>int</sub><sup>1</sup></span> across the Atlantic Ocean (AMTs 12–24).</p>
Full article ">Figure 7
<p>(<b>a</b>) Comparison between observed and predicted <span class="html-italic">Prochlorococcus</span> cell abundance integrated in the water column (<span class="html-italic">Pro<sub>int</sub><sup>3</sup></span>) across the Atlantic Ocean (locations displayed in <a href="#remotesensing-10-00847-f001" class="html-fig">Figure 1</a>) using the complete model (Equations (1)–(10) of the present work); (<b>b</b>) Observed and predicted <span class="html-italic">Pro<sub>int</sub><sup>3</sup></span> across the Atlantic Ocean (AMTs 12–24); (<b>c</b>) Comparison between observed and predicted <span class="html-italic">Prochlorococcus</span> cell abundance at the sea surface (<span class="html-italic">Pro<sub>surf</sub></span>) using Equations (3)–(5) of the present work; (<b>d</b>) Observed and predicted <span class="html-italic">Pro<sub>surf</sub></span> across the Atlantic Ocean (AMTs 12–24).</p>
Full article ">Figure 8
<p>Estimated monthly distribution of the <span class="html-italic">Prochlorococcus</span> cells integrated in the top 200 m of the water column (<span class="html-italic">Pro<sub>int</sub></span>) (cells m<sup>−2</sup>), with correspondent vertical profiles of estimated <span class="html-italic">Prochlorococcus</span> cell abundance (cells l<sup>−1</sup>) at sites (white dots) in the North Atlantic Gyre (NAG) and South Atlantic Gyre: <span class="html-italic">ProI</span>(<span class="html-italic">z</span>) is indicated by the orange dashed line, <span class="html-italic">ProII</span>(<span class="html-italic">z</span>) by the blue dashed line, and the total <span class="html-italic">Prochlorococcus</span> abundance by the solid black line. Cell abundance was calculated based on the monthly climatology of environmental variables [<a href="#B22-remotesensing-10-00847" class="html-bibr">22</a>].</p>
Full article ">Figure 9
<p>Monthly averages of (<b>a</b>) the areal extent of the North Atlantic Gyre (NAG) and (<b>b</b>) South Atlantic Gyre (SAG); (<b>c</b>) the euphotic depth <span class="html-italic">z<sub>eu</sub></span> at the NAG and (<b>d</b>) the SAG; (<b>e</b>) sea surface temperature <span class="html-italic">SST</span> at the NAG and (<b>f</b>) the SAG; (<b>g</b>) the estimated <span class="html-italic">Prochlorococcus</span> cell abundance at the sea surface (<span class="html-italic">Pro<sub>surf</sub></span>) at the NAG and (<b>h</b>) the SAG; (<b>i</b>) <span class="html-italic">Prochlorococcus</span> cell abundance at the deep maximum (<span class="html-italic">Pro<sub>max</sub></span>) at the NAG and (<b>j</b>) the SAG; and (<b>k</b>) <span class="html-italic">Prochlorococcus</span> cell abundance integrated in the water column (<span class="html-italic">Pro<sub>int</sub></span>) at the NAG and (<b>l</b>) the SAG. Subtropical Gyres were defined as regions where the surface chlorophyll concentrations were lower than 0.075 mg m<sup>−3</sup>. The euphotic depth was estimated using the calculated <span class="html-italic">K<sub>d</sub>PAR</span> using Equation (1), and <span class="html-italic">SST</span> measurements were taken from monthly satellite composites from January 2003 to December 2014 [<a href="#B22-remotesensing-10-00847" class="html-bibr">22</a>]. <span class="html-italic">Prochlorococcus</span> cell abundance was computed using monthly averaged input variables for the years 2003–2014 [<a href="#B22-remotesensing-10-00847" class="html-bibr">22</a>], at specific locations inside the Atlantic gyres: NAG: 26° N, 50° W; SAG: 20° S, 20° W.</p>
Full article ">Figure 10
<p>Time series of (<b>a</b>,<b>g</b>) the estimated <span class="html-italic">Prochlorococcus</span> cell abundance integrated in the water column (<span class="html-italic">Pro<sub>int</sub></span>), (<b>c</b>,<b>i</b>) estimated <span class="html-italic">Prochlorococcus</span> cell abundance at the deep maximum (<span class="html-italic">Pro<sub>max</sub></span>), and (<b>e</b>,<b>k</b>) estimated <span class="html-italic">Prochlorococcus</span> cell abundance at surface (<span class="html-italic">Pro<sub>surf</sub></span>), at one location (26° N, 50° W) in the North Atlantic Gyre (NAG) (<b>a</b>–<b>f</b>), and one location (20° S, 20° W) in the South Atlantic Gyre (SAG) (<b>g</b>–<b>l</b>). Figures (<b>b</b>,<b>d</b>,<b>f</b>,<b>h</b>,<b>j</b>,<b>l</b>) show the time-series anomalies (black bars) and trends (red lines). Anomalies were calculated by subtracting the monthly climatology (dashed red lines in figures (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>)) from the calculated values for each month (black lines in figures (<b>a</b>,<b>c</b>,<b>e</b>,<b>g</b>,<b>i</b>,<b>k</b>)).</p>
Full article ">
22 pages, 3910 KiB  
Article
Canopy Reflectance Modeling of Aquatic Vegetation for Algorithm Development: Global Sensitivity Analysis
by Guanhua Zhou, Zhongqi Ma, Shubha Sathyendranath, Trevor Platt, Cheng Jiang and Kang Sun
Remote Sens. 2018, 10(6), 837; https://doi.org/10.3390/rs10060837 - 27 May 2018
Cited by 22 | Viewed by 6602
Abstract
Optical remote sensing of aquatic vegetation in shallow water is an essential aid to ecosystem protection, but it is difficult because the spectral characteristics of the vegetation are sensitive to external features such as water background effects, atmospheric effects, and the structural properties [...] Read more.
Optical remote sensing of aquatic vegetation in shallow water is an essential aid to ecosystem protection, but it is difficult because the spectral characteristics of the vegetation are sensitive to external features such as water background effects, atmospheric effects, and the structural properties of the canopy. A global sensitivity analysis of an aquatic vegetation radiative transfer model provides invaluable background for algorithm development for use in optical remote sensing. Here, we use the extended Fourier Amplitude Sensitivity Test (EFAST) method for the modelling. Four different cases were identified by subdividing the ranges of water depth and leaf area index (LAI) involved. The results indicate that the reflectance of emergent vegetation is affected mainly by the concentrations of chlorophyll a + b in leaves (Cab), leaf inclination distribution function parameter (LIDFa) and LAI. The parameter LAI is influential in sparse vegetation cases whereas Cab and LIDFa are influential in dense vegetation cases. Canopy reflectance for submerged vegetation is dominated by water parameters. Relatively, LAI and Cab are highly sensitive vegetation parameters. The analysis is extended to vegetation index as well, which takes the Sentinel-2A as the reference sensor. It shows that NDAVI (Normalized Difference Aquatic Vegetation Index) is suitable for retrieving LAI in all cases except deep-sparse for emergent vegetation, whereas NDVI (Normalized Difference Vegetation Index) would be better in the deep-sparse case. NDVI, NDAVI and WAVI (Water Adjusted Vegetation Index), respectively, are suitable for retrieving Cab, Car and LIDFa in dense cases. For submerged vegetation, the sensitivity of LAI to NDAVI is relatively high only in the shallow-sparse case. The adjustment factor L in SAVI and WAVI fails to suppress the sensitivity to water constituent parameters. The sensitivity of LAI and Cab to NDVI in deep cases is relatively higher than that to the other indices, which may provide clues for the construction of inversion algorithms in macrophyte remote sensing in the aquatic environment using spectral signatures in the visible and near infrared regions. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The Schematic of vertical structures for emergent vegetation and submerged vegetation. (<b>a</b>) Emergent vegetation; (<b>b</b>) Submerged vegetation.</p>
Full article ">Figure 2
<p>The Sentinel-2A spectral response functions of bands (B1–B9) in 400–1000 nm.</p>
Full article ">Figure 3
<p>The cumulative plot of normalized total order sensitivity (<span class="html-italic">NST</span>) for emergent vegetation in four different cases. (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the cases of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively. The area enclosed by the curve and the horizontal axis (the first parameter in the list, Hw), and between the curves (except Hw) represents the value of <span class="html-italic">NST</span>.</p>
Full article ">Figure 4
<p>In four different cases, the curves of high sensitivity parameters varying with wavelength. The columns (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the case of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively. (1)–(9) represent the high sensitivity parameters, (1): N, (2): Concentration of chlorophyll a + b (Cab), (3): Concentration of carotenoid (Car), (4): Concentration of dry matter (Cm), (5): Leaf area index (LAI), (6): Leaf inclination distribution function parameter a (LIDFa), (7): Concentration of chlorophyll a, in water (Cchla), (8): Coefficient to calculate scattering of total suspended matter (Btsm), (9): Concentration of suspended matter (SPM). It should be noted that the curves of absorption coefficient of Cab and Car are also shown in (2) and (3). The vertical axis on the right shows the range of the absorption coefficient.</p>
Full article ">Figure 5
<p>The cumulative plot of <span class="html-italic">NST</span> for submerged vegetation in four different cases. (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the cases of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively. The area enclosed by the curve and the horizontal axis (the first parameter in the list, Hw), and between the curves (except Hw) represents the value of <span class="html-italic">NST</span>.</p>
Full article ">Figure 6
<p>In four different cases, the curves of high sensitivity parameters varying with wavelength. The columns (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the case of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively. (1)–(9) represent the high sensitivity parameters of the model, (1): The height of the upper water layer (Hw), (2): Concentration of chlorophyll a, in water (Cchla), (3): Coefficient to calculate scattering of total suspended matter (Btsm), (4): Concentration of suspended matter (SPM), (5): Absorption coefficient of CDOM at 375 nm (aCDOM), (6): Plant height (Hp), (7): Concentration of chlorophyll a+b (Cab), (8): Leaf area index (LAI).</p>
Full article ">Figure 7
<p>The <span class="html-italic">NST</span> of high sensitivity parameters to vegetation indices for emergent vegetation. (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the case of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively.</p>
Full article ">Figure 8
<p>The <span class="html-italic">NST</span> of high sensitivity parameters to vegetation indices for submerged vegetation. (<b>a</b>), (<b>b</b>), (<b>c</b>), (<b>d</b>) represent the cases of shallow-sparse, shallow-dense, deep-sparse and deep-dense, respectively.</p>
Full article ">
23 pages, 5888 KiB  
Article
Phytoplankton Size Structure in Association with Mesoscale Eddies off Central-Southern Chile: The Satellite Application of a Phytoplankton Size-Class Model
by Andrea Corredor-Acosta, Carmen E. Morales, Robert J. W. Brewin, Pierre-Amaël Auger, Oscar Pizarro, Samuel Hormazabal and Valeria Anabalón
Remote Sens. 2018, 10(6), 834; https://doi.org/10.3390/rs10060834 - 25 May 2018
Cited by 21 | Viewed by 7253
Abstract
Understanding the influence of mesoscale and submesoscale features on the structure of phytoplankton is a key aspect in the assessment of their influence on marine biogeochemical cycling and cross-shore exchanges of plankton in Eastern Boundary Current Systems (EBCS). In this study, the spatio-temporal [...] Read more.
