Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Biophysical modelling of phytoplankton communities from first principles using two-layered spheres: Equivalent Algal Populations (EAP) model

Open Access Open Access

Abstract

There is a pressing need for improved bio-optical models of high biomass waters as eutrophication of coastal and inland waters becomes an increasing problem. Seasonal boom conditions in the Southern Benguela and persistent harmful algal production in various inland waters in Southern Africa present valuable opportunities for the development of such modelling capabilities. The phytoplankton-dominated signal of these waters additionally addresses an increased interest in Phytoplankton Functional Type (PFT) analysis. To these ends, an initial validation of a new model of Equivalent Algal Populations (EAP) is presented here. This paper makes a first order comparison of two prominent phytoplankton Inherent Optical Property (IOP) models with the EAP model, which places emphasis on explicit bio-physical modelling of the phytoplankton population as a holistic determinant of inherent optical properties. This emphasis is shown to have an impact on the ability to retrieve the detailed phytoplankton spectral scattering information necessary for PFT applications and to successfully simulate reflectance across wide ranges of physical environments, biomass, and assemblage characteristics.

© 2014 Optical Society of America

1. Introduction

Eutrophication is a leading cause of impairment of aquatic ecosystems worldwide, and is accelerating in rate and extent due in large part to human activities [1]. Toxic algae, tainted freshwater supplies and fisheries-threatening hypoxia can all result in health risks and substantial economic losses. The need for improved water quality monitoring in vulnerable coastal regions and of inland freshwater resources is best addressed by a combined approach: increased monitoring of biogeophysical variables in these environments (both in situ and by remote sensing), coupled with the development of reliable bio-optical models in pursuit of better understanding of the biophysical relationships at play.

Existing optical modelling approaches [25] employ simple IOP models generally most suited to open-ocean (oligo- or meso-trophic) conditions and usually heavily dependent on Chlorophyll a (Chl a) specificity. Such models also typically decouple the phytoplankton absorption and backscattering terms, or use backscattering terms related only to the gross particulate: an approach that has served well in relatively low biomass oceanic waters, but is fundamentally restrictive for the analysis of the phytoplankton -specific ocean colour signal. This work has been undertaken, in part, to assess the suitability of a more holistic phytoplankton optical modelling approach for eutrophic environments, and ultimately to provide an enhanced capacity to analyse the ocean colour signal content related to the variability of phytoplankton functional types.

This EAP model represents a departure from existing modelling approaches in that it is population driven in terms of phytoplankton functional type (PFT). In other words, phytoplankton IOPs are formulated from population-specific refractive indices and so are not independent of each other. It has been established that at elevated biomass the spectral character of phytoplankton scattering becomes increasingly important [23,24]. Implementation of the EAP model allows investigation into when the resulting AOPs (including remote sensing reflectance, Rrs) may be sensitive to this variability, through the use of spectrally variant phase functions in the radiative transfer component of the model. Also, size- and assemblage-related influences are not well parameterised in the literature [7], and the EAP model additionally allows for investigation into these effects.

The very productive southern Benguela lends itself well to the development of optical modelling and validation activities. Inner shelf waters are typically dominated by 3 phytoplankton groups: potentially bloom-forming diatom and dinoflagellate assemblages, and nanophytes (small-celled assemblages, including some chlorophytes). These groups have distinctive optical characteristics and in any given natural assemblage there may be optically important components of all of them. For simplicity, modelling of only the dinoflagellate group is discussed here, as representative of frequently occurring high biomass blooms.

2. Biophysical modelling from first principles using two-layered spheres

2.1. Phytoplankton component

Assemblages are modelled using equivalent size distributions and a two-layered sphere cell geometry [6, 7], comprising a core sphere (representing the cytoplasm) and a shell sphere (chloroplast). Phytoplankton IOPs are generated from the real and imaginary parts of the cell refractive indices [7]. Where detailed refractive index data are not known at longer wavelengths (> 750 nm) and imaginary refractive indices are assumed to be very small, they are linearly interpolated from their values at 720 nm to 1e-5 at 750 nm, and then to 1e-9 at 900 nm, with subsequent quantified effect on the real part of the refractive index using established methods [7]. (This allows consistency between the RIs of different PFTs at long wavelengths).

