Skip to main content

ORIGINAL RESEARCH article

Front. Mar. Sci., 23 February 2018
Sec. Coastal Ocean Processes
Volume 4 - 2017 | https://doi.org/10.3389/fmars.2017.00437

Modeling Net Growth of Phaeocystis antarctica Based on Physiological and Optical Responses to Light and Temperature Co-limitation

  • 1Wallops Flight Facility, Goddard Space Flight Center (NASA), Wallops Island, VA, United States
  • 2Scripps Institution of Oceanography, University of California, San Diego, La Jolla, CA, United States

Temperature and light are fundamental environmental variables which regulate phytoplankton growth rates when nutrients are in excess. For polar coastal oceans that are undergoing changes in sea ice cover and warming, light, and temperature are particularly important for bloom dynamics. Using colonial Phaeocystis antarctica cultures grown at steady-state, we assessed the combined effect of these two environmental controls on net growth rate (μn), chlorophyll-specific absorption of light (aph* (λ)), and quantum yields for growth (ϕμ). Specific net growth rates (μn) varied from 0.04 to 0.34 day−1 within a matrix of light and temperature ranging from 14 to 542 μmol quanta m−2 s−1 and −1.5 to 4°C. Values of aph* (λ) varied significantly with light but only slightly with temperature. Values of ϕμ ranged from 0.003 to 0.09 mol C (mol quanta absorbed)−1 with highest values at low light and 4°C. For excess irradiances or low temperatures where growth rate is inhibited, quantum yields were low. The low ϕμ values are attributed both to increased absorption by photoprotective pigments compared to photosynthetic pigments and thermodynamic control of dark reaction enzymes. The systematic changes in photophysiological properties of P. antarctica in relation to temperature and light were used to develop a series of nested light- and temperature-dependent models for μn, aph* (λ), and ϕμ. A model for aph* (300–700 nm) was developed that takes into account the systematic changes in aph* (λ) due to pigment packaging effects and cellular concentrations of chlorophylls and photoprotective pigments. Also, a model for ϕμ was developed based on a cumulative one-hit Poisson probability function. These model parameterizations for absorption and quantum yield are combined into an overall model of net growth that can be applied easily to P. antarctica bloom dynamics using remote sensing data for temperature, light, and chlorophyll a. Furthermore, modeling based on the biophysical variables aph* (λ), and ϕμ that are shown to regulate the growth rate provides a more fundamental mechanistic approach compared to other modeling methods that do not explicitly resolve photon flux into the cell or the quantum yield.

Introduction

The colonial prymnesiophyte Phaeocystis can dominate coastal, ice edge, and open ocean blooms in polar and temperate waters with significant implications for carbon export (Smith et al., 1991; DiTullio et al., 2000; van Leeuwe et al., 2007; Pavlov et al., 2017). The importance of light, temperature, and nutrients in governing phytoplankton growth and bloom formation has motivated research on metabolic and cellular characteristics of phytoplankton in response to environmental stressors. Singly, the light-, temperature-, and nutrient- dependence of phytoplankton growth has been well-characterized for a variety of species under a wide range of experimentally controlled laboratory conditions providing the basis for generalized models of phytoplankton growth for temperature, nutrients, and light limitation. Nutrient-limited growth has been described as a function of cell quota which is determined by Michaelis-Menten uptake kinetics (Droop, 1983). These kinetic models have been modified to account for multiple nutrients by invoking Leibig's Law of the minimum (Falkowski and Raven, 2007). Temperature-dependent growth has been modeled with the Arrhenius equation (Li, 1980) and maximal growth has been based on empirical derivations determined from cultures (Eppley, 1972). Light-dependent growth has been modeled as a variety of different formulations which allows growth to increase with irradiance, saturate, and decrease at high irradiances (Goldman and Carpenter, 1974; Laws and Bannister, 1980; Moisan et al., 2002). While nutrient limitation is generally modeled on the basis of the “limiting” nutrient, phytoplankton must also acclimate to large variations in light and temperature (Behrenfeld et al., 2002; Ciotti et al., 2002). Often these regulate growth in an interactive co-limitation matrix (Sosik and Mitchell, 1994; Sunda and Huntsman, 1997, 2011) however there is lacking a mechanistic framework for modeling phytoplankton growth rate based on the biophysical variables of cellular absorption and quantum yield under simultaneous co-limitation.

Blooms of Phaeocystis sp. have been reported in diverse oceanic provinces for temperature ranging from less than 0°C to greater than 35°C and for both shallow and deep mixed layers (Gieskes et al., 2007; Gypens et al., 2007). Data sets required to specify physiological parameters for growth at various environmental conditions are relatively rare. A review of Chl:C ratios in polar and temperate microalgae at different acclimation states is found in Lacour et al. (2017)Various mathematical descriptions for modeling phytoplankton growth that include multiple limitations establish the foundations for our approach (Shuter, 1979; Laws and Bannister, 1980; Rhee and Gotham, 1981a,b; Geider et al., 1998; Behrenfeld et al., 2002; Edwards et al., 2016). These concepts have been extended to the estimates of primary production and growth rates from satellites (Behrenfeld et al., 2005; Arteaga et al., 2014, 2016). Growth rates regulated by irradiance, temperature, and nutrients have been modeled using different approaches based on algal metabolism (Laws and Bannister, 1980), elemental ratios of chlorophyll or carbon (Geider et al., 1998; Graff et al., 2016), or cell carbon quota (Zonneveld et al., 1997). Some of these models have limited ecological applicability because they require parameters that are not easily measured in natural systems. We model steady-state growth based on the mechanistic biophysical parameters of chl-a specific spectral absorption and photosynthetic quantum yeield (aph* (λ) and ϕμ,) which are directly regulated by light, temperature, and nutrients (Kiefer and Mitchell, 1983; Sakshaug et al., 1989; Cullen, 1990; Moisan and Mitchell, 1999). For brevity, all symbols are defined in Table 1. For steady-state growth, aph* (λ) and ϕμ vary in response to temperature, light, and nutrient regulation of growth (Mitchell and Kiefer, 1988; Nelson and Prézelin, 1990; Sosik and Mitchell, 1991, 1995; Moisan and Mitchell, 1999; Westberry et al., 2008). Combining estimates of phytoplankton absorption (Kishino et al., 1985) with rates of carbon fixation allows for validation of models based on aph* (λ) and ϕμ for natural communities (Cleveland et al., 1989; Smith et al., 1989; Marra et al., 1995; Sosik and Mitchell, 1995; Sosik, 1996). Thus, the model we describe here to predict aph* (λ) and ϕμ for light and temperature co-limitation can easily be applied and validated for natural communities in polar regions using routine methods and is most applicable when blooms dominated by Phaeocystis are forming and nutrients are in excess. The model is most likely not as robust during pre-bloom situations with a mixed community assemblage and for post bloom conditions when nutrients are exhausted.

TABLE 1
www.frontiersin.org

Table 1. Symbols used throughout the text.

Modeling growth rates of phytoplankton over large geographical scales will require parameterizations that take into account the taxonomic diversity of the phytoplankton community and variability driven by physiological acclimation (Moisan et al., 2017). In this paper, we have developed a model of growth rate using the colonial prymnesiophyte, P. antarctica, an ecologically and biogeochemically important organism in the polar and subpolar carbon cycle (Stuart et al., 2000; Smith et al., 2003; Pabi and Arrigo, 2006; Smith et al., 2006; Gieskes et al., 2007) and is responsible for significant amounts of new production (Arrigo et al., 1999). Temperature, light, and iron are major limiting factors that control the initiation and development of phytoplankton blooms in the Southern Ocean. We describe the range of variability and relationships of μn, aph* (λ), and ϕμ for P. antarctica for a realistic range of temperature, light, and nutrient conditions that occur during the Antarctic spring bloom. Using steady-state experimental data, we developed a series of nested models that together form a photophysiological model of net growth. Our model is based on contemporary bio-optical models for photosynthesis (Kiefer and Mitchell, 1983; Cullen, 1990; Moisan and Mitchell, 1999) which utilize aph* (λ) and ϕμ as photophysiological variables to describe phytoplankton net growth,

μn=chl aC350nm700nmaph (λ)E0(λ)ϕμ dλ.    (1a)

We determined the environmental control of aph* (λ), and ϕμ by assessing their magnitude for co-limitation using controlled light and temperature experiments. Our model for ϕμ is based on a cumulative one-hit Poisson probability distribution and introduces the concept that both light and temperature regulate the maximal quantum yield of growth for any acclimated state. In general, the maximal quantum yield for any acclimated state should be modeled as a function of the limiting growth conditions (Moisan and Mitchell, 1999) rather than using the theoretical maximum (Sakshaug et al., 1989). Our formulation is based on Cullen (1990) and we extend our previous model of ϕμ (Moisan and Mitchell, 1999) as a variable dependent on both light and temperature,

ϕμ=ϕmEo,T1expEo/EkμEo/Ekμ    (1b)

where ϕμ is the quantum yield of growth, ϕm,Eo,T is the maximal quantum yield for growth for an acclimated state within the light and temperature limitation matrix, E is the photoacclimation parameter (Falkowski and Raven, 2007), and E0 is photosynthetically available radiation.