Understanding the influence of mesoscale and submesoscale features on the structure of phytoplankton is a key aspect in the assessment of their influence on marine biogeochemical cycling and cross-shore exchanges of plankton in Eastern Boundary Current Systems (EBCS). In this study, the spatio-temporal evolution of phytoplankton size classes (PSC) in surface waters associated with mesoscale eddies in the EBCS off central-southern Chile was analyzed. Chlorophyll-a (Chl-a) size-fractionated filtration (SFF) data from in situ samplings in coastal and coastal transition waters were used to tune a three-component (micro-, nano-, and pico-phytoplankton) model, which was then applied to total Chl-a satellite data (ESA OC-CCI product) in order to retrieve the Chl-a concentration of each PSC. A sea surface, height-based eddy-tracking algorithm was used to identify and track one cyclonic (sC) and three anticyclonic (ssAC1, ssAC2, sAC) mesoscale eddies between January 2014 and October 2015. Satellite estimates of PSC and in situ SFF Chl-a data were highly correlated (0.64 < r < 0.87), although uncertainty values for the microplankton fraction were moderate to high (50 to 100% depending on the metric used). The largest changes in size structure took place during the early life of eddies (~2 months), and no major differences in PSC between eddy center and periphery were found. The contribution of the microplankton fraction was ~50% (~30%) in sC and ssAC1 (ssAC2 and sAC) eddies when they were located close to the coast, while nanoplankton was dominant (~60–70%) and picoplankton almost constant (<20%) throughout the lifetime of eddies. These results suggest that the three-component model, which has been mostly applied in oceanic waters, is also applicable to highly productive coastal upwelling systems. Additionally, the PSC changes within mesoscale eddies obtained by this satellite approach are in agreement with results on phytoplankton size distribution in mesoscale and submesoscale features in this region, and are most likely triggered by variations in nutrient concentrations and/or ratios during the eddies’ lifetimes. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area in central-southern Chile. An eight-day composite (25 January 2014–1 February 2014) of surface total chlorophyll-a (Chl-a), obtained from version 3.0 of the Ocean Colour Climate Change Initiative (OC-CCI; 4 km resolution) product, is represented in red-blue color scale. The geostrophic velocity, obtained from Ssalto/Duacs multimission altimeter AVISO product (<a href="http://www.aviso.altimetry.fr" target="_blank">http://www.aviso.altimetry.fr</a>), is shown in gray arrows. The blue star indicates the location of the COPAS coastal time series Station 18 and, together with the black dots, indicates the locations of the size-fractionated filtration (SFF) Chl-a in situ data (≤ 10 m depth), during different campaigns (November 2004–September 2015). The red lines indicate the offshore limit of the coastal zone (CZ; ~100 km from the coast) and the coastal transition zone (CTZ; coast to ~800 km offshore), respectively.</p>
Full article ">Figure 2
<p>In situ concentrations of size-fractionated Chl-a for micro- (<span class="html-italic">C<sub>M</sub></span>), nano- and pico- (<span class="html-italic">C<sub>NP</sub></span>), nano- (<span class="html-italic">C<sub>N</sub></span>), and picoplankton (<span class="html-italic">C<sub>P</sub></span>) (upper panels: <b>a</b>–<b>d</b>), and their contribution to total in situ Chl-a (lower panels: <b>e</b>–<b>h</b>) as a function of total in situ Chl-a concentration (<span class="html-italic">C</span>). The fitted three-component model of Brewin et al. [<a href="#B43-remotesensing-10-00834" class="html-bibr">43</a>] is overlaid in each case (solid black line). The subscript ‘i’ indicates the different Chl-a size classes. The location of the samples in the CZ and CTZ is differentiated by grey dots and black triangles, respectively.</p>
Full article ">Figure 3
<p>Total satellite Chl-a concentration (OC-CCI product; 4 km resolution) and size-fractionated Chl-a estimates obtained from the regional three-component model, as a function of in situ total and size fractioned Chl-a concentration. The black dotted line represents the 1:1 line (perfect relationship between in situ and satellite estimates).</p>
Full article ">Figure 4
<p>Spatial distribution of the satellite Chl-a estimates for size-fractionated concentrations (<b>a</b>–<b>c</b>; mg m<sup>−3</sup>) obtained from the regional three-component model applied to the total Chl-a satellite data, and their relative contribution to total Chl-a (<b>d</b>–<b>f</b>; dimensionless). The geostrophic velocity field is shown (black arrows). Data in this figure are an example of an eight-day composite for the same dates included in <a href="#remotesensing-10-00834-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 5
<p>The geostrophic velocity field (black arrows) and the trajectories of the tracked mesoscale eddies (upper panels: <b>a</b>,<b>d</b>,<b>g</b>,<b>j</b>), together with radius (<b>b</b>,<b>e</b>,<b>h</b>,<b>k</b>) and displacement velocity (<b>c</b>,<b>f</b>,<b>i</b>,<b>l</b>) through time (in weeks). The red circles in the upper panels represent the eddies in a time step of their complete trajectories (blue-yellow color scale). Four eddies were analyzed: subsurface anticyclone 1 (ssAC1; 27 December 2014 to 13 August 2015, 30 weeks), subsurface anticyclone 2 (ssAC2; 1 January 2014 to 21 August 2014, 30 weeks), surface anticyclone (sAC; 10 February 2014 to 1 November 2014, 34 weeks), and surface cyclone (sC; 9 November 2014 to 8 October 2015, 43 weeks). Gaps in the data for some periods (weeks) were created after screening out size-fractionated Chl-a fields that did not cover at least 50% of the spatial extent of an eddy.</p>
Full article ">Figure 6
<p>Eight-day composites of surface total Chl-a anomalies (green-blue color scale) for three different time steps of eddies trajectories. The position and radius of the eddies in each composite (red circles) and the geostrophic velocity field (gray/black arrows) are also shown. The dates of the composites are ssAC1 (<b>a</b>–<b>c</b>) 9–16 January 2015, 14–21 March 2015 and 2–9 June 2015; ssAC2 (<b>d</b>–<b>f</b>) 25 January to 1 February 2014, 22–29 March 2014 and 10–17 June 2014; sAC (<b>g</b>–<b>i</b>) 14–21 March 2014, 17–24 May 2014 and 29 August to 5 September 2014; and sC (<b>j</b>–<b>l</b>) 1–8 January 2015, 26 February to 5 March 2015 and 29 August to 5 September 2015.</p>
Full article ">Figure 7
<p>Temporal variability of phytoplankton size classes (PSC) in terms of the size-specific fractional (relative) contributions to total Chl-a (dimensionless) in the four selected eddies (center: continuous black line; periphery: dashed black line) during the period indicated in <a href="#remotesensing-10-00834-f005" class="html-fig">Figure 5</a>. The mean contribution by each fraction in the CZ for the first week of eddy tracking is denoted by the red dots. The vertical dashed gray lines indicate the weeks of the composites of surface total Chl-a anomalies presented in <a href="#remotesensing-10-00834-f006" class="html-fig">Figure 6</a>. Gaps in the data are explained in <a href="#remotesensing-10-00834-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure A1
<p>Flow-Diagram of the Methodological Procedures Used in This Study.</p>
Full article ">
25 pages, 2171 KiB  
Article
Machine Learning Regression Approaches for Colored Dissolved Organic Matter (CDOM) Retrieval with S2-MSI and S3-OLCI Simulated Data
by Ana Belen Ruescas, Martin Hieronymi, Gonzalo Mateo-Garcia, Sampsa Koponen, Kari Kallio and Gustau Camps-Valls
Remote Sens. 2018, 10(5), 786; https://doi.org/10.3390/rs10050786 - 19 May 2018
Cited by 71 | Viewed by 8919
Abstract
The colored dissolved organic matter (CDOM) variable is the standard measure of humic substance in waters optics. CDOM is optically characterized by its spectral absorption coefficient, a C D O M at at reference wavelength (e.g., ≈ 440 nm). Retrieval of CDOM is [...] Read more.
The colored dissolved organic matter (CDOM) variable is the standard measure of humic substance in waters optics. CDOM is optically characterized by its spectral absorption coefficient, a C D O M at at reference wavelength (e.g., ≈ 440 nm). Retrieval of CDOM is traditionally done using bio-optical models. As an alternative, this paper presents a comparison of five machine learning methods applied to Sentinel-2 and Sentinel-3 simulated reflectance ( R r s ) data for the retrieval of CDOM: regularized linear regression (RLR), random forest regression (RFR), kernel ridge regression (KRR), Gaussian process regression (GPR) and support vector machines (SVR). Two different datasets of radiative transfer simulations are used for the development and training of the machine learning regression approaches. Statistics comparison with well-established polynomial regression algorithms shows optimistic results for all models and band combinations, highlighting the good performance of the methods, especially the GPR approach, when all bands are used as input. Application to an atmospheric corrected OLCI image using the reflectance derived form the alternative neural network (Case 2 Regional) is also shown. Python scripts and notebooks are provided to interested users. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Polynomial regressions using the simulated Sentinel-2 Multi-Spectral Instrument(S2-MSI) and Sentinel-3 Ocean and Land Colour Instrument (S3-OLCI) configuration datasets.</p>
Full article ">Figure 2
<p>Box-plots of the residuals <span class="html-italic">S2 two ratios</span> and <span class="html-italic">S2 All bands + ratios.</span> On each box, the central mark is the median, the edges of the box are the lower hinge (defined as the 25th percentile) and the upper hinge (the 75th percentile), the whiskers extend to the most extreme data points not considered outliers.</p>
Full article ">Figure 3
<p>Comparison of the performance of the models using the linear regression representation, S2-MSI. On the y axis are the normalized CDOM values, on the x axis the value of the ratios.</p>
Full article ">Figure 4
<p>Permutation plots for four ML methods for the S2-MSI configuration.</p>
Full article ">Figure 5
<p>Box-plots of the residuals for <span class="html-italic">S3 two ratios</span> and <span class="html-italic">S3 All bands + ratios</span>.</p>
Full article ">Figure 6
<p>Comparison of the performance of the models using the linear regression representation, S3-OLCI.</p>
Full article ">Figure 7
<p>Permutation plots for four ML methods for the S3-OLCI configuration.</p>
Full article ">Figure 8
<p>Ranges and mean of reflectance spectra at S3-OLCI wavebands from the C2X dataset for C2A and C2AX subsets. The utilized S2-MSI and S3-OLCI bands are highlighted for convenience.</p>
Full article ">Figure 9
<p>Statistic box-plots of the residual errors on the top; permutation plots of the RLR and GPR model on the bottom—C2X S3-OLCI.</p>
Full article ">Figure 10
<p>Retrieval performance with respect to the C2A(X) CDOM simulations: on the top left the scatter plot of the Polyfit method with the <span class="html-italic">Ratio1</span>; on the top right the scatter plot of the Polyfit method with the <span class="html-italic">Ratio2</span>. On the bottom left RLR method using all available bands and the two ratios as input; on the bottom right the GPR method with all available bands and the two ratios.</p>
Full article ">Figure 11
<p>Absorption coefficient of CDOM at 440 nm retrieved by ONNS (IOP NNs) for the C2A (red dots) and C2AX (blue dots) spectral types. Plot is in log10 scale.</p>
Full article ">Figure 12
<p>Computational time in seconds of each model and band combination on an standard OLCI scene.</p>
Full article ">Figure 13
<p>Comparison of ADG443_NN vs. GPR model (m<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>): left ADG443_NN product; right GPR CDOM output.</p>
Full article ">Figure 14
<p>Comparison of ADG443_NN uncertainties vs. the uncertainties of the GPR model (m<math display="inline"><semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>1</mn> </mrow> </msup> </semantics></math>): left ADG443 uncertainties; right GPR CDOM uncertainties.</p>
Full article ">
27 pages, 4966 KiB  
Article
Remotely Sensing the Biophysical Drivers of Sardinella aurita Variability in Ivorian Waters
by Jean-Baptiste Kassi, Marie-Fanny Racault, Brice A. Mobio, Trevor Platt, Shubha Sathyendranath, Dionysios E. Raitsos and Kouadio Affian
Remote Sens. 2018, 10(5), 785; https://doi.org/10.3390/rs10050785 - 18 May 2018
Cited by 13 | Viewed by 6285
Abstract
The coastal regions of the Gulf of Guinea constitute one of the major marine ecosystems, producing essential living marine resources for the populations of Western Africa. In this region, the Ivorian continental shelf is under pressure from various anthropogenic sources, which have put [...] Read more.