The relative chloroplast volume is maintained at 20% while the effective diameter of the cell varies. The effective diameter of the modelled distribution is thus central to the IOP model, with the resulting IOPs depending heavily on particle size. The modelled IOPs have a constant Chl a density per cell (ci), set at 2.5 kg.m−3 which was chosen as representative from the literature [7]. Recent experiments however indicate that covarying ci with effective diameter may be more appropriate as the Chl a density of specific species becomes more important with the bloom’s monospecificity. However, this approach is not pursued further here.

For the forward modelling of example populations, a Standard normal particle size distribution (from 1 to 100 μm at 1 μm resolution) with effective variance of 0.6 is used as a reasonable approximation to the reduced species diversity typical of blooms of elevated biomass [6].

2.2. Other components

The primary intention here is to examine the modelled phytoplankton biomass and assemblage effects on ocean colour, so very simplistic scaling of other constituents is employed to demonstrate relative model performance.

A simple exponential combined gelbstoff and detrital absorption term agd(λ) [8, 9] is used as representative of commonly occurring conditions in the Benguela:

agd(λ)=agd(400)exp[S(λ400)]
The exponential slope factor S is given a constant value of 0.012 [10].

An observed relationship of

agd(400)=0.0904log[Chla]+0.1287
from measurements in the Benguela is used to scale the gelbstof/detrital exponential term, and agd(750) onwards is assumed to be zero. Non-algal backscattering is modelled after Roesler and Perry [11] who describe a small particle backscattering term (bbs) represented by a power law relationship (bbs = λ−1.2) with a constant spectral shape dependent only on wavelength but variable in magnitude. Small particle (non-algal) scatter bs is approximated as 50 times the bbs. This yields a non-algal particulate backscattering probability (bs) of 0.02 (2%). This is assumed to be reasonable given that it has been shown that the total particulate backscattering probability b varies in the range 1.2 to 3.2 % in coastal waters dominated by non-algal particles (i.e. Case 2) [12], and that generally accepted values for b in Case 1 waters is around 1% [13]. Keeping the non-phytoplankton backscattering constant with Chl a results in a dependent but non-linear relationship, resulting in an overall b that decreases as Chl a increases, while maintaining its significant optical contribution and increasing spectral variability as eutrophication occurs.

Ecolight radiative transfer software (Sequoia Scientific, 2008) was used to generate Rrs from these IOPs (given total absorption, attenuation and backscattering data), using wavelength-specific Fournier Forand phase functions as b varies across the wavelength spectrum. Fluorescence quantum efficiency ϕ was approximated as follows by Chl a concentration: below 10 mg.m−3 = 1%, 10–50 mg.m−3 = 0.6%, 50–100 mg.m−3 = 0.2% and over 100 mg.m−3 = 0.1%.These values are based on MODIS ϕsat climatologies [14], and measurements [15] to characterise the reduction in ϕ as eutrophication increases. An annual average for solar irradiance and a solar zenith of 30 degrees was used in lieu of time and location.

3. A comparison of IOP models

EAP simulated Rrs were compared to a subset of those from models described by Alvain et al. [2] and Lee et al. [3], two prominent phytoplankton IOP models both well validated for their applications, designed for Chl a concentrations ranging from 0.02 – 3 mg.m−3 [2] and 0.03 – 30 mg.m−3 [3]. The IOP approaches of both models are abbreviated in Table 1. Abbreviations used are detailed in Table 2.

Tables Icon

Table 1. IOP parameterisations of the models of Alvain [2] and Lee [3]

The models of both Alvain and Lee use Chl a specific absorption spectra aϕ*(λ) as described by Bricaud [16]. The approaches taken to gelbstof absorption are similar to the EAP model. The Alvain model neglects a separate detrital absorption term (representing all non-algal particles) on the basis that the spectral shape is similar to that of agd(λ) and that absorption by non-algal particles represents just 10% of gelbstoff absorption (as determined by Siegel et al., 2002, in [2]). Neither model includes a fluorescence term.