The culture experiments were designed to assess the broad range of temperature and light that Antarctic phytoplankton experience in nature that is relevant to formation of blooms. Since a bloom cannot form unless there are excess nutrients to support the growth of the bloom, we do not model nutrients although clearly macronutrients and iron can regulate the crop size and productivity (Sedwick, 2007; van Leeuwe and Stefels, 2007). By characterizing the light and temperature response of phytoplankton absorption and quantum yield over a realistic environmental range, we have parameterized a bio-optical model for μn for application to ocean color satellites and buoys. This effort to describe and model the variability in the biophysical parameters that mediate temperature and light control of phytoplankton growth provides important details related to phytoplankton acclimation when multiple environmental variables are regulating growth simultaneously.

Materials and Methods

Cultures of colonial P. antarctica (CCMP 1374) were grown semi-continuously at −1.5, 0, 2, and 4°C for 5–8 generations in f/2 medium (Gulliard and Ryther, 1962) under continuous blue light ranging from 14 to 542 mol quanta m−2 s−1. The spectrum of the light source was similar to the light quality observed in the upper ocean (Moisan and Mitchell, 1999). Cultures were maintained at ~30% of the carrying capacity of the culture medium. This protocol results in a steady-state acclimated condition at the time of experimental sampling. Details of the illumination source, blue filter and culturing system and our method to estimate spectral quantum flux in the cultures can be found in Moisan and Mitchell (1999).

Specific Growth Rate (μn)

Specific growth rate was estimated by a linear regression of loge transformed daily determinations of in vivo fluorescence intensity (n = 2) measured with a Turner Model 10 fluorometer at the same time each day. POC data was also collected during mid-log phase growth at the time of the experiments, allowing an estimate of the carbon-based growth rate for each experimental culture.

Fluorometric Estimates of Chl a

Chl a concentrations (n = 3) were estimated fluorometrically using a Turner Model 10-AU fluorometer according to (Welschmeyer, 1994).

High Performance Liquid Chromatography (HPLC)

Duplicate samples (n = 2) were concentrated onto Whatman GF/F filters and extracted in cold 100% acetone by grinding with a teflon-tip grinding rod. After centrifugation, the supernatant was mixed with 20% (v/v) HPLC-grade water and analyzed on a Shimadzu LC10-AD HPLC system (Wright et al., 1991). Pigment concentrations were based on absorption at 440 nm (Dynamax Model UV-1). Integrated HPLC peak area was quantified with external standards. Canthaxanthin was used as an internal standard. A spectral diode detector provided spectral absorption of peaks which were used to confirm the individual pigments by reference to a spectral library of standards. Duplicate samples (N = 2) differed by less than 2% of each other on replicate filtered samples. Note that there was a small difference (2–20%) between HPLC samples and fluorometric samples. For this reason, we have averaged the HPLC and fluorometric samples.

Chl-Specific Absorption Coefficient (aph* (λ))

For experiments at −1.5, 2, and 4°C, in vivo whole cell absorption spectra were determined (n = 4) on concentrated samples from 300 to 800 nm at 1 nm intervals with an integrating sphere accessory in a dual beam Perkin Elmer Lambda 6 UV/Vis spectrophotometer. Fresh f/2 medium was used as a reference and a blank. For cultures grown at 0°C, the integrating sphere failed and we estimated the absorption coefficient with the glass fiber filter method of Mitchell (1990) which has been shown to agree well with the integrating sphere methods after corrections for pathlength amplification (Mitchell et al., 2000). The chl-specific absorption coefficient was estimated by dividing loge absorption, aph (λ), by the corresponding chl a value:

aph(λ)=aph (λ) [chl a]1.    (2)

Particulate Carbon and Nitrogen Content

Samples (n = 3) were filtered onto pre-combusted GF/F filters, stored at −20°C and run on a Carlo Erba carbon analyzer. Samples of acetanilide were used to validate the system's calibration. See Moisan and Mitchell (1999) for details.

Cell Concentrations

Cell concentrations were estimated in Palmer Maloney chambers both fresh (unfixed) and observed at a magnification of 400X. See Moisan and Mitchell (1999) for details.

Quantum Yield for Growth

Quantum yield for growth was based on the carbon-specific net growth rate, whole cell in vivo spectral absorption, and the spectral irradiance in each treatment (Sosik and Mitchell, 1991; Moisan and Mitchell, 1999) and is a rearrangement of Equation (1a).

Non-linear Curve Fitting

Coefficients for P-E curves, and temperature and light dependent equations for growth and quantum yield for growth were fit using a Marquardt-Levenberg non-linear least squares minimization routine (Marquardt, 1963). See Moisan and Mitchell (1999) for details.

Results and Discussion

The Relationship between Growth Rate, Temperature, and Light

Growth rates of P. antarctica varied ~8-fold ranging from 0.04 day−1 to 0.34 day−1 under conditions which were limiting to inhibiting for growth (Figure 1A). Specific net growth rates (μn) increased in response to light and approached photoinhibition at higher irradiances (Table 2). The initial slope of the light-dependent growth was lowest at 4°C (0.007 average± S.E. 0.001) and highest at 0°C (0.030 average ± S.E. 0.006). A photoacclimation parameter (E) was derived utilizing a curve-fitting scheme of all of the light treatments for a single temperature and other studies. Values of E were both temperature- and light- dependent (Figure 1A).

FIGURE 1
www.frontiersin.org

Figure 1. (A) Specific growth rates (μn, day−1) of P. antarctica with respect to irradiance for growth temperatures ranging from −1.5 to 4°C. (B) Relative growth rate (μnm) with respect to temperature for irradiances in units of μmol quanta m−2 s−1. (C) An empirically-derived model for growth rate based on combined limitation of light and temperature. The continuous surface is the model prediction and discrete points are experimental observations. (D) Modeled vs. observed μ (day−1) based on Equation (4).

TABLE 2
www.frontiersin.org

Table 2. Growth and pigmentation of P. antarctica grown under nutrient-replete conditions.

Relative growth rates (μnm) for our treatments exhibited a strong dependence on temperature, with maximal values between 0 and 2°C except at photoinhibiting intensities (Figure 1B, Tilzer and Dubinsky, 1987). In general, μn had a parabolic response to temperature, with a 1.5-fold change between minima and maxima with the exception of the treatments at 14 μmol quanta m−2 s−1. At low temperatures under moderate light intensities, P. antarctica achieved maximal growth similar to those grown at high light at 4°C. For example, the maximal growth rates, μm (0.98 ± 0.05, average ± S.E.) achieved at 37 μmol quanta m−2 s−1 at temperatures <2°C were lower than those observed at inhibiting growth conditions at 542 μmol quanta m−2 s−1 at 4°C (Figure 1A, t-test; p > 0.05).