The coastal regions of the Gulf of Guinea constitute one of the major marine ecosystems, producing essential living marine resources for the populations of Western Africa. In this region, the Ivorian continental shelf is under pressure from various anthropogenic sources, which have put the regional fish stocks, especially Sardinella aurita, the dominant pelagic species in Ivorian industrial fishery landings, under threat from overfishing. Here, we combine in situ observations of Sardinella aurita catch, temperature, and nutrient profiles, with remote-sensing ocean-color observations, and reanalysis data of wind and sea surface temperature, to investigate relationships between Sardinella aurita catch and oceanic primary producers (including biomass and phenology of phytoplankton), and between Sardinella aurita catch and environmental conditions (including upwelling index, and turbulent mixing). We show that variations in Sardinella aurita catch in the following year may be predicted, with a confidence of 78%, based on a bilinear model using only physical variables, and with a confidence of 40% when using only biological variables. However, the physics-based model alone is not sufficient to explain the mechanism driving the year-to-year variations in Sardinella aurita catch. Based on the analysis of the relationships between biological variables, we demonstrate that in the Ivorian continental shelf, during the study period 1998–2014, population dynamics of Sardinella aurita, and oceanic primary producers, may be controlled, mainly by top-down trophic interactions. Finally, based on the predictive models constructed here, we discuss how they can provide powerful tools to support evaluation and monitoring of fishing activity, which may help towards the development of a Fisheries Information and Management System. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Comparison of <span class="html-italic">Sardinella aurita</span> landings (tons) in the fishing zone of Abidjan produced by the Ministry of Production of Animals (MPA) at the Abidjan Fisheries Direction with the corresponding landings for all Ivorian waters produced by the Food and Agriculture Organization (FAO) for the period 1997–2014.</p>
Full article ">Figure 2
<p>Spatial and seasonal variations of (<b>a</b>) Chlorophyll (mg m<sup>−3</sup>), (<b>b</b>) SST (°C), (<b>c</b>) upwelling index (m<sup>3</sup> s<sup>−1</sup> km<sup>−1</sup>), and (<b>d</b>) wind-induced turbulent mixing (m<sup>3</sup> s<sup>−3</sup>). Left panel: mean for 1998–2014. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan. Right panel: box area climatological monthly means for 1998–2014 with standard error (gray shading).</p>
Full article ">Figure 3
<p>Seasonal variations of (<b>a</b>) thermocline depth (m), (<b>b</b>) surface temperature (°C), (<b>c</b>) nitracline depth (m), (<b>d</b>) nitracline depth-integrated nitrate concentration (mol m<sup>−2</sup>), (<b>e</b>) phosphatocline depth (m), and (<b>f</b>) phosphatocline depth-integrated phosphate concentration (mol m<sup>−2</sup>). Left panel: blue line is upper bound of the layer and turquoise line is lower bound of the layer. All observations are based on in situ measurements collected in Abidjan fishing zone at the point of coordinate 5.5°W, 4.5°N. Water column oceanographic profiles are from the World Ocean Atlas 2013 [<a href="#B56-remotesensing-10-00785" class="html-bibr">56</a>].</p>
Full article ">Figure 4
<p>Maps of relative anomalies in annual mean chlorophyll concentration expressed in percentage in Ivorian waters. Chlorophyll concentration anomalies are calculated using the Ocean Color Climate Change Initiative (OC-CCI) product for the period 1998 to 2014. Blue (red) color indicates lower (higher) chlorophyll concentration relative to the 17 year mean. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure 5
<p>Maps of anomalies of timing of initiation (8-Day period) of the phytoplankton growth in Ivorian waters. The timing of initiation is calculated using the OC-CCI chlorophyll product and applying the threshold method presented in Racault et al. [<a href="#B57-remotesensing-10-00785" class="html-bibr">57</a>]. Anomalies are estimated for the period 1998 to 2014. Blue (red) color indicates earlier (later) phytoplankton growth initiation compared with the 17 year mean. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure 6
<p>Interannual variability during the period 1998 to 2014 in (<b>a</b>) <span class="html-italic">Sardinella aurita</span> catch (tons) in year <span class="html-italic">t +</span> 1, and timing of initiation (weeks) in year <span class="html-italic">t</span>; and (<b>b</b>) <span class="html-italic">Sardinella aurita</span> catch (tons) in year <span class="html-italic">t</span> and chlorophyll concentration (mg m<sup>−3</sup>) in year <span class="html-italic">t</span>. Note that in (<b>a</b>), the relationship is shown for phytoplankton bloom timing in year <span class="html-italic">t</span> and <span class="html-italic">S. aurita</span> catch in year <span class="html-italic">t +</span> 1, as we are investigating the influence of the bloom timing (food availability) on the <span class="html-italic">S. aurita</span> recruitment success, which is shown in the catch success in following year. On the other hand, in (<b>b</b>), the relationship is shown for chlorophyll concentration and <span class="html-italic">S. aurita</span> catch both in the same year, <span class="html-italic">t</span>, as we are investigating grazing pressure of <span class="html-italic">S. aurita</span> on chlorophyll concentration in the same year. All data were averaged over the area of the fishing zone of Abidjan (please see box-area location in <a href="#remotesensing-10-00785-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 7
<p>Maps of upwelling index anomalies (m<sup>3</sup> s<sup>−1</sup> km<sup>−1</sup>). Anomalies are estimated for the period 1998 to 2014. Blue (red) color indicates stronger (weaker) upwelling conditions compared to the 17 year mean. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure 8
<p>Maps of wind-induced turbulent mixing anomalies (m<sup>3</sup> s<sup>−3</sup>). Anomalies are estimated for the period 1998 to 2014. Blue (red) color indicates stronger (weaker) turbulent conditions compared with the 17 year mean. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure 9
<p>Interannual variability during the period 1998 to 2014 in (<b>a</b>) <span class="html-italic">Sardinella aurita</span> catch (tons) in year <span class="html-italic">t +</span> 1 and upwelling index (annual mean in m<sup>3</sup> s<sup>−1</sup> km<sup>−1</sup>) in year <span class="html-italic">t</span>; and (<b>b</b>) <span class="html-italic">Sardinella aurita</span> catch (in tons) in year <span class="html-italic">t +</span> 1 and wind-induced turbulent mixing (mean June–December in m<sup>3</sup> s<sup>−3</sup>) in year <span class="html-italic">t</span>. Note that in (a), the relationship is shown for upwelling index and wind-induced turbulent mixing in year <span class="html-italic">t</span>, and <span class="html-italic">S. aurita</span> catch in year <span class="html-italic">t +</span> 1, as we are investigating the influence of the ocean physics on the <span class="html-italic">S. aurita</span> recruitment success, which is shown in the catch success in following year. All data were averaged over the area of the fishing zone of Abidjan (please see box-area location in <a href="#remotesensing-10-00785-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 10
<p>Plots of empirical relationships between interannual variations in <span class="html-italic">Sardinella aurita</span> catch, phytoplankton chlorophyll, phytoplankton timing of initiation, turbulent mixing, and upwelling index. Annual mean chlorophyll concentration in mg m<sup>−3</sup>; timing of initiation of phytoplankton growth in days; June to December mean wind-induced turbulent mixing in m<sup>3</sup> s<sup>−3</sup>; and annual mean upwelling index in m<sup>3</sup> s<sup>−1</sup> km<sup>−1</sup>. Note that in (<b>a</b>), the relationship is shown for chlorophyll concentration and <span class="html-italic">S. aurita</span> catch both in the same year <span class="html-italic">t</span>, as we are investigating grazing pressure on chlorophyll concentration in that year. Similarly, in (<b>c</b>–<b>f</b>) the relationships are shown for variables in the year <span class="html-italic">t</span>, as we are investigating the influence of ocean physics on the concentration and timing of the first trophic level in the same year, and in (<b>i</b>,<b>j</b>), as we are investigating autocorrelation between physical variables and between biological variables. However, in (<b>b</b>,<b>g</b>,<b>h</b>), the relationship is shown for biophysical variables in year <span class="html-italic">t</span> and <span class="html-italic">S. aurita</span> catch in year <span class="html-italic">t +</span> 1, as we are investigating the influence of the biophysical variables on the <span class="html-italic">S. aurita</span> catch success in following year. All data were averaged over the area of the fishing zone of Abidjan (please see box-area location in <a href="#remotesensing-10-00785-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 11
<p>Comprehensive assessment of predictive performance of diagnostic models using all possible combinations of input variables. (<b>a</b>–<b>o</b>) Regression analyses between observed <span class="html-italic">Sardinella aurita</span> catch and predicted <span class="html-italic">Sardinella aurita</span> catch based on combinations of four predictive variables, i.e., anomalies of timing of initiation, annual mean chlorophyll concentration, annual mean upwelling index, and June to December mean wind-induced turbulent mixing in year <span class="html-italic">t</span>. The regression statistics are reported using adjusted <span class="html-italic">R</span><sup>2</sup>, which increases only if added variables are relevant ones. Significances are calculated using <span class="html-italic">F</span>-test. All data were averaged over the area of the fishing zone of Abidjan (please see box-area location in <a href="#remotesensing-10-00785-f002" class="html-fig">Figure 2</a>).</p>
Full article ">Figure 12
<p>Conceptual diagram for the year-to-year variations in <span class="html-italic">Sardinella aurita</span> catch in relation to first trophic level and ocean physics along the Ivorian continental shelf. (<b>a</b>) Year when <span class="html-italic">Sardinella</span> catch is high; (<b>b</b>) Year when <span class="html-italic">Sardinella</span> catch is low. Lowercase letters <b>a</b>–<b>j</b> refer to the empirical relationships presented in <a href="#remotesensing-10-00785-f010" class="html-fig">Figure 10</a>.</p>
Full article ">Figure A1
<p>Maps of annual mean chlorophyll concentration in mg Chl m<sup>−3</sup> for the years 1998 to 2014. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure A2
<p>Maps of timing of initiation of phytoplankton growth in days for the years 1998 to 2014. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure A3
<p>Maps of annual mean upwelling index in m<sup>3</sup> s<sup>−1</sup> m<sup>−1</sup> for the years 1998 to 2014. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">Figure A4
<p>Maps of annual mean wind-induced turbulent mixing in m<sup>3</sup> s<sup>−3</sup> for the years 1998 to 2014. The boxed area against the coast indicates the geographical location of the fishing zone of Abidjan.</p>
Full article ">
21 pages, 8863 KiB  
Article
Machine Learning Automatic Model Selection Algorithm for Oceanic Chlorophyll-a Content Retrieval
by Katalin Blix and Torbjørn Eltoft
Remote Sens. 2018, 10(5), 775; https://doi.org/10.3390/rs10050775 - 17 May 2018
Cited by 38 | Viewed by 6255
Abstract
Ocean Color remote sensing has a great importance in monitoring of aquatic environments. The number of optical imaging sensors onboard satellites has been increasing in the past decades, allowing to retrieve information about various water quality parameters of the world’s oceans and inland [...] Read more.
Ocean Color remote sensing has a great importance in monitoring of aquatic environments. The number of optical imaging sensors onboard satellites has been increasing in the past decades, allowing to retrieve information about various water quality parameters of the world’s oceans and inland waters. This is done by using various regression algorithms to retrieve water quality parameters from remotely sensed multi-spectral data for the given sensor and environment. There is a great number of such algorithms for estimating water quality parameters with different performances. Hence, choosing the most suitable model for a given purpose can be challenging. This is especially the fact for optically complex aquatic environments. In this paper, we present a concept to an Automatic Model Selection Algorithm (AMSA) aiming at determining the best model for a given matchup dataset. AMSA automatically chooses between regression models to estimate the parameter in interest. AMSA also determines the number and combination of features to use in order to obtain the best model. We show how AMSA can be built for a certain application. The example AMSA we present here is designed to estimate oceanic Chlorophyll-a for global and optically complex waters by using four Machine Learning (ML) feature ranking methods and three ML regression models. We use a synthetic and two real matchup datasets to find the best models. Finally, we use two images from optically complex waters to illustrate the predictive power of the best models. Our results indicate that AMSA has a great potential to be used for operational purposes. It can be a useful objective tool for finding the most suitable model for a given sensor, water quality parameter and environment. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The feature ranking stage of the AMSA.</p>
Full article ">Figure 2
<p>Regression stage of the AMSA (A).</p>
Full article ">Figure 3
<p>Regression stage of the AMSA (B).</p>
Full article ">Figure 4
<p>The ML AMSA for oceanic Chl-a content estiamtion.</p>
Full article ">Figure 5
<p>Illustration of the AMSA for application.</p>
Full article ">Figure 6
<p>Position of the data for the MERIS (<b>left</b>) and MODIS-Aqua (<b>right</b>) global dataset. The red and black markers indicate oligotrophic and eutrophic conditions, respectively.</p>
Full article ">Figure 7
<p>Estimated Chl-a map for the coast of East USA by using the GPR (<b>left-column</b>) and SVR (<b>right-column</b>) model with bands centered at 510, 560 and 620 nm. The bottom row shows the enlarged area indicated by the red squares.</p>
Full article ">Figure 8
<p>Estimated Chl-a map for the southern Baltic sea by using the GPR (<b>left-column</b>) and SVR (<b>right-column</b>) model with bands centered at 510, 560 and 620 nm. The bottom row shows the enlarged area indicated by the red squares.</p>
Full article ">
21 pages, 3380 KiB  
Article
A Statistical Modeling Framework for Characterising Uncertainty in Large Datasets: Application to Ocean Colour
by Peter E. Land, Trevor C. Bailey, Malcolm Taberner, Silvia Pardo, Shubha Sathyendranath, Kayvan Nejabati Zenouz, Vicki Brammall, Jamie D. Shutler and Graham D. Quartly
Remote Sens. 2018, 10(5), 695; https://doi.org/10.3390/rs10050695 - 2 May 2018
Cited by 4 | Viewed by 6005
Abstract
Uncertainty estimation is crucial to establishing confidence in any data analysis, and this is especially true for Essential Climate Variables, including ocean colour. Methods for deriving uncertainty vary greatly across data types, so a generic statistics-based approach applicable to multiple data types is [...] Read more.