A significant difference between these two approaches is the determination of the backscatter. The Alvain model uses a spectrally invariant Petzold phase function for the total particulate backscatter b. This results in a constant total backscatter fraction of 0.0189. The Lee model separates the backscatter into its two components, calculating b from bϕ with a 1% Fournier Forand phase function, and bbdet from bbdet with a Petzold phase function.

Both models have included dynamic ranges of IOP values in order to account for the variability observed in natural waters. For clarity when making the comparison to the EAP model, IOP values were selected from within these dynamic ranges which most closely approximate those of the EAP model. For the ag(λ) and agd(λ) contributions this selection was based on the EAP value of agd(400). Likewise, a constant detrital scattering function approximating that of the EAP was selected from the ranges offered by the Lee and Alvain models. This ensures that the comparison between models is made under consistent (i.e. Benguela-like) water type conditions. Furthermore, the Alvain and Lee models were constrained for the purposes of this comparison to use the same Chl a specific phytoplankton absorption spectra [16]. A subset of phytopankton scattering functions was selected by comparing the coefficients at 550 nm to those of the EAP model.

While these constraints inevitably reduce the range of resulting Rrs values, these subsets should be acceptable for the purpose of commenting primarily on the different approaches to modelling the backscattering and the significance of these approaches as phytoplankton biomass increases to eutrophic concentrations.

The Alvain IOPs were processed to Rrs using Ecolight’s two component model for Chl a concentrations of 1, 2, 5, 10, 15 and 30 mg.m−3, using total a and b and a Petzold phase function for all particulate matter. This results in a total particulate backscatter fraction of up to 1.9% which is considered too high for Chl a concentrations over about 1 mg.m−3 [17].

The Lee IOPs were also processed to Rrs using Ecolight’s two component model for Chl a concentrations of 1, 2, 5, 10, 15 and 30 mg.m−3 using the total absorption and scattering coefficients. A combined backscattering phase function was calculated for each wavelength using the 1% Fournier Forand phase function to retrieve the b from the bϕ, and the Petzold to retrieve the bbdet from the bdet. The sum of these, i.e. the total bb, was then used as input for Ecolight’s wavelength-specific variable b Fournier Forand phase function selection option.

4. Results and Discussion

4.1. Comparison of the constrained Alvain, Lee and EAP models

The Rrs generated (Fig. 1) are beyond the expected performance of the Alvain model (i.e. Chl a concentrations up to 3 mg.m−3) and up to the very limit of the Lee model (up to 30 mg.m−3) in order to examine the impacts of the different approaches to the backscattering terms as biomass increases. Rrs from the 3 models agree well for the Chla = 1 and 2 mg.m−3 experiments. Divergence is observed after Chl a = 5 mg.m−3 (Fig. 1).

Most notable perhaps at first glance are the comparatively large differences in magnitude of Rrs between EAP Chl a = 1 and 2 mg.m−3 with respect to the other models’. It should be emphasised that the EAP model, in this its simplest implementation, describes a homogenous dinoflagellate population with an idealised size distribution. At these low phytoplankton concentrations, natural populations would likely be variable both in species and in particle size.

 figure: Fig. 1

Fig. 1 Comparison of Ecolight modelled Rrs from 3 IOP models: EAP, Alvain [2] and Lee [3]. Chl a values are 1, 2, 5, 10, 15 and 30 mg.m−3 in each case.

Download Full Size | PDF

4.2. Preliminary validation of Rrs

A preliminary validation exercise was carried out using measured Tethered Surface Radiometer Buoy (TSRB) data from 2002–2008 in the Benguela, processed to Rrs using ProSoft proprietary software (Satlantic). Measurements were selected for each chosen Chl a class, and the mean spectrum is presented using one standard deviation as an indication of variability in the optical measurements of natural populations (Fig. 2).