Empirical Model of Growth Rate

The systematic dependence of net growth rate (μn) on temperature and light allowed for the derivation of a simple empirical model for μn (Figure 1C). We modeled the light-dependence of growth with a hyperbolic tangent equation commonly used for photosynthesis-irradiance relationships (Platt et al., 1980, Figure 1A) which allows μn to increase with irradiance and become saturated or photoinhibited at higher irradiances (Table 3). We modeled μm as a function of temperature (Eppley, 1972). Values of α were fit as linear functions of temperature and light (R2 = 0.85) as has been observed for Antarctic phytoplankton (Lizotte and Sullivan, 1991). Net growth for P. antarctica can be estimated with knowledge of light and temperature using the following equations,

μm=log(2)0.851(1.066T),    (3a)
α=α1+(α2T),    (3b)
μn=μsb[1exp(αPARμsb)]exp(βPARμsb)    (3c)
μsb=μmax[α+β/α][β/α+β]β/α    (3d)

where μn is net growth rate, μm is the maximal growth rate (day−1), T is temperature (°C), α is the initial slope of the μn vs. irradiance relationship, β is the inhibition of growth at high irradiance, α1 and α2 are the y-intercepts and slope for α with temperature, respectively, μsb (analogous to Psb in Platt et al., 1980) is a fitting parameter and PAR is photosynthetically available radiation (350–700 nm). Equations (3c,d) represent a growth equation originally derived by Platt et al. (1980) as a function of P vs. I. parameters. The values for α1 and α2 are 0.0069 ± 0.001 and (−0.001 ± 0.0003), respectively (p < 0.00001). We have very few treatments where irradiance was sufficient to inhibit growth. However, other investigators have clearly documented that phytoplankton growth is inhibited at higher irradiances (Goldman and Carpenter, 1974; Laws and Bannister, 1980; Platt et al., 1980). Therefore, we allow β to be constrained in a relatively large range to accomodate photoinhibition at high light under low temperature. These dynamics are in general agreement with increases in photoinhibition of short-term photosynthetic carbon fixation at high light (Platt et al., 1980).

TABLE 3
www.frontiersin.org

Table 3. Growth fitting parameters of Phaeocystis grown between −1.5 and 4°C.

This simple empirical model has an excellent ability for predicting μn based only on light and temperature as input (Figure 1D; r2 = 0.89, p < 0.05). However, it does not include explicit information about mechanisms of physiological acclimation. Therefore, we have also developed a model to predict growth using a series of nested models that predict the biophysical parameters, aph* (λ) and ϕμ, based on acclimation to temperature and light.

Temperature and Light Dependence of aph* (λ): Observations and Modeling of Variability

The in vivo whole cell chlorophyll-specific absorption properties, aph* (λ), of P. antarctica varied with both irradiance and temperature (Figures 2A,B). Values of aph* (λ) increased with increasing irradiance (Figures 2C–E). Values of aph* (λ) at 436 and 676 nm generally increased with decreasing temperature but the magnitude of the temperature dependence was less than the irradiance dependence (Figures 2C,D, ANOVA p > 0.05). Values of aph* (330), due to mycosporine-like amino acids (Moisan and Mitchell, 2001), had less dependence on temperature than estimates at 436 and 676 nm (Figure 2E). Our observations are within the range of aph* (λ) observed during Phaeocystis-dominated blooms (Cota et al., 1994; Arrigo et al., 1998; Reynolds et al., 2001).

FIGURE 2
www.frontiersin.org

Figure 2. Absorption characteristics of P. antarctica. In vivo chl-specific absorption spectrum (aph* (λ)) for P. antarctica grown at (A) 4C and (B)−1.5°C between a range of irradiances (μmol quanta m−2 s−1). Irradiance- and temperature- dependent values of aph* (circles) at (C) 436 nm (D) 676 nm and (E) 330 nm. Multiple linear regressions based on temperature and light (continuous surface) are overlaid on the discrete points of aph* (λ). Each point represents the average of 4 replicates. C.V. for aph* (λ) values are 5%.

Pigment package effects, changes in pigment ratios (Table 2), and thylakoid stacking played an important role in determining both the spectral shape and large changes in the magnitude of aph* (λ) (Moisan and Mitchell, 1999; Moisan et al., 2006). Ratios of aph (330) to aph (676) increased linearly from 1.8 to 13 due to changes in MAA to Chl a concentrations as a photo-protective response to increased irradiance (Moisan and Mitchell, 2001).

We explored the dependence of aph* (λ) on irradiance and temperature in an attempt to specify a simple parameterization for spectral absorption that could accurately capture the changes in both shape and magnitude caused by changes in pigment ratios and packaging. Singly, aph* (676) was highly predictable by a linear regression with either irradiance (r2 = 0.62, Figure 3A) or temperature (r2 = 0.63, data not shown). However, the accuracy of modeled estimates of aph* (λ) was improved by using a multiple linear regression with light and temperature,

aph(λ) = Co + CEPAR + CTT,    (4)

where PAR is in units of μmol quanta m−2 s−1, temperature (T) is in °C, Co is the intercept of the multiple linear regression, CE and CT are the multiple linear regression coefficients for PAR and T. The multiple regression fit for aph* (676) for temperature and light had an r2 = 0.82 (Figure 3B), a considerable improvement compared to the fit for only irradiance (r2 = 0.62, data not shown). A comparison of modeled (10 nm resolution) and observed (1 nm hyperspectral) aph* (λ) are shown in Figures 3C,D. Our modeled values are in excellent agreement with measured aph* (λ) in the visible region. Variability in aph* (λ) between 300 and 400 nm is dominated by irradiance (Moisan and Mitchell, 2001) with negligible dependence on temperature (data not shown). For consistency, a single parameterization based on PAR and T is used for the full spectrum (300–700 nm). Coefficients for the model based on light and temperature are given in Table 4. The model does an excellent job of describing aph* (λ) as illustrated for an extreme range of acclimation states representing high (Figure 3C) and low (Figure 3D) degrees of pigment packaging and changes in photosynthetic and photo-protective pigments relative to chlorophyll a. High relative error in the UV for the high light sample (Figure 3C) is caused in part because aph* (330) is ~10 times smaller than the low light example (Figure 3D) caused by changes in cellular concentrations of mycosporine-like amino acids relative to chlorophyll a (Moisan and Mitchell, 2001). There was insignificant irradiance in the system below 350 nm so the large variance in the aph* (λ) model for data between 300 and350 nm did not affect the estimates of ϕμ or μn; see section Performance of the Biophysical Model of Net Growth for P. antarctica for dertails on model application and Figures 5C,D for results modeling ϕμ and μn.

FIGURE 3
www.frontiersin.org

Figure 3. Modeled vs. observed values of aph* at 676 nm based on a linear regression using (A) light and (B) a multiple linear regression based on temperature and light. Comparison of aph* (λ) observed (1 nm resolution continuous line) vs. modeled values (10 nm resolution) based on temperature and light for cultures grown at (C) 4°C 542 μmol quanta m−2 s−1 and −1.5°C and (D) 37 μmol quanta m−2 s−1. Note scale difference in aph* in (C,D). Standard error bars are shown for model estimates.

TABLE 4
www.frontiersin.org

Table 4. Coefficients for multiple-linear regression model for aph* (λ) based on temperature and light.

Temperature and Light Dependence of the Quantum Yield for Growth: Observations and Modeling

The quantum yield for growth, ϕμ, is the molar ratio of carbon fixed for net growth to the quanta absorbed. Understanding how ϕμ varies as a function of environmental conditions that are easily measured either in situ or remotely is critical in the parameterization of photosynthetic bio-optical models that have ecological relevance. We have chosen to estimate ϕμ in relation to light and temperature because they limit bloom formation, control cellular absorption, and enzymatic activity. Although phytoplankton may exhibit chromatic adaptation that optimizes the absorbing pigments to selectively absorb the modified under water light spectrum (Dubinsky and Stambler, 2009), the quantum yield of photosynthesis (oxygen evolved or carbon fixed per photon absorbed) is relatively flat from about 400–650 nm (Luning and Dring, 1985). In the classic review Photosynthesis by Rabinowitch and Govindjee (1969) they show the spectral quantum yield determined by Emerson and Lewis (1943) for Chlorella pyrenoidosa ranged from 0.07 to 0.09 over the spectral range 400–680 nm with the lowest values in the area of photoprotective carotenoids 480–510 nm. While this classic work shows very modest spectral variation in the spectral quantum yield of ±15% from 400 to 680 nm, there are other articles that show larger variations more like ±30%, including our work on the chlorophyll-a fluorescence yield of P. antarctica (Moisan and Mitchell, 2001) with a similar dip in the photoprotective carotenoid band and very low yields below 350 nm where mycosporine amino acids absorb. The work by (Luning and Dring, 1985) for 4 genera of macro algae indicated quantum yield varied by ±15% from 400 to 670 nm. There is spectral variation in the quantum yield that we did not determine in this study, but for the PAR region below 680 nm the variations are modest, depending on which study is cited.