Uncertainty estimation is crucial to establishing confidence in any data analysis, and this is especially true for Essential Climate Variables, including ocean colour. Methods for deriving uncertainty vary greatly across data types, so a generic statistics-based approach applicable to multiple data types is an advantage to simplify the use and understanding of uncertainty data. Progress towards rigorous uncertainty analysis of ocean colour has been slow, in part because of the complexity of ocean colour processing. Here, we present a general approach to uncertainty characterisation, using a database of satellite-in situ matchups to generate a statistical model of satellite uncertainty as a function of its contributing variables. With an example NASA MODIS-Aqua chlorophyll-a matchups database mostly covering the north Atlantic, we demonstrate a model that explains 67% of the squared error in log(chlorophyll-a) as a potentially correctable bias, with the remaining uncertainty being characterised as standard deviation and standard error at each pixel. The method is quite general, depending only on the existence of a suitable database of matchups or reference values, and can be applied to other sensors and data types such as other satellite observed Essential Climate Variables, empirical algorithms derived from in situ data, or even model data. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The distribution of data used in the matchups database. The map shows the number of matchups in each 1° × 1° cell.</p>
Full article ">Figure 2
<p>A simple uncertainty model in which ln(chl<sub>SAT</sub>) is the only explanatory variable. (<b>a</b>) The points are ln(chl<sub>IS</sub>) plotted against ln(chl<sub>SAT</sub>), and the black line is the 1:1 line. The blue line shows the best fitting pb(ln(chl<sub>SAT</sub>)) mean model, and the solid red line shows the best fitting ps(ln(chl<sub>SAT</sub>)) mean model. The remaining lines are offset above and below the latter. The solid pink line is offset by the standard error, the dashed pink line (almost overlapping with the dashed red line) by the best fitting ps(ln(chl<sub>SAT</sub>)) standard deviation model, and the dashed red line by the square root of the sum of squares of standard error and standard deviation. (<b>b</b>) Contributions to absolute uncertainty using the ps(ln(chl<sub>SAT</sub>)) mean model. The points are the absolute difference between ln(chl<sub>SAT</sub>) and ln(chl<sub>IS</sub>); the solid red line is the absolute bias; the solid pink line is the standard error; the dashed pink line is the standard deviation; the dashed red line is the square root of the sum of squares of standard error and standard deviation; and the solid black line is the square root of the sum of squares of bias, standard error, and standard deviation.</p>
Full article ">Figure 3
<p>Dependencies of the best fitting model. Each graph shows the dependency of a model parameter (mean or standard deviation) on a single explanatory variable (the ‘partial dependency’), all others being held constant. The red line is the model prediction, the pink lines are one standard error either side of this, and the pale blue circles are the prediction plus the model residual at each data point. (<b>a</b>) Mean model dependencies, in the order that they were added by gamlss: (top row) pb(ln(chlSAT)), ps(day length), and ps(<span class="html-italic">R<sub>rs</sub></span>(412)); (second row) pbc(day of year), ps(satellite age), and ps(<span class="html-italic">R<sub>rs</sub></span>(469)); (third row) ps(<span class="html-italic">R<sub>rs</sub></span>(531)), ps(time difference), ps(airmass); (bottom row) ps(1/cos(view zenith angle)), ps(<span class="html-italic">R<sub>rs</sub></span>(547)), and ps(<span class="html-italic">R<sub>rs</sub></span>(555)). All graphs have the same scale on the vertical axis. (<b>b</b>) Dependency of ln(standard deviation) on ps(<span class="html-italic">R<sub>rs</sub></span>(645)).</p>
Full article ">Figure 4
<p>(<b>a</b>) Chl<sub>SAT</sub> (mg m<sup>−3</sup>) from a MODIS-Aqua overpass on 5 May 2013 at 12:10 (original satellite projection at ~1 km resolution, central section removed); (<b>b</b>) overall uncertainty in chl<sub>SAT</sub>, chl<sub>SAT</sub> × <math display="inline"><semantics> <mo>√</mo> </semantics></math>(bias<sup>2</sup> + standard deviation<sup>2</sup> + standard error<sup>2</sup>), estimated using GAMLSS; (<b>c</b>) chl<sub>SAT</sub> with bias subtracted; (<b>d</b>) uncertainty in bias-subtracted chl<sub>SAT</sub>, chl<sub>SAT</sub> × <math display="inline"><semantics> <mo>√</mo> </semantics></math>(standard deviation<sup>2</sup> + standard error<sup>2</sup>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Mean δln(chl), or bias; (<b>b</b>) standard deviation; (<b>c</b>) standard error.</p>
Full article ">Figure 6
<p>(<b>a</b>) Unweighted composite of chl from 1 to 8 May 2013 using the mean of ln(chl), shown at the top of the scale bar in mg m<sup>−3</sup>; (<b>b</b>) number of chl values contributing to each pixel, shown at the bottom of the scale bar.</p>
Full article ">Figure 7
<p>Weighted composites and their uncertainty. (<b>a</b>) Uncorrected weighted composite of chl in mg m<sup>−3</sup> from 1 to 8 May 2013 using the mean of ln(chl); (<b>b</b>) uncertainty in the uncorrected composite due to bias, standard deviation, standard error, and estimated natural variability of ln(chl); (<b>c</b>) bias corrected weighted chl composite; (<b>d</b>) uncertainty in the corrected composite due to standard deviation, standard error, and estimated natural variability of ln(chl).</p>
Full article ">Figure 7 Cont.
<p>Weighted composites and their uncertainty. (<b>a</b>) Uncorrected weighted composite of chl in mg m<sup>−3</sup> from 1 to 8 May 2013 using the mean of ln(chl); (<b>b</b>) uncertainty in the uncorrected composite due to bias, standard deviation, standard error, and estimated natural variability of ln(chl); (<b>c</b>) bias corrected weighted chl composite; (<b>d</b>) uncertainty in the corrected composite due to standard deviation, standard error, and estimated natural variability of ln(chl).</p>
Full article ">Figure 8
<p>Natural chl variation used to account for variability greater than the uncertainty predicted by the model. Light grey regions have fewer than two measurements, white regions are zero.</p>
Full article ">Figure A1
<p>RGB composites of absolute bias (red), SD (green), and SE (blue), scaled from zero to one. Light grey is missing data, white is bias, SD, and SE all greater than one. (<b>a</b>) 5 May 2013 at 12:10, mapped to the composite grid; (<b>b</b>) mean composite from 1 to 8 May 2013.</p>
Full article ">
20 pages, 29419 KiB  
Article
Long-Term Changes in Colored Dissolved Organic Matter from Satellite Observations in the Bohai Sea and North Yellow Sea
by Cong Xiao, Deyong Sun, Shengqiang Wang, Zhongfeng Qiu, Yu Huan and Jiabao Zhang
Remote Sens. 2018, 10(5), 688; https://doi.org/10.3390/rs10050688 - 29 Apr 2018
Cited by 4 | Viewed by 4608
Abstract
Spatial and temporal variations in colored dissolved organic matter (CDOM) are of great importance to understanding the dynamics of the biogeochemical properties of water bodies. This study proposed a remote sensing approach for estimating CDOM concentrations (CCDOM) based on in [...] Read more.
Spatial and temporal variations in colored dissolved organic matter (CDOM) are of great importance to understanding the dynamics of the biogeochemical properties of water bodies. This study proposed a remote sensing approach for estimating CDOM concentrations (CCDOM) based on in situ observations from the Bohai Sea (BS) and the North Yellow Sea (NYS). Cross-validation demonstrated that the accuracy of the CDOM algorithm is R2 = 0.78, APD = 15.9%, RMSE = 0.92 (ppb). The CDOM algorithm was applied to estimate the 14-year (2003–2016) sea surface CCDOM in the BS and NYS areas using Moderate Resolution Imaging Spectroradiometer (MODIS) monthly products. The results showed a significant fluctuation in CDOM variations on a long-term scale. The highest values of CDOM were observed in the BS, the middle values were observed in the Bohai Strait, and the lowest values were observed in the NYS. Seasonal variations were observed with long-lasting high CDOM values from June to August in coastal waters, while relatively low values were observed in the NYS in the summer. In the spring and fall, a distinct increase appeared in the NYS. High CDOM values in the nearshore coastal waters were mostly related to terrestrial inputs, while CDOM in the offshore regions was mainly due to autochthonous production. Furthermore, ocean currents played an important role in the variations in CDOM in the BS and NYS areas, especially for variations in CDOM in the Bohai Strait. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Study area and sampling stations. Red stations indicate those with in situ colored dissolved organic matter (CDOM) measurements coinciding with the in situ remote sensing reflectance (Rrs) measurements that were used for the algorithm development. Black stations indicate those where the in situ CDOM measurements coincide with the salinity measurements.</p>
Full article ">Figure 2
<p>In situ measured Rrs in a spectral range of 400–700 nm.</p>
Full article ">Figure 3
<p>Relationships between CDOM and salinity during different cruises in the study area (see regression equations in <a href="#remotesensing-10-00688-t001" class="html-table">Table 1</a>).</p>
Full article ">Figure 4
<p>Comparison between measured and modeled CDOM. (<b>a</b>) CDOM derived using in situ measured Rrs data; (<b>b</b>) CDOM derived using MODIS data.</p>
Full article ">Figure 5
<p>Validation of absorption coefficient of CDOM (<span class="html-italic">a</span><sub>CDOM</sub>) estimated by using Carder et al.’s (2003) and Tassan et al.’s (1994) models, based on our in situ measured Rrs data.</p>
Full article ">Figure 6
<p>Annually averaged CDOM distributions derived from the MODISA monthly data (2003–2016). Note that the first subfigure shows the CDOM distribution averaged by the 14-year annual CDOM data. Red boxes refer to six subareas for the detailed description of the CDOM spatial and temporal distributions (A: the Liaodong Bay; B: the central of the Bohai Sea; C: the Bohai Bay; D: the Laizhou Bay; E: the Bohai Strait; F: the central of the North Yellow Sea).</p>
Full article ">Figure 7
<p>CDOM variations in six subareas derived from the MODISA data. (<b>a</b>) Annual CDOM averaged by the 12-month results of each year; (<b>b</b>) Monthly mean CDOM averaged among specific months over 14 years (2003–2016). A: the Liaodong Bay; B: the central of the Bohai Sea; C: the Bohai Bay; D: the Laizhou Bay; E: the Bohai Strait; F: the central North Yellow Sea. The Y-axis on the left represents the values of A, B, C, D, E, and the Y-axis on the right represents the values of F.</p>
Full article ">Figure 8
<p>Monthly CDOM distributions derived from the MODISA data averaged among specific months over 14 years (2003–2016).</p>
Full article ">Figure 9
<p>Spatial and monthly CDOM variations at two transects (S1: 118.5°E–124°E, 38.5°N; S2: 118°E–124°E, 38.5°N; see the first subfigure in <a href="#remotesensing-10-00688-f008" class="html-fig">Figure 8</a>).</p>
Full article ">Figure 10
<p>Distributions of in situ-measured salinity, CDOM and Chla values collected from three cruises (Aug. 2015, Jul. 2016, and Jan. 2017).</p>
Full article ">Figure 11
<p>Distribution of eight-year (2009–2016) salinity averages for specific seasons obtained from HYCOM (HYbrid Coordinate Ocean Model) and the eight-year CDOM and Chla averages for specific seasons from MODISA. The HYCOM data is available from <a href="http://hycom.org/" target="_blank">http://hycom.org/</a>. Considering that the HYCOM data range is from September 2008 to the present, we focused on the period of 2009–2016. Detailed information about the HYCOM validation and parameter sets can be seen in [<a href="#B66-remotesensing-10-00688" class="html-bibr">66</a>,<a href="#B67-remotesensing-10-00688" class="html-bibr">67</a>] studies.</p>
Full article ">Figure 12
<p>Seasonal variation in 14-year averaged monthly CDOM (ppb) and Chla (mg m<sup>−3</sup>) from January to December in the six subareas: A: the Liaodong Bay; B: the central Bohai Sea; C: the Bohai Bay; D: the Laizhou Bay; E: the Bohai Strait; F: the central North Yellow Sea.</p>
Full article ">Figure 13
<p>Distribution of two main water masses (i.e., Bohai Sea Central Water (BSCW) and Bohai Sea Coastal Water (BSCoW)) during the different seasons in our study area.</p>
Full article ">
12 pages, 3405 KiB  
Article
Influence of Tropical Instability Waves on Phytoplankton Biomass near the Marquesas Islands
by Elodie Martinez, Hirohiti Raapoto, Christophe Maes and Keitapu Maamaatuaihutapu
Remote Sens. 2018, 10(4), 640; https://doi.org/10.3390/rs10040640 - 20 Apr 2018
Cited by 8 | Viewed by 4826
Abstract
The Marquesas form an isolated group of small islands in the Central South Pacific where quasi-permanent biological activity is observed. During La Niña events, this biological activity, shown by a net increase of chlorophyll-a concentration (Chl, a proxy of phytoplankton biomass), is particularly [...] Read more.