In the lower biomass classes (up to 5 mg.m−3) the models all perform reasonably well considering the simplifications and constraints placed on them. As biomass increases past Chl a of 10 mg.m−3 however, the effect on the Rrs of the differences in the models’ approaches to back-scattering becomes evident. The notably reduced brightness of Rrs in the measurements is most accurately modelled using the EAP model’s spectrally variant total backscattering probability which varies in magnitude with both Chl a concentration and wavelength.

 figure: Fig. 2

Fig. 2 Preliminary validation of the 3 models with measured Rrs, for Chl a classes of 1, 2, 5, 10, 15 and 30 mg.m−3. The mean measured spectrum in each class is presented with 1 standard deviation as an indication of natural variability. The mean measured Chl a for each set of measurements is presented at the top of each example with the standard deviation and the number of measurements, N.

Download Full Size | PDF

It should be noted here that the use of the Bricaud aϕ* by both Lee and Alvain implies a heavy dependence on size with increasing Chl a concentrations that make them unsuitable for use over very wide ranges of biomass. The Bricaud measurements were mostly performed in ocean basin areas with Chl a concentrations of around 0.05 to 8 mg.m−3 [18], with one sampling station going up to 24 mg.m−3 in the St. Lawrence estuary. These aϕ* are appropriate for oligo/mesotrophic waters where a relationship of increasing cell size effects with increasing Chl a is generally observed. In highly eutrophic waters this relationship does not necessarily hold [19]. Figure 3 shows how the EAP model is forced to choose implausibly large sizes at high Chl a concentrations to maintain consistency with the Bricaud absorption spectra. The absorption profiles modelled above fit EAP size classes with effective diameters of between 6 and 16 μm, which were chosen according to observed dominant phytoplankton species’ effective diameters rather than Bricaud-equivalent spectra (which would force the EAP model to go up to a Deff of 45 μm at Chl a of 30 mg.m−3, representing an unusually large cell).

 figure: Fig. 3

Fig. 3 Bricaud aϕ* for Chl a = 1, 2, 5, 10, 15 and 30 mg.m−3 matched to EAP aϕ* by effective diameter (Deff), implying EAP assemblage Deff of 6, 9, 16, 23, 26 and 45 μm respectively.

Download Full Size | PDF

4.3. Employing equivalent total absorption coefficients

To account for the differences in absorption, avoid the problem of an inherent dependency on large sizes at very high biomass, and in order to be able to comment more fully on the backscattering which is central to the EAP model and performing accurate IOP modelling in eutrophic waters, the comparative models were re-created, all using the EAP phytoplankton absorption spectra for each Chl a class (Fig. 4). Also, the EAP fluorescence term was neglected to match the other models. The Lee model’s detrital absorption term adet (λ) was also neglected as it is not explicitly included in the other models, and as Lee ag(λ) values were selected for this comparison to most closely match the EAP combined agd(λ), that term can be considered to include detrital absorption (Fig. 4).

Immediately noticeable in the resulting Rrs is the well constrained 600 to 650 nm region in the EAP model with respect to the Alvain model.

 figure: Fig. 4

Fig. 4 Comparison of Ecolight modelled Rrs from the 3 IOP models: EAP, Alvain [2] and Lee [3]. Chl a values are 1, 2, 5, 10, 15 and 30 mg.m−3 in each case. This comparison uses identical (EAP) aϕ*(λ)s in order to examine the effect of the different approaches to phytoplankton backscattering.

Download Full Size | PDF

The Alvain model’s total scattering compares well with EAP b(λ) in magnitude (Fig. 5, top left) but the choice of one spectrally invariant phase function for both phytoplankton and non-algal particles (Petzold) results in an unrealistically high backscatter fraction which is constant over varying biomass (Fig. 5, bottom left). In turn this results in greatly magnified backscattering at high Chl a concentrations.

 figure: Fig. 5

Fig. 5 Scattering and backscattering characteristics of EAP, Alvain [2] and Lee [3] at Chl a concentrations 1, 2, 5, 10, 15 and 30 mg.m−3. Phytoplankton and detrital backscatter (bottom right) are shown for Chl a of 1 (lower line) and 30 mg.m−3 (upper line). The increased spectral detail of the EAP model’s phytoplankton scatter and backscatter becomes increasingly important with increasing biomass as is evident in the spectral variation in the resulting total backscattering probability (bottom left).