We spectrally quantified total photon absorption and total carbon growth so we do not have resolution of the spectral dependence of ϕμ. The quantum yield for growth at steady state for experimental samples was estimated by rearranging Equation (1a),

ϕμ=μnChl:C350nm700nmaph(λ)E0(λ)dλ.    (5)

For our study, ϕμ varied by 30-fold (Figure 4A) ranging from 0.003 to 0.09 mol C (mol quanta absorbed)−1. Values of ϕμ were highest at low irradiances and decreased at high light intensity but with a large difference in the relationship with light for different temperatures. For a given light level, ϕμ generally decreased with decreasing temperature (Figure 4A); the curve fitted to the 4°C was consistently higher than all other temperatures. Values of ϕμ for the 0 and 2°C treatments were similar to each other for the corresponding light treatments (t-test, p > 0.05). Our values for ϕμ are in the range previously reported for Phaeocystis-dominated blooms (SooHoo et al., 1987; Cota et al., 1994; Carder et al., 1995; Vaillancourt et al., 2003).

FIGURE 4
www.frontiersin.org

Figure 4. (A) Light and temperature dependence of ϕμ (mol C (mol quanta absorbed)−1) for P. antarctica (symbols). Our modeled ϕμ relationship with light and temperature (solid line) is calculated with a constant E and a modeled ϕmEo,T (Equation 6). (B) ϕmEo,T response to different temperature and irradiance treatments (Symbols are experimental observations, lines are model predictions). A fixed E value was used to calculate ϕmEo,T.

The quantum yield for net photosynthesis has previously been modeled as a product of the maximal quantum yield for net photosynthesis, ϕm and a Poisson probability function that an open photosynthetic unit will be hit. This model was originally formulated and validated with experiments on O2 evolution by short-term flashes of light (Dubinsky et al., 1986; Falkowski and Raven, 2007) and adopted to model the quantum yield for growth (Sakshaug et al., 1989; Cullen, 1990). Typically, ϕm has been set for all acclimation states to the theoretical maximal value which is equal to 0.125 mol C (mol quanta absorbed)−1 (e.g., Sakshaug et al., 1989). However, as reviewed by Laws et al. (2002), absorption by non-photosynthetic pigments will automatically result in quantum yields of photosynthesis less than the theoretical maximum. To address this important modeling issue, Moisan and Mitchell (1999) proposed a novel approach where the maximal quantum yield (ϕm) for each acclimated state was dependent on the irradiance. Here, we extend this concept of an environmentally controlled maximal quantum yield by also parameterizing it based on temperature which controls the thermodynamics of photosynthetic enzymes and therefore affects the quantum yield for growth. Thus, as we have shown for aph* above, here we present a mechanistic control of ϕm based on co-limitation by light and temperature.

We found lower quantum yield for growth values at low light (<20 μmol quanta m−2s−1) whereas the cumulative hit Poisson Probability does not allow this roll off at low light. Johnsen and Barber (2003)have shown excellent experimental data that demonstrates that there are different photophysiological mechanisms at reduced light levels which contribute to a reduced quantum yield. Despite low quantum yields for growth and photosynthesis, there is experimental evidence that photosynthetic efficiency is not always maximal at low irradiances (Johnsen and Barber, 2003). Low photosynthetic quantum yield has been attributed to the Kok effect, S-state decay (S3 to S2), and imbalances in photosystem excitation (Diner and Mauzerall, 1973). Furthermore, it is expected that the ratio of respiration to photosynthesis may increase at very low light, leading to the observed reduction in quantum yield at very low light.

Given observed values of ϕμ, a function for ϕmEo,T was derived which exponentially decreased from low to high irradiances (Figure 4B). The equations used to predict ϕmEo,T are given below and the final estimate of ϕμ compared to observed values are plotted in in Figure 5C. Values of ϕmEo,T were described as a function of irradiance (Figure 4B) and the temperature dependence was incorporated into the coefficients a and b which are modeled as linear functions of temperature (Table 5),

ϕmEo,T=aebEo,    (6a)
a=0.0125T + 0.058                r2=0.70     (6b)
b=0.0010T + 0.0075       r2=0.85    (6c)

where T is temperature in °C and Eo is in units of μmol quanta m−2 s−1. Values of a and b which are in Equation (6a) are linear relationships of parameters that rely on temperature. Values of ϕmEo,T and for Ekμ as estimated below were introduced into Equation (1b) to estimate ϕμ.

FIGURE 5
www.frontiersin.org

Figure 5. (A) The relationship between modeled aph* (λ) based on temperature and light (Equation 6, Table 4) and observed values at individual wavelengths. (B) Relative growth rates measured from various Phaeocystis species grown under a variety of temperature and irradiance regimes. Note that there are overlapping data points. (C) Modeled ϕμ vs. observed values. There are overlapping points. (D) Relationship between the observed carbon-specific growth rate and model predictions of carbon-specific rate of net growth (μn). Model is based on Equation (1) and Table 4.

TABLE 5
www.frontiersin.org

Table 5. Model variables for maximal quantum yield based on irradiance at discrete temperatures.

Performance of the Biophysical Model of Net Growth for P. antarctica

To test the performance of the biophysical model for net photosynthesis, we predicted aph* (λ) and ϕμ using the temperature and light dependent model parameterizations as described above. Overall, the predicted values of aph* (λ) underestimate the observed values by about 10% (Figures 3C,D, 5A) with greatest accuracy achieved in the region between 350 and 700 nm. Relatively more scatter was seen in the UV region between 300 and 350 nm. Figure 5B is a compilation of net growth normalized to the maximum value of the experiment vs. irradiance for the data presented here, Moisan and Mitchell (1999), and data from the literature for the genus Phaeocystis. To estimate an E value to introduce into Equation (1b), we used a non-linear curve fit (Platt et al., 1980) for the full data synthesis in Figure 5B. This fit resulted in a single value of μm and α for the full data set in Figure 5B that we used to estimate a generalized value of Em/α) for Phaeocystis spp. This resulted in a E value of 110 μmol quanta m−2 s−1 overall fit with r2 = 0.83 (Figure 5B). We also explored modeling using E determined from each individual growth vs. irradiance data set at the different temperatures (Figure 1A) but that did not improve the overall variance in our biophysical model so we kept the model simpler by using one generalized value determined from our data and literature synthesis (Figure 5B).

We predicted values of ϕμ using a cumulative one hit Poisson probability parameterization where ϕmEo,T in Equation (1b) is predicted from temperature and light (Equation 5). The ϕμ model also requires an input variable for E. Figure 5B combines data for Phaeocystis spp. presented in Moisan and Mitchell (1999) and the new data for −1.5, 0, and 2°C treatments that are presented here. Using the combined data set, we have derived an E. Our sensitivity analysis shows that the ϕμ is the most sensitive to ϕmEo,T as compared to E.We prefer the use of a composite E at this time although in principle it should depend on environmental acclimation, in this case the steady-state conditions we imposed in our experiment. Although we acknowledge that in nature both light and temperature may fluctuate on time scales faster than division rates and hence impact acclimation, our experiments were not designed to resolve that additional, and potentially important, variability. We conducted a sensitivity analysis of ϕμ that indicates that the relative magnitude of the error is greatest at high intensities; however, the absolute error is minor. These assumptions are justified on the basis that the entire equation is dependent on φmax. We do observe some variability in E between the lower temperature treatments and the 4°C treatment (Figure 1A). We derived a composite E from the intersection between the initial slope of the growth curve and the maximal growth rate (Figure 5B). The use of a generic E is appropriate because of the overall variability in growth curves with respect to irradiance for various Phaeocystis spp. Also, a sensitivity analysis of E within the model revealed that ϕμ was relatively insensitive to the changes that we observed in E. Using an E value of 110 μmol quanta m−2 s−1 and a temperature-light dependent model of ϕmEo,T results in an excellent prediction of ϕμ (Figure 5C, r2 = 0.80). Using these individual component models for aph* (λ) and ϕmEo,T in Equations (1a,b), a comparison between observed and predicted values of μn (Figure 5D; r2 = 0.66). The simple empirical model using only temperature and light resulted in a slightly better estimate (r2 = 0.70; Figure 1C) however that empirical approach does not include the mechanistic dynamics of changes in aph* (λ) and ϕmEo,T that represent how light and temperature co-limitation regulate the cellular physiology.