The Marquesas form an isolated group of small islands in the Central South Pacific where quasi-permanent biological activity is observed. During La Niña events, this biological activity, shown by a net increase of chlorophyll-a concentration (Chl, a proxy of phytoplankton biomass), is particularly strong. It has been hypothesized that this strong activity is due to iron-rich waters advected from the equatorial region to the Marquesas by tropical instability waves (TIWs). Here we investigate this hypothesis over 18 years by combining satellite observations, re-analyses of ocean data, and Lagrangian diagnostics. Four La Niña events ranging from moderate to strong intensity occurred during this period, and our results show that the Chl plume within the archipelago can be indeed influenced by such equatorial advection, but this was observed during the strong 1998 and 2010 La Niña conditions only. Chl spatio-temporal patterns during the occurrence of other TIWs rather suggest the interaction of large-scale forcing events such as an uplift of the thermocline or the enhancement of coastal upwelling induced by the tropical strengthening of the trades with the islands leading to enhancement of phytoplankton biomass within the surface waters. Overall, whatever the conditions, our analyses suggest that the influence of the TIWs is to disperse, stir, and, therefore, modulate the shape of the existing phytoplankton plume. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The 1998–2014 annual average of chlorophyll-a concentrations (Chl, mg/m<sup>3</sup>) from the satellite-derived GlobColour Chl AVE product. The purple line delineates the French Polynesian Exclusive Economic Zone (EEZ).</p>
Full article ">Figure 2
<p>Annual average conditions calculated over the whole-time period for the Marquesas physical and biological environment. (<b>a</b>) Finite-size lyapunov exponents (FSLEs) (d<sup>−1</sup>); (<b>b</b>) Sea surface temperature (SST) (°C); (<b>c</b>) surface sigma (kg/m<sup>3</sup>); (<b>d</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s). The islands are shown in black.</p>
Full article ">Figure 3
<p>Root mean square (RMS) of FLSEs in 1998. The black box delineates the area over which data are averaged to provide time-series of Figure 6. The islands are shown in black.</p>
Full article ">Figure 4
<p>(<b>Left</b>) SST (°C) and FSLEs (d<sup>−1</sup>, isocontours are plotted from 0.07 to 1 every 0.1); (<b>Centre</b>) surface sigma (kg/m<sup>3</sup>); (<b>Right</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s) on 6, 14, and 22 September 1998 (top to bottom, respectively). The islands are shown in black.</p>
Full article ">Figure 5
<p>Annual RMS of FLSE for (<b>a</b>) a neutral year (as in 2003); and during La Niña events as in (<b>b</b>) 1999; (<b>c</b>) 2007; and (<b>d</b>) 2010.</p>
Full article ">Figure 6
<p>Time series of (<b>a</b>) El Niño 3–4 index provided by the Climate Prediction Center (CPC)/National Centers for Environmental Prediction (NCEP) services (the y-axis is inverted). Values lower than the −1 threshold (dash line) highlight moderate to strong La Niña years; (<b>b</b>) FSLE (d<sup>−1</sup>; blue line and left axis) and SST (°C; black line and right axis) monthly anomalies averaged over the Marquesas archipelago (11°S–18°S/142°W–138°W); (<b>c</b>) Number of particles launched from the northeastern equatorial area and reaching the Marquesas after 40, 60, and 90 days of drift (red, blue, and black lines, respectively); (<b>d</b>) Chl monthly anomalies (mg/m<sup>3</sup>) averaged over the same area as in (<b>b</b>).</p>
Full article ">Figure 7
<p>(<b>Left</b>) SST (°C) and FSLEs (d<sup>−1</sup>, isocontours are plotted from 0.07 to 1 every 0.1); (<b>Centre</b>) surface sigma (kg/m<sup>3</sup>); (<b>Right</b>) Chl (mg/m<sup>3</sup>) and surface current (m/s) during the five moderate to strong La Niña events over 1997–2014 (top to bottom). The islands are shown in black.</p>
Full article ">Figure 8
<p>Annual mean of eddy kinetic energy (EKE; cm²/s<sup>2</sup>) issued from (<b>a</b>) the Geostrophic and Ekman Current Observatory (GECKO) climatology with a ¼° spatial resolution [<a href="#B36-remotesensing-10-00640" class="html-bibr">36</a>] and (<b>b</b>) a climatological 1/45° resolution simulation from the Regional Ocean Modeling System (ROMS model) (see [<a href="#B37-remotesensing-10-00640" class="html-bibr">37</a>]); (<b>c</b>) Daily vorticity field (10<sup>−5</sup>/s<sup>1</sup>) at 10 m for 14 June of Year 6 from the ROMS climatological simulation; (<b>d</b>) Chl (mg/m<sup>3</sup>) for 20 July 2006, from the satellite-derived Chl AVE GlobColour product.</p>
Full article ">
17 pages, 3717 KiB  
Article
Comparison of Satellite-Derived Phytoplankton Size Classes Using In-Situ Measurements in the South China Sea
by Shuibo Hu, Wen Zhou, Guifen Wang, Wenxi Cao, Zhantang Xu, Huizeng Liu, Guofeng Wu and Wenjing Zhao
Remote Sens. 2018, 10(4), 526; https://doi.org/10.3390/rs10040526 - 27 Mar 2018
Cited by 16 | Viewed by 5152
Abstract
Ocean colour remote sensing is used as a tool to detect phytoplankton size classes (PSCs). In this study, the Medium Resolution Imaging Spectrometer (MERIS), Moderate Resolution Imaging Spectroradiometer (MODIS), and Sea-viewing Wide Field-of-view Sensor (SeaWiFS) phytoplankton size classes (PSCs) products were compared with [...] Read more.
Ocean colour remote sensing is used as a tool to detect phytoplankton size classes (PSCs). In this study, the Medium Resolution Imaging Spectrometer (MERIS), Moderate Resolution Imaging Spectroradiometer (MODIS), and Sea-viewing Wide Field-of-view Sensor (SeaWiFS) phytoplankton size classes (PSCs) products were compared with in-situ High Performance Liquid Chromatography (HPLC) data for the South China Sea (SCS), collected from August 2006 to September 2011. Four algorithms were evaluated to determine their ability to detect three phytoplankton size classes. Chlorophyll-a (Chl-a) and absorption spectra of phytoplankton (aph(λ)) were also measured to help understand PSC’s algorithm performance. Results show that the three abundance-based approaches performed better than the inherent optical property (IOP)-based approach in the SCS. The size detection of microplankton and picoplankton was generally better than that of nanoplankton. A three-component model was recommended to produce maps of surface PSCs in the SCS. For the IOP-based approach, satellite retrievals of inherent optical properties and the PSCs algorithm both have impacts on inversion accuracy. However, for abundance-based approaches, the selection of the PSCs algorithm seems to be more critical, owing to low uncertainty in satellite Chl-a input data Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Map of the study area and locations of the in situ water sampling sites in the South China Sea.</p>
Full article ">Figure 2
<p>Maps of the matching HPLC data sets for MERIS, MODIS, and SeaWiFS.</p>
Full article ">Figure 3
<p>Histogram of the in situ HPLC data sets for the South China Sea (SCS).</p>
Full article ">Figure 4
<p>Scatter plots of estimated phytoplankton size classes (PSCs) versus in situ measurements: Uitz2006 (<b>a</b>–<b>c</b>); Brewin2010 (<b>d</b>–<b>f</b>); Hirata2011 (<b>g</b>–<b>i</b>); Roy2013 (<b>j</b>–<b>l</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">Figure 4 Cont.
<p>Scatter plots of estimated phytoplankton size classes (PSCs) versus in situ measurements: Uitz2006 (<b>a</b>–<b>c</b>); Brewin2010 (<b>d</b>–<b>f</b>); Hirata2011 (<b>g</b>–<b>i</b>); Roy2013 (<b>j</b>–<b>l</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">Figure 5
<p>Scatter plots of satellite-derived PSCs versus in situ measurements: Uitz2006 (<b>a</b>–<b>c</b>); Brewin2010 (<b>d</b>–<b>f</b>); Hirata2011 (<b>g</b>–<b>i</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">Figure 6
<p>Scatter plots of satellite-derived (Roy2013) PSCs versus in situ measurements: a<sub>ph</sub>(676) was estimated by the Garver–Siegel–Maritorena (GSM) model (<b>a</b>–<b>c</b>); a<sub>ph</sub>(676) was estimated by the Generalized Inherent Optical Property (GIOP) model (<b>d</b>–<b>f</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">Figure 7
<p>Scatter plots of satellite-derived values versus in situ measurements in the SCS: GSM derived a<sub>ph</sub>(676) (<b>a</b>); GIOP-derived a<sub>ph</sub>(676) (<b>b</b>); Chl-a (<b>c</b>); GIOP-based a * <sub>ph</sub>(676) (<b>d</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">Figure 7 Cont.
<p>Scatter plots of satellite-derived values versus in situ measurements in the SCS: GSM derived a<sub>ph</sub>(676) (<b>a</b>); GIOP-derived a<sub>ph</sub>(676) (<b>b</b>); Chl-a (<b>c</b>); GIOP-based a * <sub>ph</sub>(676) (<b>d</b>). The solid line is the 1:1 line. The red dashed lines are the 1:2 and 2:1 lines.</p>
Full article ">
25 pages, 7834 KiB  
Article
Improving the Remote Sensing Retrieval of Phytoplankton Functional Types (PFT) Using Empirical Orthogonal Functions: A Case Study in a Coastal Upwelling Region
by Marco Correa-Ramirez, Carmen E. Morales, Ricardo Letelier, Valeria Anabalón and Samuel Hormazabal
Remote Sens. 2018, 10(4), 498; https://doi.org/10.3390/rs10040498 - 22 Mar 2018
Cited by 6 | Viewed by 6196
Abstract
An approach that improves the spectral-based PHYSAT method for identifying phytoplankton functional types (PFT) in satellite ocean-color imagery is developed and applied to one study case. This new approach, called PHYSTWO, relies on the assumption that the dominant effect of chlorophyll-a (Chl-a) in [...] Read more.