Download Full Size | PDF

The spectrally less variant total particulate scatter of the Lee model, and choice of spectrally flat phase functions (Fournier Forand 1% and Petzold) for both the phytoplankton and non-algal components results in total backscattering profiles with more or less indistinct spectral features, although the average magnitudes compare well to EAP in the mid biomass classes. The increased spectral variability of EAP backscattering is translated into the Rrs as a shift of the 680 nm peak towards 709 as biomass increases - a phenomena also caused by the increase in the red Chl a absorption band and the rapidly increasing water absorption with increasing wavelength. This is an important observed feature at the ”red edge” of eutrophic water reflectances [20].

Modelling backscattering as flattened spectra with reduced spectral variability may enable adequate Rrs modelling at low biomass but as Chl a concentration and the importance of non-algal scattering increases, such simple bb models are unable to reproduce the spectral shape and magnitude variations in the Rrs successfully. The EAP model reflects an approach to backscatter derived from phytoplankton community optics rather than a calculation based on attenuation and/or the assumption of spectrally invariant particle scattering characteristics.

4.4. Characterising eutrophic water Rrs

Having shown the validity of the EAP model Rrs with respect to measurements made in the Benguela, the model can now be used to examine the characteristics of much more eutrophic water conditions. Figure 6 shows Ecolight modelled Rrs using IOPs generated using the EAP model, for biomass increasing from 1 to 200 mg.m−3. A Standard normal particle size distribution from 1 to 100 μm with effective diameter 16 μm is used here, representing a homogenous assemblage displaying the typical optical characteristics of a dinoflagellate population. This simulation (again using Ecolight) includes contributions from the combined gelbstoff/detrital absorption term agd(λ) and small particle backscattering bbs(λ) as described earlier, and uses a choice of Fournier Forand phase function with a variable b at each output wavelength.

 figure: Fig. 6

Fig. 6 Modelled EAP Rrs (above) for typical dinoflagellate assemblage with effective diameter of 16 μm. Chl a concentrations are 1, 2, 5, 10, 20, 30, 50, 100, 150, 200 mg.m−3. Below the shift from maximum peak reflectance height in the blue/green to the red is shown (dotted lines), for increasing Chl a. The first derivatives of these slopes (solid lines) cross at a Chl a of around 15 mg.m−3, the point at which the red features of high biomass reflectance spectra start to dominate.

Download Full Size | PDF

As biomass increases, the modelled Rrs show most distinctly a shift in the fluorescence peak at around 680 nm (in low biomass) towards 710 nm at Chl a concentrations of above 100 mg.m−3. This is caused by the combined effects of a decrease in the fluorescence quantum efficiency [20] and increased phytoplankton absorption as this region becomes the dominant identifier of high concentrations of Chl a. Water is also strongly absorbing in this region which magnifies this effect. The switch in dominance of the 443:555 to the 709:555 band ratio (Fig. 6, as per the SeaWiFs OC4 Chl a retrieval algorithm [21]) shows the importance of signal in the red for Chl a estimation in waters with concentrations over about 25 mg.m−3.

Other notable features of Rrs with increasing eutrophication are convergence at around 660 nm, as well as around 560 nm, and the well constrained region between these points. The decrease in Rrs in the blue is indicative of decreased influence of gelbstoff absorption with respect to Chl a absorption as the latter becomes more dominant spectrally.

4.5. EAP Rrs validation at very high biomass

Some very high biomass TSRB measured Rrs are presented here. The measurements (N=4, 6 and 5 respectively) represent a range of assemblage types and size distributions, which the EAP forward modelled Rrs do not consider here. This brief validation exercise shows that a chosen effective diameter of 12 μm most accurately matches all three high biomass blooms (for Chl a of 110, 150 and 180 mg.m−3 respectively). Most of the Rrs measurements were from a Prorocentrum triestinum-dominated bloom in 2005, with varying lesser proportions of Dinophysis acuminata, D. fortii and P. reticulatum [22]. P. triestinum is a small dinoflagellate approximately 18–22 μm in length and 6–11 μm in width. The variability in resulting Rrs from size effects is clearly seen.