Although we used a single value of E a sensitivity analysis proved that using the estimated E for the different temperatures did not improve our model fit. Also, our quantum yields for growth are a net process of growth after respiration which can be important in nature (Grégori et al., 2002; Siegel et al., 2002; Marra and Barber, 2004; López-Urrutia et al., 2006). Our quantum yield for net growth calculations can be defined by:

ϕμn=μnCEabsorbed=PREabsorbed.

Although respiration is often modeled as a fixed ratio to growth rate, or a basal a rate plus a fixed fraction of growth rate, this likely is not so simple. There are relatively few data sets in the literature for phytoplankton respiration over a range of growth conditions and we did not estimate respiration for our study. Respiration data that is commonly used in models is very noisy (Laws and Bannister, 1980; Geider et al., 1998) and is mostly limited to a modest range of intensity from about 50–200 μmol quanta m−2 s−1 so the fidelity of respiration models may not be robust over the full range of light intensity that is relevant to phytoplankton communities. One expects that the ratio of respiration to net growth (R/μn) will actually go to infinity at the limit as light goes to zero where there is no photosynthesis but basal respiration continues. Also, for the growth inhibiting irradiances greater than about 150 μmol quanta m−2 s−1 we observed at lower temperatures (Figure 1A), we speculate that the ratio of R/μn also must increase. Models that assume a basal respiration rate plus a fixed fraction of growth rate will not predict this increase in R/μn at higher growth inhibiting irradiances but will instead predict lower respiration as growth rate decreases. This seems flawed since inhibiting irradiances would cause damage from oxidative stress and hence very likely a greater respiration rate, not a lower rate. These two thought experiments based on simple logic indicate that it is important to obtain more data that includes respiration at very low light that and also at high growth inhibiting irradiances, but such data is lacking in the literature.

Our biophysical model includes explicit descriptions of the regulation of absorption and quantum yield by environmental controls which results both in an accurate model of net growth and captures details of light and temperature regulation of the biophysical variables that mediate light utilization. Our quantum yields for growth were calculated in a traditional way of modeling based on carbon, chlorophyll a, phytoplankton absorption, and spectral irradiance flux. However, our estimates are relatively high compared to other phytoplankton; our high C: chl a ratios are above the typical ratio of 75–150 for phytoplankton (Falkowski and Raven, 2007). This is likely due to the fact that Phaeocystis can accumulate a carbohydrate/glucan-rich mucilage that is not metabolically active in addition to metabolically active cellular carbon. Mathot et al. (2000) estimates that mucus-related carbon at peak bloom is about 14 ± 4% of total colony carbon but mucus-related carbon may be as high as 90% for a senescent bloom (Verity and Medlin, 2003). The colonial matrix acts as a spillover mechanism probably for storage of low molecular weight compounds (Lancelot and Mathot, 1985) and hence may accumulate when there is excess light.

Although natural systems may have substantial variations in both light and temperature over time-scales that are much less than the doubling times that will affect acclimation of algae in a more complicated way than the simple steady-state experiments we conducted, it is very time consuming to try to add that complexity to the matrix of growth and acclimation control. There is a relatively small literature for the steady-state co-limitation of light, nutrients and temperature acclimation of cellular pigments, absorption, quantum yield and growth (Laws and Bannister, 1980; Kiefer and Mitchell, 1983; Sosik and Mitchell, 1991, 1994; Geider et al., 1997). There is even more limited knowledge for cold water obligate psychrophiles like P. antarctica, for example the study of Luxem et al. (2017) who explored light and iron limitation for this organism. In the area of applied phycology aimed at commercial production, recent work with a similar experimental design for light and temperature as ours, for the seaweed Gracilaria, Nejrup et al. (2013), reported growth rates and metabolic capacity. Grobbelaar et al. (1996) have explored microalgae acclimated to different light levels, then exposed to light fluctuations that would be found in commercial cultivation. However, neither of these studies quantified spectral absorption or quantum yield, that ultimately underlie the responses for growth and bulk PvsE relationships. Our work advances the state of knowledge with greater detail of how light and temperature combined affect acclimation of the fundamental biophysical parameters light absorption and quantum yield of algae and provides a level of detail not previously reported for the ecologically important prymnesiophyte, Phaeocystis antarctica, that is a key structural component of the Southern Ocean and that may show changes in its role as stratification and heating of the Southern Ocean change both light and temperature climate in the future.

Conclusions

Understanding the contribution of Phaeocystis in the overall global carbon cycle is important because this genus can dominate mesoscale blooms and make a significant contribution to the carbon and sulfur cycle for open and coastal oceans as well in the vicinity of ice edges. The strain of P. antarctica we used for these experiments exhibits flexibility in its ability to absorb light due to changes in cell size and the cellular concentrations of photoprotective and photosynthetic pigments that are regulated by light and temperature co-limitation. These cellular changes have consequences for how P. antarctica captures light required for carbon fixation. Photo-protective pigments that divert absorbed photons from the photosynthetic reaction centers will automatically result in photosynthetic quantum yields below the theoretical maximum. Using data from our controlled laboratory experiments, we have developed both empirical and biophysical models that predict growth rates based on knowledge of chlorophyll a, light, and temperature. It is important to note that unlike some proposed models that depend only on easily measured variables our biophysical model, aph* (λ) and ϕμ is parameterized based on light and temperature since nutrients are not regulating during the onset and development of blooms. We propose a conceptually important improvement over previous models by accepting the fact that the maximum quantum yield for any acclimated state may be considerably less than the theoretical maximum (Moisan and Mitchell, 1999) and that it should be regulated by environmental control, in this case light and temperature co-regulation. In general, aph* (λ), Ekm and ϕμ, should all be modeled as variables determined by the simultaneous interaction of light, temperature, and nutrients. Furthermore, future ocean color missions with hyperspectral ability, such as PACE, may be capable of remote sensing phytoplankton spectral shape that we document changes significantly with different acclimation states (Figure 2). Such a capability might allow a link between community structure, photosynthetic processes and the biophysical model for absorption, cellular pigmentation and pigment packaging that is dependent on light and temperature co-limitation (see also Wang et al., 2016; Moisan et al., 2017). While this level of detail remains aspirational for our community, an accurate understanding of how biophysical variables (Equation 1) depend on the full matrix of environmental forcing requires these essential but time-consuming laboratory studies for representative taxa over the full range of environmental variability experienced by phytoplankton to advance our ability to apply advanced remote sensing methods to coastal processes of primary production and the carbon cycle.

Author Contributions

TM: Supervised all experiment, sample analysis and data analysis; TM and BGM: Contributed equally to writing, editing, graphics, and tables.

Funding

Support for this work was provided by ONR grant N00014-91-J-1186 and NASA Award # NNX15AC98G (BGM) and the NASA Biodiversity Program (05-TEB/05-0016) (TM).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank S. Cheng for excellent technical assistance; M. Vernet for insightful recommendations throughout this study and comments on earlier versions of the manuscript; M. Olaizola and R. Goericke for assistance with HPLC analysis and thoughtful discussions regarding pigment data; N. Delaney, N. Crawford, M. Cape, and M. Linkswiler for programming assistance; and M. Anderson for final manuscript preparation.

References

Arrigo, K. R., Robinson, D. H., Worthen, D. L., Dunbar, R. B., DiTullio, G. R., VanWoert, M., et al. (1999). Phytoplankton community structure and the drawdown of nutrients and CO2 in the southern ocean. Science 283, 365–367. doi: 10.1126/science.283.5400.365

PubMed Abstract | CrossRef Full Text | Google Scholar

Arrigo, K. R., Worthen, D., Schnell, A., and Lizotte, M. P. (1998). Primary production in Southern Ocean waters. J. Geophys. Res. 103, 15587–15600.