An approach that improves the spectral-based PHYSAT method for identifying phytoplankton functional types (PFT) in satellite ocean-color imagery is developed and applied to one study case. This new approach, called PHYSTWO, relies on the assumption that the dominant effect of chlorophyll-a (Chl-a) in the normalized water-leaving radiance (nLw) spectrum can be effectively isolated from the signal of accessory pigment biomarkers of different PFT by using Empirical Orthogonal Function (EOF) decomposition. PHYSTWO operates in the dimensionless plane composed by the first two EOF modes generated through the decomposition of a space–nLw matrix at seven wavelengths (412, 443, 469, 488, 531, 547, and 555 nm). PFT determination is performed using orthogonal models derived from the acceptable ranges of anomalies proposed by PHYSAT but adjusted with the available regional and global data. In applying PHYSTWO to study phytoplankton community structures in the coastal upwelling system off central Chile, we find that this method increases the accuracy of PFT identification, extends the application of this tool to waters with high Chl-a concentration, and significantly decreases (~60%) the undetermined retrievals when compared with PHYSAT. The improved accuracy of PHYSTWO and its applicability for the identification of new PFT are discussed. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>The mean daily chlorophyll-a (Chl-a) concentration derived from MODIS-Aqua for the dates of the PHYTOFRONT cruise (4–6 February 2014). The black lines and dots indicate the two transects and the stations sampled. The square symbol indicates the location of the time-series station St. 18.</p>
Full article ">Figure 2
<p>In situ FluoroProbe profiles during the PHYTOFRONT cruise: north and south transects (see <a href="#remotesensing-10-00498-f001" class="html-fig">Figure 1</a>). The Chl-a concentration (color scheme; mg m<sup>−3</sup>) associated to: Diatom (<b>a</b>,<b>b</b>); and Green Algae and Cryptophyta (GA+C; (<b>c</b>,<b>d</b>)); (<b>e</b>,<b>f</b>) color dissolved organic matter (CDOM) relative concentration; and (<b>g</b>,<b>h</b>) the average Chl-a concentration (black line) in surface waters (first 20 m depth) and the relative contribution of Diatom (red dashed lines) and GA+C (green dashed lines) to surface Chl-a.</p>
Full article ">Figure 3
<p>(<b>a</b>) An empirical reference model nLw<sup>ref</sup> of PHYSAT [<a href="#B22-remotesensing-10-00498" class="html-bibr">22</a>] with the mean nLw radiances at 300 Chl-a concentrations (range: 0.1–3 mg·m<sup>−3</sup>, every 0.1 mg·m<sup>−3</sup>); and (<b>b</b>) a regional reference model nLw<sup>upw</sup> with the mean nLw radiances at 1490 Chl-a concentrations (range: 0.1–15 mg·m<sup>−3</sup>), based on the MODIS-A nLw data from the coastal upwelling region off central Chile (35–38°S and 72–76°W) during the upwelling season (January–March) of 2014. The Rrs retrievals where nLw(555) was &gt;1.3 W·m<sup>−2</sup>·μm<sup>−1</sup>·sr<sup>−1</sup> and the aerosol optical thickness (AOT) was &gt;0.15 were excluded to reduce the presence of biased values arising from high concentrations of suspended sediments or errors in the atmospheric correction.</p>
Full article ">Figure 4
<p>The acceptable PHYSAT ranges of the nLw anomalies (dashed lines) for each phytoplankton functional type (PFT) [<a href="#B24-remotesensing-10-00498" class="html-bibr">24</a>]. The continuous bold lines represent the mean values for each group. Abbreviations: COB: Coccolithophorids bloom; PHA: <span class="html-italic">Phaeocystis</span>-like; DIA: Diatoms; SLC: <span class="html-italic">Synechococcus</span>; PRO: <span class="html-italic">Prochlorococcus</span>; and NEU: Nanoeukaryotes.</p>
Full article ">Figure 5
<p>The PFT distribution calculated from a 3-day average of MODIS-A nLw data (4–6 February 2014), using: (<b>a</b>) the PHYSAT standard reference model (nLw<sup>ref</sup>); and (<b>b</b>) the PHYSAT regional reference model (nLw<sup>upw</sup>) for the coastal upwelling and transition zones in the region off central Chile. Abbreviations: COB: Coccolithophorids bloom; PHA: <span class="html-italic">Phaeocystis</span>-like; DIA: Diatoms; SLC: <span class="html-italic">Synechococcus</span>; PRO: <span class="html-italic">Prochlorococcus</span>; and NEU: Nanoeukaryotes.</p>
Full article ">Figure 6
<p>The spatial pattern of the: first <span class="html-italic">U</span><sub>1</sub> (<b>a</b>); and second <span class="html-italic">U</span><sub>2</sub> (<b>b</b>) orthogonal modes derived from the SVD of a space–nLw matrix, composed by the average MODIS-A nLw data from the 4–6 February 2014.</p>
Full article ">Figure 7
<p>The dimensionless <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane derived from the singular value decomposition (SVD) of <span class="html-italic">NNs</span> matrix. The colored symbols represent the location of the typical values (synthetic matrix <span class="html-italic">Rs</span>) of the PFTs at different Chl-a concentrations; the black dots correspond to values derived from the observations at each pixel in the <span class="html-italic">R</span> matrix. Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 8
<p>(<b>a</b>) The dimensionless <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane derived from the joint SVD of the synthetic matrix <span class="html-italic">Rs</span> and an <span class="html-italic">R</span> matrix composed by the average MODIS-A nLw data during 4–6 February 2014. The colors represent the respective PFTs assigned to each dot (pixel) considering their closeness to the unadjusted PFT orthomodels (black dots) obtained from the acceptable ranges proposed by PHYSAT. (<b>b</b>) The unadjusted PFT estimation of PHYSTWO by rearranging the assignments of (<b>a</b>) in a longitude–latitude plane. Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 9
<p>The schematic diagram summarizing the PHYSTWO method. The dashed arrow represents the process of orthomodels adjustment and their incorporation into the synthetic matrix.</p>
Full article ">Figure 10
<p>The dimensionless <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane obtained from the joint SVD of the synthetic matrix <span class="html-italic">Rs</span> and an <span class="html-italic">R</span> matrix composed by the matchups of nLw data corresponding to in situ measurements performed during the PHYTOFRONT cruise and in St. 18. (<b>a</b>) The relative concentration (percent of total Chl-a concentration) of Diatoms (green circles) and Nanoeukaryotes (grey circles) derived from FluoroProbe measurements associated with Diatom and Green Algae plus Cryptophyta, respectively. In (<b>b</b>,<b>c</b>), the colored area represents the relative concentration interpolated for Diatoms and Nanoeukaryotes, respectively, with blue (red) colors corresponding to low (high) values. The orange dots in (<b>b</b>) denote the new addition to the orthomodel for Diatoms (yellow line); the light gray dots in (<b>c</b>) denote the points excluded in the orthomodel for Nanoeukaryotes (grey line). Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>; F-coding refers to the PHYTOFRONT stations in <a href="#remotesensing-10-00498-f001" class="html-fig">Figure 1</a>.</p>
Full article ">Figure 11
<p>The relative concentration of <span class="html-italic">Phaeocystis</span> (gray tones) in the orthogonal <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane, calculated from the SVD of an <span class="html-italic">R</span> matrix composed of 141 registers from the MAREDAT database in the period between 2002 and 2009. The orthomodels for <span class="html-italic">Phaeocystis</span> and Diatoms are shown as red and yellow dots, respectively. The bright red dots show the proposed adjustment for the <span class="html-italic">Phaeocystis</span> orthomodel.</p>
Full article ">Figure 12
<p>The fitted orthogonal models (first <span class="html-italic">U</span><sub>1</sub> and second <span class="html-italic">U</span><sub>2</sub> modes) for each PFT. The section on the right is an enlargement of the square area in the section on the left. Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 13
<p>(<b>a</b>) The dimensionless <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane derived from the joint SVD of the typical values matrix <span class="html-italic">Rt</span> and an <span class="html-italic">R</span> matrix composed of the average MODIS-A nLw data during 4–6 February 2014. The colors represent the respective PFT assigned to each dot (pixel) considering their closeness to the fitted PFT orthomodels (black dots) shown in <a href="#remotesensing-10-00498-f012" class="html-fig">Figure 12</a>. (<b>b</b>) The adjusted PFT estimation of PHYSTWO were obtained by rearranging the assignments of (<b>a</b>) in a longitude–latitude plane. (<b>c</b>) A global view of the PFT estimation performed by PHYSTWO for the same dates. Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 13 Cont.
<p>(<b>a</b>) The dimensionless <span class="html-italic">U</span><sub>1</sub><span class="html-italic">–U</span><sub>2</sub> plane derived from the joint SVD of the typical values matrix <span class="html-italic">Rt</span> and an <span class="html-italic">R</span> matrix composed of the average MODIS-A nLw data during 4–6 February 2014. The colors represent the respective PFT assigned to each dot (pixel) considering their closeness to the fitted PFT orthomodels (black dots) shown in <a href="#remotesensing-10-00498-f012" class="html-fig">Figure 12</a>. (<b>b</b>) The adjusted PFT estimation of PHYSTWO were obtained by rearranging the assignments of (<b>a</b>) in a longitude–latitude plane. (<b>c</b>) A global view of the PFT estimation performed by PHYSTWO for the same dates. Abbreviations are the same as in <a href="#remotesensing-10-00498-f005" class="html-fig">Figure 5</a>.</p>
Full article ">Figure 14
<p>The relationship between MODIS-A Chl-a and the spatial pattern of the: first (<b>a</b>); and second (<b>b</b>) modes resulting from of the SVD of a <span class="html-italic">R</span> matrix composed by MODIS-A nLw data off central Chile (35–38°S, 72–76°W) on the dates of the PHYTOFRONT cruise (4–6 February 2014). (<b>c</b>) The regional reference model (nLw<sup>M1</sup>) derived from the reconstruction of the first SVD orthogonal mode <span class="html-italic">M</span><sub>1</sub>. (<b>d</b>) The relationship between nLw<sup>M1</sup> and nLw<sup>upw</sup> for each wavelength.</p>
Full article ">Figure A1
<p>The typical nLw radiances for PFTs in environments with Chl-a concentrations in the range between 0.01 and 3 mg·m<sup>−3</sup>, contained in the <span class="html-italic">Rs</span> synthetic matrix.</p>
Full article ">
18 pages, 4181 KiB  
Article
Comparison of Machine Learning Techniques in Inferring Phytoplankton Size Classes
by Shuibo Hu, Huizeng Liu, Wenjing Zhao, Tiezhu Shi, Zhongwen Hu, Qingquan Li and Guofeng Wu
Remote Sens. 2018, 10(3), 191; https://doi.org/10.3390/rs10030191 - 8 Mar 2018
Cited by 46 | Viewed by 6342
Abstract
The size of phytoplankton not only influences its physiology, metabolic rates and marine food web, but also serves as an indicator of phytoplankton functional roles in ecological and biogeochemical processes. Therefore, some algorithms have been developed to infer the synoptic distribution of phytoplankton [...] Read more.
The size of phytoplankton not only influences its physiology, metabolic rates and marine food web, but also serves as an indicator of phytoplankton functional roles in ecological and biogeochemical processes. Therefore, some algorithms have been developed to infer the synoptic distribution of phytoplankton cell size, denoted as phytoplankton size classes (PSCs), in surface ocean waters, by the means of remotely sensed variables. This study, using the NASA bio-Optical Marine Algorithm Data set (NOMAD) high performance liquid chromatography (HPLC) database, and satellite match-ups, aimed to compare the effectiveness of modeling techniques, including partial least square (PLS), artificial neural networks (ANN), support vector machine (SVM) and random forests (RF), and feature selection techniques, including genetic algorithm (GA), successive projection algorithm (SPA) and recursive feature elimination based on support vector machine (SVM-RFE), for inferring PSCs from remote sensing data. Results showed that: (1) SVM-RFE worked better in selecting sensitive features; (2) RF performed better than PLS, ANN and SVM in calibrating PSCs retrieval models; (3) machine learning techniques produced better performance than the chlorophyll-a based three-component method; (4) sea surface temperature, wind stress, and spectral curvature derived from the remote sensing reflectance at 490, 510, and 555 nm were among the most sensitive features to PSCs; and (5) the combination of SVM-RFE feature selection techniques and random forests regression was recommended for inferring PSCs. This study demonstrated the effectiveness of machine learning techniques in selecting sensitive features and calibrating models for PSCs estimations with remote sensing. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Framework for model development.</p>
Full article ">Figure 2
<p>Scatter plots of satellite-derived versus high performance liquid chromatography (HPLC) microplankton size fractions (Fm): (<b>a</b>) random forests using features selected with SVM-RFE, (<b>b</b>) SVM using features selected with SVM-RFE, (<b>c</b>) SVM using ocean color features selected with SVM-RFE, and (<b>d</b>) three-component method. The dashed line is a 1:1 line, and the solid line is a regression line. Plot (<b>e</b>) shows the frequency distributions of their relative errors, and the numbers along the color ramp indicates the pixel density after log transformation (y = ln(x)).</p>
Full article ">Figure 3
<p>Scatter plots of satellite-derived versus high performance liquid chromatography (HPLC) nanoplankton size fractions (Fn): (<b>a</b>) random forests using features selected with SVM-RFE, (<b>b</b>) SVM using features selected with SVM-RFE, (<b>c</b>) SVM using ocean color features selected with SVM-RFE, and (<b>d</b>) three-component method. The dashed line is a 1:1 line, and the solid is a regression line. Plot (<b>e</b>) shows the frequency distributions of their relative errors, and the numbers along the color ramp indicates the pixel density after log transformation (y = ln(x)).</p>
Full article ">Figure 4
<p>Scatter plots of satellite-derived versus high performance liquid chromatography (HPLC) picoplankton size fractions (Fp): (<b>a</b>) random forests using features selected with SVM-RFE, (<b>b</b>) SVM using features selected with SVM-RFE, (<b>c</b>) SVM using ocean color features selected with SVM-RFE, and (<b>d</b>) three-component method. The dash line is 1:1 line, and the solid is regression line. Plot (<b>e</b>) shows the frequency distributions of their relative errors, and the numbers along the color ramp indicates the pixel density after log transformation (y = ln(x)).</p>
Full article ">
17 pages, 3947 KiB  
Article
Deriving Total Suspended Matter Concentration from the Near-Infrared-Based Inherent Optical Properties over Turbid Waters: A Case Study in Lake Taihu
by Wei Shi, Yunlin Zhang and Menghua Wang
Remote Sens. 2018, 10(2), 333; https://doi.org/10.3390/rs10020333 - 23 Feb 2018
Cited by 38 | Viewed by 6011
Abstract
Normalized water-leaving radiance spectra nLw(λ), particle backscattering coefficients bbp(λ) in the near-infrared (NIR) wavelengths, and total suspended matter (TSM) concentrations over turbid waters are analytically correlated. To demonstrate the use of bbp(λ [...] Read more.