As biomass increases, the need for the employment of wavelength-dependent phase functions for at very least the phytoplankton component is evident. Figure 7 (bottom right) shows the increasingly variable values of the backscatter fraction as Chl a increases (for consistent assemblage Deff and phytoplankton type). These range from an essentially spectrally constant backscatter fraction at low Chl a (1 mg.m−3) to variability across the wavelength spectrum from 0.002 (i.e. 0.2%) to 0.024 (i.e. 2.4%) in eutrophic water conditions.

 figure: Fig. 7

Fig. 7 High Biomass Validation of EAP Rrs, with EAP total backscatter probability shown for Chl a 1, 10, 50, 100, 150, 200 and 300 mg.m−3.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Contribution of Phytoplankton to total IOPs and Rrs signal, for Chl a values of 2 and 150 mg.m−3, for an idealised dinoflagellate assemblage with effective diameter of 16 μm.

Download Full Size | PDF

4.6. Proportional contribution of phytoplankton to total Rrs

From EAP modelled data, the proportional (percentage) contributions of Chl a to the total absorption, total scatter and backscatter are shown (Fig. 8) for Chl a = 2 mg.m−3 and 150 mg.m−3, keeping the effective diameter constant at 16 μm (representing a dinoflagellate assemblage). It can be concluded that at lower biomass, gelbstoff and detrital/small particles are optically significant, while at high biomass the signal is almost exclusively due to phytoplankton pigment. That being said, however, the relative contribution of detritus and small particles to backscatter remains important even at Chl a of 150 mg.m−3 and this is translated into the Rrs signal especially between 550 and 680 nm (Fig. 8).

[Note: These proportions include absorption and scatter of the medium itself (water). Phytoplankton Rrsp is calculated with only water + phytoplankton specific IOPs. Other contributions are total RrsRrsp. Any optical interactions between phytoplankton and other components are therefore neglected here.]

5. Conclusion

The EAP model is shown to be appropriate for IOP and Rrs modelling in high biomass Case 1 waters. Here it is presented in its simplest, forward form but there is much scope for added complexity and sensitivity. The modelling of mixed assemblages (mixed phytoplankton types of varying ci and pigment concentrations, and different population size distributions) can be replicated as required. Cells containing vacuoles and other anomalous scattering characteristics can be considered too. The importance of detailed spectral backscattering is emphasised here. A full investigation into the sensitivity of the model to the use of discretised phase functions is necessary to determine the extent of reliance on the detailed spectral backscattering component.

Tables Icon

Table 2. Symbols and Abbreviations

Acknowledgments

Funding awarded to L. Robertson Lain from the Centre for Scientific and Industrial Research (CSIR) and University of Cape Town (UCT) PhD Scholarship Programme is gratefully acknowledged, as is funding from the CSIR/DST SWEOS Strategic Research Programme. UCT BASICS and ACCESS programmes are acknowledged for support of H. Evers-King.

References and links

1. M. F. Chislock, E. Doster, and R. A. Zitomer, “Eutrophication: Causes, Consequences, and Controls in Aquatic Ecosystems,” Nature Education Knowledge 4, 10 (2013).

2. S. Alvain, H. Loisel, and D. Dessailly, “Theoretical analysis of ocean color radiances anomalies and implications for phytoplankton groups detection in case 1 waters,” Opt. Express 20, 1070–1083 (2012). [CrossRef]   [PubMed]  

3. Z.P. Lee, ed. “Remote Sensing of Inherent Optical Properties: Fundamentals, Tests of Algorithms, and Applications,” Reports of the International Ocean Colour Coordinating Group 5, 1–122 (2006).