Google Scholar

Arteaga, L., Pahlow, M., and Oschlies, A. (2014). Global patterns of phytoplankton nutrient and light co-limitation inferred from an optimality GÇÉ based model. Glob. Biogeochem. Cycle 28, 648–661. doi: 10.1002/2013GB004668

CrossRef Full Text | Google Scholar

Arteaga, L., Pahlow, M., and Oschlies, A. (2016). Modeled Chl: C ratio and derived estimates of phytoplankton carbon biomass and its contribution to total particulate organic carbon in the global surface ocean. Glob. Biogeochem. Cycle 30, 1791–1810. doi: 10.1002/2016GB005458

CrossRef Full Text | Google Scholar

Behrenfeld, M. J., Boss, E., Siegel, D. A., and Shea, D. M. (2005). Carbon-based ocean productivity and phytoplankton physiology from space. Glob. Biogeochem. Cycle 19:GB1006. doi: 10.1029/2004GB002299

CrossRef Full Text | Google Scholar

Behrenfeld, M. J., Marañón, E., Siegel, D. A., and Hooker, S. B. (2002). Photoacclimation and nutrient-based model of light-saturated photosynthesis for quantifying oceanic primary production. Mar. Ecol. Prog.Ser. 228, 103–117. doi: 10.3354/meps228103

CrossRef Full Text | Google Scholar

Carder, K. L., Lee, Z. P., Marra, J., Steward, R. G., and Perry, M. J. (1995). Calculated quantum yield of photosynthesis of phytoplankton in the marine light-mixed layers (59oN, 21oW). J. Geophys. Res. 100, 6655–6663.

Google Scholar

Ciotti, A. M., Lewis, M. R., and Cullen, J. J. (2002). Assessment of the relationships between dominant cell size in natural phytoplankton communities and the spectral shape of the absorption coefficient. Limnol. Oceanogr. 47, 404–417. doi: 10.4319/lo.2002.47.2.0404

CrossRef Full Text | Google Scholar

Cleveland, J. S., Perry, M. J., Kiefer, D. A., and Talbot, M. C. (1989). Maximum quantum yield of photosynthesis in the northwestern Sargasso Sea. J. Mar. Res. 47, 869–892. doi: 10.1357/002224089785076055

CrossRef Full Text | Google Scholar

Cota, G. F., Smith, W. O. Jr., and Mitchell, B. G. (1994). Photosynthesis of Phaeocystis in the Greenland Sea. Limnol. Oceanogr. 39, 948–953. doi: 10.4319/lo.1994.39.4.0948

CrossRef Full Text | Google Scholar

Cullen, J. J. (1990). On models of growth and photosynthesis in phytoplankton. Deep Sea Res. Part A 37, 667–683.

Google Scholar

Diner, B., and Mauzerall, D. (1973). Feedback controlling oxygen production in a cross-reaction between two photosystems in photosynthesis. Biochim. Biophys. Acta 305, 329–352.

PubMed Abstract | Google Scholar

DiTullio, G. R., Grebmeier, J. M., Arrigo, K. R., Lizotte, M. P., Robinson, D. H., Leventer, A., et al. (2000). Rapid and early export of Phaeocystis antarctica blooms in the Ross Sea, Antarctica. Nature 404, 595–598. doi: 10.1038/35007061

PubMed Abstract | CrossRef Full Text | Google Scholar

Droop, M. R. (1983). 25 years of algal growth kinetics: a personal view. Bot. Mar. 26, 99–112.

Google Scholar

Dubinsky, Z., Falkowski, P. G., and Wyman, K. (1986). Light harvesting and utilization by phytoplankton. Plant Cell Physiol. 27, 1335–1349. doi: 10.1093/oxfordjournals.pcp.a077232

CrossRef Full Text | Google Scholar

Dubinsky, Z. V., and Stambler, N. (2009). Photoacclimation processes in phytoplankton: mechanisms, consequences, and applications. Aquat. Microb. Ecol. 56, 163–176. doi: 10.3354/ame01345

CrossRef Full Text | Google Scholar

Edwards, K. F., Thomas, M. K., Klausmeier, C. A., and Litchman, E. (2016). Phytoplankton growth and the interaction of light and temperature: a synthesis at the species and community level. Limnol. Oceanogr. 61, 1232–1244. doi: 10.1002/lno.10282

CrossRef Full Text | Google Scholar

Emerson, R., and Lewis, C. M. (1943). The dependence of the quantum yield of photosynthesis on the wavelength of light. Am. J. Bot. 30, 165–178.

Google Scholar

Eppley, R. W. (1972). Temperature and phytoplankton growth in the sea. Fish. Bull. Nat. Ocean. Atmos. Adm. 70, 1063–1085.

Google Scholar

Falkowski, P. G., and Raven, J. (2007). Aquatic Photosynthesis, 2nd Edn. Princeton, NJ: Princeton University Press.

Google Scholar

Geider, R. J., MacIntyre, H. L., and Kana, T. M. (1997). Dynamic model of phytoplankton growth and acclimation: responses of the balanced growth rate and the chlorophyll a: carbon ratio to light, nutrient-limitation and temperature. Mar. Ecol. Prog. Ser. 148l, 187–200.

Google Scholar

Geider, R. J., MacIntyre, H. L., and Kana, T. M. (1998). A dynamic regulatory model of phytoplanktonic acclimation to light, nutrients, and temperature. Limnol. Oceanogr. 43, 679–694.

Google Scholar

Gieskes, W. W. C., Leterme, S. C., Peletier, H., Edwards, M., and Reid, P. C. (2007). Phaeocystis colony distribution in the North Atlantic Ocean since 1948, and interpretation of long-term changes in the Phaeocystis hotspot in the North Sea. Biochemistry 83, 49–60. doi: 10.1007/s10533-007-9082-6

CrossRef Full Text | Google Scholar

Goldman, J. C., and Carpenter, E. J. (1974). A kinetic approach to the effect of temperature on algal growth. Limnol. Oceanogr. 19, 756–766.

Google Scholar

Graff, J. R., Westberry, T. K., Milligan, A. J., Brown, M., Dall'Olmo, G., Reifel, K. M., et al. (2016). Photoacclimation of natural phytoplankton communities. Mar. Ecol. Prog. Ser. 542, 51–62. doi: 10.3354/meps11539

CrossRef Full Text | Google Scholar

Grégori, G., Denis, M., Lefèvre, D., and Beker, B. (2002). A flow cytometric approach to assess phytoplankton respiration. Methods Cell Sci. 24, 99–106. doi: 10.1023/A:1024110418448

PubMed Abstract | CrossRef Full Text | Google Scholar

Grobbelaar, J. U., Nedbal, L., and Tichy, V. (1996). Influence of high frequency light/dark fluctuations on photosynthetic characteristics of microalgae photoacclimated to different light intensities and implications for mass algal cultivation. J. Appl. Phycol. 8, 335–343. doi: 10.1007/BF02178576

CrossRef Full Text | Google Scholar

Gulliard, R. R., and Ryther, J. H. (1962). Studies on marine planktonic diatoms. I. Cyclotella nana Hustedt and Detonula confervacea (Cleve) Gran. Can. J. Microbiol. 8, 229–239.

Google Scholar

Gypens, N., Lacroix, G., and Lancelot, C. (2007). Causes of variability in diatom and Phaeocystis blooms in Belgian coastal waters between 1989 and 2003: a model study. J. Sea Res. 57, 19–35. doi: 10.1016/j.seares.2006.07.004

CrossRef Full Text | Google Scholar

Johnsen, Z., and Barber, R. T. (2003). The low-light reduction in the quantum yield of photosynthesis: potential errors and biases when calculating the maximal quantum yield. Photosyn. Res. 75, 85–95. doi: 10.1023/A:1022440305765

CrossRef Full Text | Google Scholar

Kiefer, D. A., and Mitchell, B. G. (1983). A simple steady state description of phytoplankton growth based on absorption cross section and quantum efficiency. Limnol. Oceanogr. 28, 770–776.

Google Scholar

Kishino, M., Takahashi, M., Okami, N., and Ichimura, S. (1985). Estimation of spectral absorption coefficients of phytoplankton in the sea. Bull. Mar. Sci. 37, 634–642.