Normalized water-leaving radiance spectra nLw(λ), particle backscattering coefficients bbp(λ) in the near-infrared (NIR) wavelengths, and total suspended matter (TSM) concentrations over turbid waters are analytically correlated. To demonstrate the use of bbp(λ) in the NIR wavelengths in coastal and inland waters, we used in situ optics and TSM data to develop two TSM algorithms from measurements of the Visible Infrared Imaging Radiometer Suite (VIIRS) onboard the Suomi National Polar-orbiting Partnership (SNPP) using backscattering coefficients at the two NIR bands bbp(745) and bbp(862) for Lake Taihu. The correlation coefficients between the modeled TSM concentrations from bbp(745) and bbp(862) and the in situ TSM are 0.93 and 0.92, respectively. A different in situ dataset acquired between 2012 and 2016 for Lake Taihu was used to validate the performance of the NIR TSM algorithms for VIIRS-SNPP observations. TSM concentrations derived from VIIRS-SNPP observations with these two NIR bbp(λ)-based TSM algorithms matched well with in situ TSM concentrations in Lake Taihu between 2012 and 2016. The normalized root mean square errors (NRMSEs) for the two NIR algorithms are 0.234 and 0.226, respectively. The two NIR-based TSM algorithms are used to compute the satellite-derived TSM concentrations to study the seasonal and interannual variability of the TSM concentration in Lake Taihu between 2012 and 2016. In fact, the NIR-based TSM algorithms are analytically based with minimal in situ data to tune the coefficients. They are not sensitive to the possible nLw(λ) saturation in the visible bands for highly turbid waters, and have the potential to be used for estimation of TSM concentrations in turbid waters with similar NIR nLw(λ) spectra as those in Lake Taihu. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Maps of China’s inland Lake Taihu. Locations of the in situ TSM measurements. Between 2012 and 2016 are marked as “×” in Lake Taihu.</p>
Full article ">Figure 2
<p>Scatter plots of in situ-derived <span class="html-italic">b<sub>bp</sub></span>(<span class="html-italic">λ</span>) versus in situ TSM concentration in Lake Taihu for <span class="html-italic">b<sub>bp</sub></span>(<span class="html-italic">λ</span>) at wavelengths of (<b>a</b>) 551 nm, (<b>b</b>) 671 nm, (<b>c</b>) 745 nm, and (<b>d</b>) 862 nm.</p>
Full article ">Figure 3
<p>Scatter plots for (<b>a</b>) <span class="html-italic">b<sub>bp</sub></span>(745)-derived versus in situ-measured TSM, (<b>b</b>) <span class="html-italic">b<sub>bp</sub></span>(862)-derived versus in situ-measured TSM, and (<b>c</b>) <span class="html-italic">b<sub>bp</sub></span>(862)-derived versus <span class="html-italic">b<sub>bp</sub></span>(745)-derived TSM.</p>
Full article ">Figure 4
<p>Scatter plots of (<b>a</b>) VIIRS <span class="html-italic">b<sub>bp</sub></span>(745)-derived <span class="html-italic">TSM</span><sup>(745)</sup> versus in situ-measured TSM and (<b>b</b>) VIIRS <span class="html-italic">b<sub>bp</sub></span>(862)-derived <span class="html-italic">TSM</span><sup>(862)</sup> versus in situ-measured TSM.</p>
Full article ">Figure 5
<p>Seasonal climatology <span class="html-italic">nL<sub>w</sub></span>(745) images (<b>a</b>–<b>d</b>) and <span class="html-italic">nL<sub>w</sub></span>(862) images (<b>e</b>–<b>h</b>) for spring, summer, autumn, and winter from VIIRS 2012–2016 measurements, respectively.</p>
Full article ">Figure 6
<p>Seasonal climatology <span class="html-italic">b<sub>bp</sub></span>(862) images (<b>a</b>–<b>d</b>) and <span class="html-italic">TSM</span><sup>(862)</sup> images (<b>e</b>–<b>h</b>) for spring, summer, autumn, and winter from VIIRS 2012–2016 measurements, respectively.</p>
Full article ">Figure 7
<p>VIIRS-derived yearly composite images of <span class="html-italic">b<sub>bp</sub></span>(862) (<b>a</b>–<b>e</b>) and <span class="html-italic">TSM</span><sup>(862)</sup> (<b>f</b>–<b>j</b>) in the corresponding years of 2012–2016.</p>
Full article ">Figure 8
<p>Variations of VIIRS-derived (<b>a</b>) <span class="html-italic">b<sub>bp</sub></span>(745) and <span class="html-italic">b<sub>bp</sub></span>(862), (<b>b</b>) <span class="html-italic">TSM</span><sup>(745)</sup> and <span class="html-italic">TSM</span><sup>(862)</sup> for the entirety of Lake Taihu.</p>
Full article ">
15 pages, 15256 KiB  
Article
Chlorophyll-a Concentration Retrieval in the Optically Complex Waters of the St. Lawrence Estuary and Gulf Using Principal Component Analysis
by Julien Laliberté, Pierre Larouche, Emmanuel Devred and Susanne Craig
Remote Sens. 2018, 10(2), 265; https://doi.org/10.3390/rs10020265 - 8 Feb 2018
Cited by 20 | Viewed by 7921
Abstract
Empirical methods based on band ratios to infer chlorophyll-a concentration by satellite do not perform well over the optically complex waters of the St. Lawrence Estuary and Gulf. Using a dataset of 93 match-ups, we explore an alternative method relying on empirical orthogonal [...] Read more.
Empirical methods based on band ratios to infer chlorophyll-a concentration by satellite do not perform well over the optically complex waters of the St. Lawrence Estuary and Gulf. Using a dataset of 93 match-ups, we explore an alternative method relying on empirical orthogonal functions (EOF) to develop an algorithm that relates the satellite-derived remote sensing reflectances to in situ chlorophyll-a concentration for the Sea-viewing Wide Field-of-view Sensor (SeaWiFS). Results show that an accuracy of 41% at retrieving chlorophyll-a concentration can be reached using the EOF method compared to 140% for the widely-used Ocean Chlorophyll 4 (OC4v4) empirical algorithm, 53% for the Garver-Siegel-Maritorena (GSM01) and 54% for the Generalized Inherent Optical Property (GIOP) semi-analytical algorithms. This result is possible because the EOF approach is able to extract region-specific radiometric features from the satellite remote sensing reflectances that are related to absorption properties of optical components (water, coloured dissolved organic matter and chlorophyll-a) using the visible SeaWiFS channels. The method could easily be used with other ocean-colour satellite sensors (e.g., MODIS, MERIS, VIIRS, OLCI) to extend the time series for the St. Lawrence Estuary and Gulf waters. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Frequency distribution of the 2927 original Chl measurements, 93 of which were considered valid match-ups. (<b>a</b>) Chl concentration frequency distributions (inset presents the same data binned by order of magnitude). Temporal distributions of Chl concentration for (<b>b</b>) the original and (<b>c</b>) match-up datasets.</p>
Full article ">Figure 2
<p>Representation of the different water-types. (<b>a</b>) Classification of the SLEG climatological Rrs using the methodology of [<a href="#B37-remotesensing-10-00265" class="html-bibr">37</a>]. The black line represents the perfect Case-1 relation between Rrs(412)/Rrs(443) and Rrs(555)/Rrs(490). (<b>b</b>) Map of the normalized difference between the climatological SeaWiFS Rrs spectra and the perfect Case-1 line (see the text), with black crosses representing the location of the original Chl samples and grey dots representing the retained Chl samples. The same colour scale applies to both panels.</p>
Full article ">Figure 3
<p>Band ratio algorithms. (<b>a</b>) Comparison of the St. Lawrence Estuary and Gulf (SLEG) match-ups (black dots) and SeaBASS [<a href="#B55-remotesensing-10-00265" class="html-bibr">55</a>] datasets (light grey dots). The grey line corresponds to the OC4v4 polynomial fit, and the black line is the linear regression of in situ Chl and corresponding SLEG remote sensing reflectances ratios, with R443/555 = Rrs(443)/Rrs(555), R490/555 = Rrs(490)/Rrs(555) and R510/555 = Rrs(510)/Rrs(555). (<b>b</b>) Scatterplot of in situ versus satellite-derived Chl using the OC4L algorithm.</p>
Full article ">Figure 4
<p>Features of the EOF method: (<b>a</b>) stability of the EOF; (<b>b</b>) scatter plot of in situ Chl versus satellite-derived Chl using the EOF, with the dashed line as the 1:1 ratio; (<b>c</b>) all modes of oscillation (solid coloured lines are linear interpolation over wavelengths of the discrete spectral data and are used as a guide to aid visualizing the spectral signature).</p>
Full article ">Figure 5
<p>Relative error as a function of (<b>a</b>) Chl concentration and (<b>b</b>) spatial distribution.</p>
Full article ">Figure 6
<p>SLEG chlorophyll for 12 October 2003, as predicted by the EOF algorithm.</p>
Full article ">
18 pages, 2563 KiB  
Article
Hue-Angle Product for Low to Medium Spatial Resolution Optical Satellite Sensors
by Hendrik Jan Van der Woerd and Marcel Robert Wernand
Remote Sens. 2018, 10(2), 180; https://doi.org/10.3390/rs10020180 - 26 Jan 2018
Cited by 78 | Viewed by 8369
Abstract
In the European Citclops project, with a prime aim of developing new tools to involve citizens in the water quality monitoring of natural waters, colour was identified as a simple property that can be measured via a smartphone app and by dedicated low-cost [...] Read more.
In the European Citclops project, with a prime aim of developing new tools to involve citizens in the water quality monitoring of natural waters, colour was identified as a simple property that can be measured via a smartphone app and by dedicated low-cost instruments. In a recent paper, we demonstrated that colour, as expressed mainly by the hue angle (α), can also be derived accurately and consistently from the ocean colour satellite instruments that have observed the Earth since 1997. These instruments provide superior temporal coverage of natural waters, albeit at a reduced spatial resolution of 300 m at best. In this paper, the list of algorithms is extended to the very first ocean colour instrument, and the Moderate Resolution Imaging Spectroradiometer (MODIS) 500-m resolution product. In addition, we explore the potential of the hue angle derivation from multispectral imaging instruments with a higher spatial resolution but reduced spectral resolution: the European Space Agency (ESA) multispectral imager (MSI) on Sentinel-2 A,B, the Operational Land Imager (OLI) on the National Aeronautics and Space Administration (NASA) Landsat-8, and its precursor, the Enhanced Thematic Mapper Plus (ETM+) on Landsat-7. These medium-resolution imagers might play a role in an upscaling from point measurements to the typical 1-km pixel size from ocean colour instruments. As the parameter α (the colour hue angle) is fairly new to the community of water remote sensing scientists, we present examples of how colour can help in the image analysis in terms of water-quality products. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Plot of the available spectral bands in the visual range (400–710 nm) as function of spatial resolution for seven types of sensors. The multispectral imager (MSI) provides three different resolution modes, Landsat 8 has one band more than Landsat 7 at 30-m resolution. The Medium Resolution Imaging Spectrometer (MERIS), Moderate Resolution Imaging Spectroradiometer (MODIS) and Ocean and Land Colour Instrument (OLCI) provide products at two resolutions.</p>
Full article ">Figure 2
<p>Original hyperspectral data plus red squares at the position of MERIS bands (<b>a</b>); the reconstructed spectrum from Rrs at nine MERIS spectral bands (<b>b</b>); a comparison of the two (<b>c</b>); and Colour Matching Function (CMF) curves for red, green, and blue that weight the spectral information to derive the hue angle (<b>d</b>).</p>
Full article ">Figure 3
<p>Original hyperspectral data plus green squares at the central positions of MSI bands (<b>a</b>); MSI measures the visual spectrum in only five wide bands (<b>b</b>); and the reconstructed spectrum will start to deviate from the original spectrum, especially at green and red wavelengths (<b>c</b>,<b>d</b>).</p>
Full article ">Figure 4
<p>Deviation in degrees from the hyperspectral hue angle for six instruments. The horizontal axis gives the hue angle based on the reconstructed spectra (scaled by 100 for practical reasons). Note that, although the shape follows the same pattern for all of the instruments, the vertical scale is very different. The drawn line is the fifth order polynomial fit to the data.</p>
Full article ">Figure 5
<p>Validation of the hue angle reconstruction for six instruments.</p>
Full article ">Figure 6
<p>Plot of the standard deviation in the parameter (<span class="html-italic">α</span>-sensor minus <span class="html-italic">α</span>-hyperspectral) for nine sensors. On the horizontal axis, the hyperspectral hue interval is given in degrees, together with the number of spectra (N) that were used to calculate the standard deviation. The upper panel (<b>a</b>) is based on the International Ocean Colour Coordinating Group (IOCCG) data, and the lower panel (<b>b</b>) is based on the field data.</p>
Full article ">Figure 7
<p>Relation between the hue angle and chlorophyll-<span class="html-italic">a</span> concentration (CHL) for eight classes of Coloured Dissolved Organic Matter (CDOM) concentration. The first four classes (0.001 &lt; g440 &lt; 0.1) can be found in the open ocean, while the next two (0.1 &lt; g440 &lt; 0.5) represent regional seas and coastal waters. The very CDOM-rich waters (0.5 &lt; g440 &lt; 2.5) can be found in rivers and lakes.</p>
Full article ">Figure 8
<p>Relation between the hue angle and the absorption (a) of radiation at 440 nm.</p>
Full article ">
22 pages, 3246 KiB  
Article
High-Chlorophyll-Area Assessment Based on Remote Sensing Observations: The Case Study of Cape Trafalgar
by Iria Sala, Gabriel Navarro, Marina Bolado-Penagos, Fidel Echevarría and Carlos M. García
Remote Sens. 2018, 10(2), 165; https://doi.org/10.3390/rs10020165 - 25 Jan 2018
Cited by 18 | Viewed by 4937
Abstract
Cape Trafalgar has been highlighted as a hotspot of high chlorophyll concentrations, as well as a source of biomass for the Alborán Sea. It is located in an unique geographical framework between the Gulf of Cádiz (GoC), which is dominated by long-term seasonal [...] Read more.