4. K. L. Carder, S. K. Hawes, K. A. Baker, R. C. Smith, R. G. Steward, and B. G. Mitchell, “Reflectance model for quantifying chlorophyll a in the presence of productivity degradation products,” J. Geophys. Res.: Oceans 96, 20599–20611 (1991). [CrossRef]  

5. J. Fischer and F. Fell, “Simulation of MERIS measurements above selected ocean waters,” Int. J. Remote Sens. 20, 1787–1807 (1999). [CrossRef]  

6. S. Bernard, F. A. Shillington, and T. A. Probyn, “The use of equivalent size distributions of natural phytoplankton assemblages for optical modeling,” Opt. Express 15, 1995–2007 (2007). [CrossRef]   [PubMed]  

7. S. Bernard, T. A. Probyn, and A. Quirantes, “Simulating the optical properties of phytoplankton cells using a two-layered spherical geometry,” Biogeosciences Discussions 6, 1497–1563 (2009). [CrossRef]  

8. A. Bricaud, A. Morel, and L. Prieur, “Absorption by dissolved organic matter of the sea (yellow substance) in the UV and visible domains,” Limnol. Oceanogr. 26, 43–53 (1981). [CrossRef]  

9. C. S. Roesler, M. J. Perry, and K. L. Carder, “Modeling in situ phytoplankton absorption from total absorption spectra in productive inland marine waters,” Limnol. Oceanogr. 34, 1510–1523 (1989). [CrossRef]  

10. S. Bernard, T. A. Probyn, and F. A. Shillington, “Towards the validation of SeaWiFS in southern African waters: the effects of gelbstoff,” S. Afr. J. Marine Sci. 19, 15–25 (1998). [CrossRef]  

11. C. S. Roesler and M. J. Perry, “In situ phytoplankton absorption, fluorescence emission, and particulate backscattering spectra determined from reflectance,” J. Geophys. Res. 100, 13279–13294 (1995). [CrossRef]  

12. M. Chami, E. B. Shybanov, G. A. Khomenko, M. E. G. Lee, O. V. Martynov, and G. K. Korotaev, “Spectral variation of the volume scattering function measured over the full range of scattering angles in a coastal environment,” Appl. Optics 45, 3605–3619 (2006). [CrossRef]  

13. M. S. Twardowski, E. Boss, J. B. Macdonald, W. S. Pegau, A. H. Barnard, and J. R. V. Zaneveld, “A model for estimating bulk refractive index from the optical backscattering ratio and the implications for understanding particle composition in case I and case II waters,” J. Geophys. Res. 106, 14129–14142 (2001). [CrossRef]  

14. M. J. Behrenfeld, T. K. Westberry, and E. Boss, “Satellite-detected fluorescence reveals global physiology of ocean phytoplankton,” Biogeosciences 6, 779–794 (2009). [CrossRef]  

15. M. Ostrowska, B. WoŸniak, and J. Dera, “Modelled quantum yields and energy efficiency of fluorescence, photosynthesis and heat production by phytoplankton in the World Ocean,” Oceanologia 54, 565–610 (2012).

16. A. Bricaud, A.-L. Bédhomme, and A. Morel, “Optical properties of diverse phytoplanktonic species: Experimental results and theoretical interpretation,” J. Plankton Res. 10, 851–873 (1988). [CrossRef]  

17. A. Morel and S. Maritorena, “Bio-optical properties of oceanic waters: A reappraisal,” J. Geophys. Res. 106, 7163–7180 (2001). [CrossRef]  

18. A. Bricaud, M. Babin, A. Morel, and H. Claustre, “Variability in the chlorophyll-specific absorption coefficients of natural phytoplankton: Analysis and parameterization,” J. Geophys. Res. 100, 13321–13332 (1995). [CrossRef]  

19. M. Crichton, L. Hutchings, T. Lamont, and A. Jarre, “From physics to phytoplankton: prediction of dominant cell size in St Helena Bay in the Southern Benguela,” J. Plankton Res. 35, 526–541 (2013). [CrossRef]  

20. H. M. Dierssen, R. M. Kudela, and J. P. Ryan, “Red and black tides: Quantitative analysis of water-leaving radiance and perceived color for phytoplankton, colored dissolved organic matter, and suspended . . .,” Limnol. Oceanogr. 51, 2646–2659 (2006). [CrossRef]  