Google Scholar

Lacour, T., Lariviere, J., and Babin, M. (2017). Growth, Chl a content, photosynthesis, and elemental composition in polar and temperate microalgae. Limnol. Oceanogr. 1, 43–48. doi: 10.1002/lno.10369

CrossRef Full Text | Google Scholar

Lancelot, C., and Mathot, S. (1985). Biochemical fractionation of primary production by phytoplankton in Belgian coastal waters during short- and long-term incubations with C-bicarbonate II. Phaeocystis poucheti colonial population. Mar. Biol. 86, 227–232.

Google Scholar

Laws, E. A., and Bannister, T. T. (1980). Nutrient- and light-limited growth of Thalassiosira fluviatilis in continuous culture, with implications for phytoplankton growth in the ocean. Limnol. Oceanogr. 25, 457–473.

Google Scholar

Laws, E., Sakshaug, E., Babin, M., Dandonneau, Y., Falkowski, P., Geider, R., et al. (2002). Photosynthesis and primary productivity in marine ecosystems: Practical aspects and application of techniques. Scientific Committee on Oceanic Research, JGOFS Report No. 36, 1–89.

Google Scholar

Li, W. K. W. (1980). “Temperature adaptation in phytoplankton: cellular and photosynthetic characteristics,” in Primary Productivity in the Sea, ed P. G. Falkowski (New York, NY: Plenum), 259–279.

Google Scholar

Lizotte, M. P., and Sullivan, C. W. (1991). Photosynthesis-irradiance relationships in microalgae associated with Antarctic pack ice: evidence for in situ activity. Mar. Ecol. Prog. Ser. 71, 175–184. doi: 10.3354/meps071175

CrossRef Full Text | Google Scholar

López-Urrutia, A., San Martin, E., Harris R. P., and Irigoien, X. (2006). Scaling the metabolic balance of the oceans. Proc. Natl. Acad. Sci. U.S.A. 103, 8739–8744. doi: 10.1073/pnas.0601137103

PubMed Abstract | CrossRef Full Text | Google Scholar

Luning, K., and Dring, M. J. (1985). Action spectra and spectral quantum yield of photosynthesis in marine macroalgae with thin and thick thalli. Mar. Biol. 87, 119–129. doi: 10.1007/BF00539419

CrossRef Full Text | Google Scholar

Luxem, K. E., Ellwood, M. J., and Strzepek, R. F. (2017). Intraspecific variability in Phaeocystis antarctica's response to iron and light stress. PLoS ONE 12:e0179751. doi: 10.1371/journal.pone.0179751

PubMed Abstract | CrossRef Full Text | Google Scholar

Marquardt, D. W. (1963). An algorithm for least squares estimations of parameters. J. Soc. Ind. Appl. Math. 11, 431–441.

Google Scholar

Marra, J., and Barber, R. T. (2004). Phytoplankton and heterotrophic respiration in the surface layer of the ocean. Geophys. Res. Lett. 31:L09314. doi: 10.1029/2004GL019664

CrossRef Full Text | Google Scholar

Marra, J., Langdon, C., and Knudson, C. A. (1995). Primary production, water column changes, and the demise of Phaeocystis bloom at the Marine Light-Mixed Layers site (59oN, 21oW) in the northeast Atlantic Ocean. J. Geophys. Res. 100, 6633–6643.

Google Scholar

Mathot, S., Smith, W. O. Jr, Carlson, C. A., Garrison, D. L., Gowing, M. M., and Vickers, C. L. (2000). Carbon partitioning within Phaeocystis antarctica (Prymnesiophyceae) colonies in the Ross Sea, Antarctica. J. Phycol. 36, 1049–1056. doi: 10.1046/j.1529-8817.2000.99078.x

CrossRef Full Text | Google Scholar

Mitchell, B. G. (1990). “Algorithms for determining the absorption coefficient of aquatic particulates using the quantitative filter technique (QFT),” in Ocean Optics X, ed R. Spinrad (Bellingham, WA: SPIE), 37–148.

Mitchell, B. G., Bricaud, A., Carder, K., Cleveland, J., Ferrari, G., Gould, R., et al. (2000). “Determination of spectral absorption coefficients of particles, dissolved material and phytoplankton for discrete water samples,” in Ocean Optics Protocols For Satellite Ocean Color Sensor Validation (Greenbelt, MD: NASA/TM-2000-209966), 125–153.

Google Scholar

Mitchell, B. G., and Kiefer, D. A. (1988). Variability in pigment specific particulate fluorescence and absorption spectra in the northeastern Pacific Ocean. Deep Sea Res. Part I 35, 665–689.

Google Scholar

Moisan, J. R., Moisan, T. A., and Abbot, M. R. (2002). Modeling the effects of temperature on the maximum growth rates of phytoplankton growth. Ecol. Model. 153, 197–215. doi: 10.1016/S0304-3800(02)00008-X

CrossRef Full Text | Google Scholar

Moisan, T. A., Ellisman, M. H., Buitenhuys, C. W., and Sosinsky, G. E. (2006). Differences in chloroplast ultrastructure of Phaeocystis antarctica in low and high light. Mar. Biol. 149, 1281–1290. doi: 10.1007/s00227-006-0321-5

CrossRef Full Text | Google Scholar

Moisan, T. A., and Mitchell, B. G. (1999). Photophysiological adaptation of Phaeocystis antarctica Karsten under PAR light limitation. Limnol. Oceanogr. 44, 247–258. doi: 10.4319/lo.1999.44.2.0247

CrossRef Full Text | Google Scholar

Moisan, T. A., and Mitchell, B. G. (2001). UV absorption by mycosporine-like amino acids in Phaeocystis antarctica Karsten. Mar. Biol. 138, 217–227. doi: 10.1007/s002270000424

CrossRef Full Text | Google Scholar

Moisan, T. A., Rufty, K. M., Moisan, J. R., and Linkswiler, M. A. (2017). Satellite observations of phytoplankton functional type spatial distributions, phenology, diversity, and ecotones. Front. Mar. Sci. 4:189. doi: 10.3389/fmars.2017.00189

CrossRef Full Text | Google Scholar

Nejrup, L. B., Staehr, P. A., and Thomsen, M. S. (2013). Temperature- and light-dependent growth and metabolism of the invasive red algae Gracilaria vermiculophylla- a comparison with two native macroalgae. Eur. J. Phycol. 48, 295–308. doi: 10.1080/09670262.2013.830778

CrossRef Full Text | Google Scholar

Nelson, N. B., and Prézelin, B. B. (1990). Chromatic light effects and physiological modeling of absorption properties of Heterocapsa pygmaea (= Glenodinium sp.). Mar. Ecol. Prog. Ser. 63, 37–46.

Google Scholar

Pabi, S., and Arrigo, K. R. (2006). Satellite estimation of marine particulate organic carbon in waters dominated by different phytoplankton taxa. J. Geophys. Res. 111:C09003,

Google Scholar

Pavlov, A. K., Taskjelle, T., Kauko, H. M., Hamre, B., Hudson, S. R., Assmy, P., et al. (2017). Altered inherent optical properties and estimates of the underwater light field during an Arctic under-ice bloom of Phaeocystis pouchetii. J. Geophys. Res. Oceans 122, 4939–4961. doi: 10.1002/2016JC012471

CrossRef Full Text | Google Scholar

Platt, T., Gallegos, C. L., and Harrison, W. G. (1980). Photoinhibition of photosynthesis in natural assemblages of marine phytoplankton. J. Mar. Res. 38, 687–701.

Google Scholar

Rabinowitch, E., and Govindjee, R. (1969). Photosynthesis. New York, NY: J.Wiley.

Google Scholar

Reynolds, R. A., Stramski, D., and Mitchell, B. G. (2001). A chlorophyll-dependent semi-analytical reflectance model derived from field measurements of absorption and backscattering coefficients within the Southern Ocean. J. Geophys. Res. 106, 7125–7138. doi: 10.1029/1999JC000311

CrossRef Full Text | Google Scholar

Rhee, G. Y., and Gotham, I. J. (1981a). The effect of environmental factors on phytoplankton growth: light and the interactions of light with nitrate limitation. Limnol. Oceanogr. 26, 649–659.

Google Scholar

Rhee, G. Y., and Gotham, I. J. (1981b). The effect of environmental factors on phytoplankton growth: temperature and the interactions of temperature with nutrient limitation. Limnol. Oceanogr. 26, 635–648.