Cape Trafalgar has been highlighted as a hotspot of high chlorophyll concentrations, as well as a source of biomass for the Alborán Sea. It is located in an unique geographical framework between the Gulf of Cádiz (GoC), which is dominated by long-term seasonal variability, and the Strait of Gibraltar, which is mainly governed by short-term tidal variability. Furthermore, here bathymetry plays an important role in the upwelling of nutrient-rich waters. In order to study the spatial and temporal variability of chlorophyll-a in this region, 10 years of ocean colour observations using the MEdium Resolution Imaging Spectrometer (MERIS) were analysed through different approaches. An empirical orthogonal function decomposition distinguished two coastal zones with opposing phases that were analysed by wavelet methods in order to identify their temporal variability. In addition, to better understand the physical–biological interaction in these zones, the co-variation between chlorophyll-a and different environmental variables (wind, river discharge, and tidal current) was analysed. Zone 1, located on the GoC continental shelf, was characterised by a seasonal variability weakened by the influence of other environmental variables. Meanwhile, Zone 2, which represented the dynamics in Cape Trafalgar but did not show any clear pattern of variability, was strongly correlated with tidal current whose variability was probably determined by other drivers. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>(<b>a</b>) Location map of the Gulf of Cádiz (GoC), Strait of Gibraltar (SoG), and the Alborán Sea. The dashed line frames the region of interest of this study, located in the south-west of Spain. (<b>b</b>) Bathymetry of Cape Trafalgar region. Black solid lines represent the bathymetry (15, 20, 25, 50, 75, 100, 120 m). The black cross represents the mooring location (36.13<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>N–6.03<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>W). The black star represents the location of Camarinal Sill (35.91<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>N–5.74<math display="inline"> <semantics> <msup> <mrow/> <mo>∘</mo> </msup> </semantics> </math>W).</p>
Full article ">Figure 2
<p>Monthly climatology of MERIS-OC4Me chlorophyll-<span class="html-italic">a</span> concentration (mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) centred on the Cape Trafalgar region. Black lines represent the bathymetry (15, 20, 25, 50, 75, 100, 120 m).</p>
Full article ">Figure 3
<p>Empirical orthogonal function analysis of fortnightly tidal cycle dataset of MERIS-OC4Me chlorophyll-<span class="html-italic">a</span> concentration (mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>). (<b>a</b>) Fortnightly tidal cycle climatology. Spatial coefficients maps for the (<b>b</b>) first, (<b>c</b>) second, (<b>d</b>) third, and (<b>e</b>) fourth mode. Black lines represent the bathymetry (15, 20, 25, 50, 75, 100 m). Pixels belonging to Cádiz Bay were erased in order to avoid their influence in this analysis.</p>
Full article ">Figure 4
<p>Empirical orthogonal function analysis of fortnightly tidal cycle dataset of MERIS-OC4Me chlorophyll-<span class="html-italic">a</span> concentration (mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>). Temporal coefficient evolution for the (<b>a</b>) first, (<b>b</b>) second, (<b>c</b>) third, and (<b>d</b>) fourth mode. Dotted lines represent the photoperiod, calculated as the fraction of hours of light versus the hours of darkness per day (Equation (<a href="#FD2-remotesensing-10-00165" class="html-disp-formula">2</a>)). Grey background represents winter and spring seasons versus white background, representing summer and autumn seasons.</p>
Full article ">Figure 5
<p>Location of the study zones distinguished by the empirical orthogonal function analyses: Zone 1 (coastal zone in front of Cádiz Bay), Zone 2 (Cape Trafalgar region), and the Open Ocean. Pixels belonging to Cádiz Bay were erased in order to avoid its influence in this analysis.</p>
Full article ">Figure 6
<p>Analysis of the chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) time series computed for the Open Ocean. (<b>a</b>) Temporal series of tidal cycle dataset chl-<span class="html-italic">a</span> concentration. (<b>b</b>) Distribution of chl-<span class="html-italic">a</span> concentration. Dotted lines (<b>a</b>,<b>b</b>) show average chl-<span class="html-italic">a</span> concentration. (<b>c</b>) Wavelet power spectrum. The colour code varies from dark blue (low values), to dark red (high values). The black line indicates the cone of influence. (<b>d</b>) Average wavelet power spectrum. Dotted lines (<b>c</b>,<b>d</b>) show the <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> = 5% significance level computed based on 1000 Markov bootstrapped series.</p>
Full article ">Figure 7
<p>Analysis of the chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) time series computed for Zone 1. (<b>a</b>) Temporal series of tidal cycle dataset chl-<span class="html-italic">a</span> concentration. (<b>b</b>) Distribution of chl-<span class="html-italic">a</span> concentration. Dotted lines (<b>a</b>,<b>b</b>) show average chl-<span class="html-italic">a</span> concentration. (<b>c</b>) Wavelet power spectrum. Colour code varies from dark blue (low values), to dark red (high values). The black line indicates the cone of influence. (<b>d</b>) Average wavelet power spectrum. Dotted lines (<b>c</b>,<b>d</b>) show the <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> = 5% significance level computed based on 1000 Markov bootstrapped series.</p>
Full article ">Figure 8
<p>Analysis of the chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) time series computed for Zone 2. (<b>a</b>) Temporal series of tidal cycle dataset chl-<span class="html-italic">a</span> concentration. (<b>b</b>) Distribution of chl-<span class="html-italic">a</span> concentration. Dotted lines (<b>a</b>,<b>b</b>) show average chl-<span class="html-italic">a</span> concentration. (<b>c</b>) Wavelet power spectrum. The colour code varies from dark blue (low values), to dark red (high values). The black line indicates the cone of influence. (<b>d</b>) Average wavelet power spectrum. Dotted lines (<b>c</b>,<b>d</b>) show the <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> = 5% significance level computed based on 1000 Markov bootstrapped series.</p>
Full article ">Figure 9
<p>Left-hand panels: Average wavelet cross-spectrum between chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) and the zonal component of wind (u-wind) in Zone 1 (<b>a</b>) and Zone 2 (<b>c</b>). Dotted lines show the <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> = 5% significance levels computed based on 1000 Markov bootstrapped series. Right-hand panels: Phases of the two time series, chl-<span class="html-italic">a</span> concentration (red line) and u-wind (blue line), computed for the peak of highest coherence in Zone 1 (<b>b</b>) and Zone 2 (<b>d</b>). Dashed lines represent the phase difference between both signals.</p>
Full article ">Figure 10
<p>Left-hand panel: Average wavelet cross-spectrum between chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) and Guadalquivir river discharge in Zone 1 (<b>a</b>) and Zone 2 (<b>c</b>). Dotted lines show the <math display="inline"> <semantics> <mi>α</mi> </semantics> </math> = 5% significance levels computed based on 1000 Markov bootstrapped series. Right-hand panel: Phases of the two time series, chl-<span class="html-italic">a</span> concentration (red line), and river discharge (blue line), computed for the peak of highest coherence in Zone 1 (<b>b</b>) and Zone 2 (<b>d</b>). Dashed lines represent the phase difference between both signals.</p>
Full article ">Figure 11
<p>Left-hand panel: Average wavelet cross-spectrum between chlorophyll-<span class="html-italic">a</span> concentration (chl-<span class="html-italic">a</span>; mg m<math display="inline"> <semantics> <msup> <mrow/> <mrow> <mo>−</mo> <mn>3</mn> </mrow> </msup> </semantics> </math>) and zonal component of tidal current (u-tide) for the three selected periods in Zone 1 (<b>a</b>) and Zone 2 (<b>c</b>). Right-hand panel: Phases of the two time series, chl-<span class="html-italic">a</span> concentration (red line), and u-tide (blue line), for the three selected periods, computed for the peak of highest coherence in Zone 1 (<b>b</b>) and Zone 2 (<b>d</b>). Dashed lines represent the phase difference between both signals.</p>
Full article ">Figure A1
<p>Spatial coefficient maps corresponding to the first four modes of the empirical orthogonal function (EOF) analysis performed with the spring tides dataset.</p>
Full article ">Figure A2
<p>Spatial coefficient maps corresponding to the first four modes of the EOF analysis performed with the neap tides dataset.</p>
Full article ">

Other

Jump to: Research

13 pages, 23574 KiB  
Technical Note
Annual New Production of Phytoplankton Estimated from MODIS-Derived Nitrate Concentration in the East/Japan Sea
by Huitae Joo, Dabin Lee, Seung Hyun Son and Sang Heon Lee
Remote Sens. 2018, 10(5), 806; https://doi.org/10.3390/rs10050806 - 22 May 2018
Cited by 6 | Viewed by 5281
Abstract
Our main objective in this study was to determine the inter-annual variation of the annual new production in the East/Japan Sea (EJS), which was estimated from MODIS-aqua satellite-derived sea surface nitrate (SSN). The new production was extracted from northern (>40° N) and southern [...] Read more.
Our main objective in this study was to determine the inter-annual variation of the annual new production in the East/Japan Sea (EJS), which was estimated from MODIS-aqua satellite-derived sea surface nitrate (SSN). The new production was extracted from northern (>40° N) and southern (>40° N) part of EJS based on Sub Polar Front (SPF). Based on the SSN concentrations derived from satellite data, we found that the annual new production (Mean ± S.D = 85.6 ± 10.1 g C m−2 year−1) in the northern part of the EJS was significantly higher (t-test, p < 0.01) than that of the southern part of the EJS (Mean ± S.D = 65.6 ± 3.9 g C m−2 year−1). Given the relationships between the new productions and sea surface temperature (SST) in this study, the new production could be more susceptible in the northern part than the southern part of the EJS under consistent SST warming. Since the new production estimated in this study is only based on the nitrate inputs into the euphotic depths during the winter, new productions from additional nitrate sources (e.g., the nitrate upward flux through the MLD and atmospheric deposition) should be considered for estimating the annual new production. Full article
(This article belongs to the Special Issue Remote Sensing of Ocean Colour)
Show Figures

Figure 1

Figure 1
<p>Field measurement stations in the East/Japan Sea (EJS). Squared dots indicate the NIFS (National Institute of Fisheries Sciences, Korea) stations. Cross marks are the JMA (Japan Meteorological Agency, Japan) stations. Dots are the MOF (Ministry of Oceans and Fisheries, Korea) stations. Red dots are the locations of C/N ratio measured previously.</p>
Full article ">Figure 2
<p>Relationships between Sea Surface Nitrate (SSN) and Sea Surface Temperature (SST) in the northern (<b>a</b>) and southern (<b>b</b>) East/Japan Sea.</p>
Full article ">Figure 3
<p>Correlations between field-measured Sea Surface Nitrate (SSN) and estimated SSN in the northern (<b>a</b>) and southern (<b>b</b>) East/Japan Sea. Dashed lines represent 1:1 lines. Red lines represent the 95% prediction bounds. Blue lines show the regression curves between in situ SSN and estimated SSN from the algorithm developed by Goes et al. (2000).</p>
Full article ">Figure 4
<p>Climatological monthly images of Sea Surface Nitrate (SSN) in the East/Japan Sea (EJS).</p>
Full article ">Figure 5
<p>Time series of monthly values in Sea Surface Nitrate (SSN) for the northern and southern East/Japan Sea (EJS) from July 2002 to December 2015.</p>
Full article ">Figure 6
<p>Long-term pattern of the annual Sea Surface Nitrate (SSN) in the East/Japan Sea (EJS) from 2003 to 2015.</p>
Full article ">Figure 7
<p>Long-term pattern of the annual new production in the East/Japan Sea (EJS) from 2003 to 2015.</p>
Full article ">Figure 8
<p>Anomalies of the annual new productions in the northern and southern East/Japan Sea (EJS) during the study period. The blue color bars and red color bars represent the northern and southern East/Japan Sea (EJS), respectively.</p>
Full article ">Figure 9
<p>Relationships between the annual new production and Sea Surface Temperature (SST) in March in the northern (<b>a</b>) and southern (<b>b</b>) East/Japan Sea (EJS).</p>
Full article ">
Back to TopTop