21. J. E. O’Reilly, S. Maritorena, B. G. Mitchell, D. A. Siegel, K. L. Carder, S. A. Garver, M. Kahru, and C. McClain, “Ocean color chlorophyll algorithms for SeaWiFS,” J. Geophys. Res. 103, 24937–24953 (1998). [CrossRef]  

22. A. Fawcett, G. C. Pitcher, S. Bernard, and A. Cembella, “Contrasting wind patterns and toxigenic phytoplankton in the southern Benguela upwelling system,” Mar. Ecol. Prog. Ser. 348, 19–31 (2007). [CrossRef]  

23. A. Whitmire, E. Boss, T. Cowles, and S. W. Scott Pegau, “Spectral variability of the particulate backscattering ratio,” Opt. Express 15, 7019–7031 (2007). [CrossRef]   [PubMed]  

24. R. D. Vaillancourt, “Light backscattering properties of marine phytoplankton: relationships to cell size, chemical composition and taxonomy,” J. Plankton Res. 26, 191–212 (2004). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (8)

Fig. 1
Fig. 1 Comparison of Ecolight modelled Rrs from 3 IOP models: EAP, Alvain [2] and Lee [3]. Chl a values are 1, 2, 5, 10, 15 and 30 mg.m−3 in each case.
Fig. 2
Fig. 2 Preliminary validation of the 3 models with measured Rrs, for Chl a classes of 1, 2, 5, 10, 15 and 30 mg.m−3. The mean measured spectrum in each class is presented with 1 standard deviation as an indication of natural variability. The mean measured Chl a for each set of measurements is presented at the top of each example with the standard deviation and the number of measurements, N.
Fig. 3
Fig. 3 Bricaud a ϕ * for Chl a = 1, 2, 5, 10, 15 and 30 mg.m−3 matched to EAP a ϕ * by effective diameter (Deff), implying EAP assemblage Deff of 6, 9, 16, 23, 26 and 45 μm respectively.
Fig. 4
Fig. 4 Comparison of Ecolight modelled Rrs from the 3 IOP models: EAP, Alvain [2] and Lee [3]. Chl a values are 1, 2, 5, 10, 15 and 30 mg.m−3 in each case. This comparison uses identical (EAP) a ϕ * ( λ )s in order to examine the effect of the different approaches to phytoplankton backscattering.
Fig. 5
Fig. 5 Scattering and backscattering characteristics of EAP, Alvain [2] and Lee [3] at Chl a concentrations 1, 2, 5, 10, 15 and 30 mg.m−3. Phytoplankton and detrital backscatter (bottom right) are shown for Chl a of 1 (lower line) and 30 mg.m−3 (upper line). The increased spectral detail of the EAP model’s phytoplankton scatter and backscatter becomes increasingly important with increasing biomass as is evident in the spectral variation in the resulting total backscattering probability (bottom left).
Fig. 6
Fig. 6 Modelled EAP Rrs (above) for typical dinoflagellate assemblage with effective diameter of 16 μm. Chl a concentrations are 1, 2, 5, 10, 20, 30, 50, 100, 150, 200 mg.m−3. Below the shift from maximum peak reflectance height in the blue/green to the red is shown (dotted lines), for increasing Chl a. The first derivatives of these slopes (solid lines) cross at a Chl a of around 15 mg.m−3, the point at which the red features of high biomass reflectance spectra start to dominate.
Fig. 7
Fig. 7 High Biomass Validation of EAP Rrs, with EAP total backscatter probability shown for Chl a 1, 10, 50, 100, 150, 200 and 300 mg.m−3.
Fig. 8
Fig. 8 Contribution of Phytoplankton to total IOPs and Rrs signal, for Chl a values of 2 and 150 mg.m−3, for an idealised dinoflagellate assemblage with effective diameter of 16 μm.

Tables (2)

Tables Icon

Table 1 IOP parameterisations of the models of Alvain [2] and Lee [3]

Tables Icon

Table 2 Symbols and Abbreviations

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

a g d ( λ ) = a g d ( 400 ) exp [ S ( λ 400 ) ]
a g d ( 400 ) = 0.0904 log [ Chl a ] + 0.1287
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.