Google Scholar

Sakshaug, E., Andresen, K., and Kiefer, D. A. (1989). A steady state description of growth and light absorption in the marine planktonic diatom Skeletonema costatum. Limnol. Oceanogr. 34, 198–205. doi: 10.4319/lo.1989.34.1.0198

CrossRef Full Text | Google Scholar

Sedwick, P. N. (2007). Evidence of high iron requirements of colonial Phaeocystis antarctica at low irradiance. Biochemistry 83, 83–97. doi: 10.1007/s10533-007-9081-7

CrossRef Full Text | Google Scholar

Shuter, B. (1979). A model of physiological adaptation in unicellular algae. J. Theor. Biol. 78, 519–552.

PubMed Abstract | Google Scholar

Siegel, D. A., Doney, S. C., and Yoder, J. A. (2002). The North Atlantic spring Phytoplankton bloom and Sverdrup's critical depth hypothesis. Science 296, 730–733. doi: 10.1126/science.1069174

PubMed Abstract | CrossRef Full Text | Google Scholar

Smith, R. C., Prézelin, B. B., Bidigare, R. R., and Baker, K. S. (1989). Bio-optical modeling of photosynthetic production in coastal waters. Limnol. Oceanogr. 34, 1524–1544.

Google Scholar

Smith, W. O. Jr., Codispoti, L. A., Nelson, D. M., Manley, T., Buskey, E. J., Niebauer, H. J., et al. (1991). Importance of Phaeocystis blooms in the high-latitude ocean carbon cycle. Nature 352, 514–516.

Google Scholar

Smith, W. O., Dennett, M. R., Mathot, S., and Caron, D. A. (2003). The temporal dynamics of the flagellated and colonial stages of Phaeocystis antarctica in the Ross Sea. Deep Sea Res. Part II 50, 605–617. doi: 10.1016/S0967-0645(02)00586-6

CrossRef Full Text | Google Scholar

Smith, W. O., Shields, A. R., Peloquin, J. A., Catalano, G., Tozzi, S., Dinniman, M. S., et al. (2006). Interannual variations in nutrients, net community production, and biogeochemical cycles in the Ross Sea. Deep Sea Res. Part II 53, 815–833. doi: 10.1016/j.dsr2.2006.02.014

CrossRef Full Text | Google Scholar

SooHoo, J. B., Palmisano, A. C., Kottmeier, S. T., Lizotte, M. P., SooHoo, S. L., and Sullivan, C. W. (1987). Spectral light absorption and quantum yield of photosynthesis in sea ice microalgae and a bloom of Phaeocystis pouchetii from McMurdo Sound, Antarctica. Mar. Ecol. Prog. Ser. 39, 175–189.

Google Scholar

Sosik, H. M. (1996). Bio-optical modeling of primary production: consequences of variability in quantum yield and specific absorption. Mar. Ecol. Prog. Ser. 143, 225–238.

Google Scholar

Sosik, H. M., and Mitchell, B. G. (1991). Absorption, fluorescence and quantum yield for growth in nitrogen limited Dunaliella tertiolecta. Limnol. Oceanogr. 36, 910–921. doi: 10.4319/lo.1991.36.5.0910

CrossRef Full Text | Google Scholar

Sosik, H. M., and Mitchell, B. G. (1994). Effects of temperature on growth, light absorption, and quantum yield in Dunalliela tertiolecta (chlorophyceae). J. Phycol. 30, 833–840.

Google Scholar

Sosik, H. M., and Mitchell, B. G. (1995). Light absorption by phytoplankton, photosynthetic pigments, and detritus in the California current system. Deep Sea Res. Part I 42, 1717–1748. doi: 10.1016/0967-0637(95)00081-G

CrossRef Full Text | Google Scholar

Stuart, V., Sathyendranath, S., Head, E. J. H., Platt, T., Irwin, B., and Maass, H. (2000). Bio-optical characteristics of diatoms and prymnesiophyte populations in the Labrador Sea. Mar. Ecol. Prog. Ser. 201, 91–106. doi: 10.3354/meps201091 (Sverdrup's Critical Depth Hypothesis. Science 296, 730–733).

CrossRef Full Text | Google Scholar

Sunda, W. G., and Huntsman, S. A. (1997). Interrelated influence of iron, light and cell size on marine phytoplankton growth. Nature 390, 389–392. doi: 10.1038/37093

CrossRef Full Text | Google Scholar

Sunda, W. G., and Huntsman, S. A. (2011). Interactive effects of light and temperature on iron limitation in a marine diatom: implications for marine productivity and carbon cycling. Limnol. Oceanogr. 56, 1475–1488. doi: 10.4319/lo.2011.56.4.1475

CrossRef Full Text | Google Scholar

Tilzer, M. M., and Dubinsky, Z. (1987). Effects of temperature and day length on the mass balance of Antarctic Phytoplankton. Polar Biol. 7, 35–42.

Google Scholar

Vaillancourt, R. D., Marra, J., Barber, R. T., and Smith, W. O. (2003). Primary productivity and in situ quantum yields in the Ross Sea and Pacific Sector of the Antarctic circumpolar current. Deep Sea Res. Partt. II 50, 559–578. doi: 10.1016/S0967-0645(02)00584-2

CrossRef Full Text | Google Scholar

van Leeuwe, M. A., and Stefels, J. (2007). Photosynthetic responses in Phaeocystis antarctica towards varying light and iron conditions. Biochemistry 83, 61–70. doi: 10.1007/s10533-007-9083-5

CrossRef Full Text | Google Scholar

van Leeuwe, M. A., Stefels, J., Belviso, S., Lancelot, C., Verity, P. G., and Gieskes, W. W. C. (2007). Phaeocystis, Major Link in the Biogeochemical Cycling of Climate-Relevant Elements. Dordrecht: Springer.

Google Scholar

Verity, P. G., and Medlin, L. K. (2003). Observations on colony formation by the cosmopolitan phytoplankton genus Phaeocystis. J. Mar. Syst. 43, 153–164. doi: 10.1016/j.jmarsys.2003.09.001

CrossRef Full Text | Google Scholar

Wang, G., Lee, Z. P., Mishra, D. R., and Ma, R. (2016). Retrieving absorption coefficients of multiple phytoplankton pigments from hyperspectral remote sensing reflectance measured over cyanobacteria bloom waters. Limnol.Oceanogr. Methods 14, 432–447. doi: 10.1002/lom3.10102

CrossRef Full Text | Google Scholar

Welschmeyer, N. A. (1994). Fluorometric analysis of chlorophyll a in the presence of chlorophyll b and phaeopigments. Limnol. Oceanogr. 39, 1985–1992.

Google Scholar

Westberry, T., Behrenfeld, M. J., Siegel, D. A., and Boss, E. (2008). Carbon-based primary productivity modeling with vertically resolved photoacclimation. Glob. Biogeochem. Cycle 22:GB2024. doi: 10.1029/2007GB003078

CrossRef Full Text

Wright, S. W., Jeffrey, S. W., Mantoura, R. F. C., Llewellyn, C. A., Björnland, T., Repeta, D., et al. (1991). Improved HPLC method for the analysis of chlorophylls and carotenoids from marine phytoplankton. Mar. Ecol. Prog. Ser. 77, 183–196.

Google Scholar

Zonneveld, C., van den Berg, H. A., and Kooijaman, S. A. L. M. (1997). Modeling carbon cell quota in light-limited phytoplankton. J. Theor. Biol. 188, 215–222.

Google Scholar

Keywords: Phaeocystis, phytoplankton, photosynthesis, modeling, co-limitation, photophysiology, quantum yield, absorption

Citation: Moisan TA and Mitchell BG (2018) Modeling Net Growth of Phaeocystis antarctica Based on Physiological and Optical Responses to Light and Temperature Co-limitation. Front. Mar. Sci. 4:437. doi: 10.3389/fmars.2017.00437

Received: 06 July 2017; Accepted: 20 December 2017;
Published: 23 February 2018.

Edited by:

François G. Schmitt, Centre National de la Recherche Scientifique (CNRS), France

Reviewed by:

Fabrice Lizon, Lille University of Science and Technology, France
Raghab Ray, UMR6539 Laboratoire des Sciences de L'environnement Marin (LEMAR), France

Copyright © 2018 Moisan and Mitchell. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: B. Greg Mitchell, gmitchell@ucsd.edu

Deceased.

Download