+ All Categories
Home > Documents > Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning...

Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning...

Date post: 01-Nov-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
12
Phytoplankton succession explains size-partitioning of new production following upwelling-induced blooms N. Van Oostende a, , J.P. Dunne b , S.E. Fawcett a , B.B. Ward a a Department of Geosciences, Princeton University, USA b NOAA Geophysical Fluid Dynamics Laboratory, USA abstract article info Article history: Received 1 October 2014 Received in revised form 23 January 2015 Accepted 30 January 2015 Available online 7 February 2015 Keywords: Phytoplankton Cell size Community succession New production Coastal upwelling Nitrate uptake Biogeochemical model Large and chain-forming diatoms typically dominate the phytoplankton biomass after initiation of coastal upwelling. The ability of these diatoms to accelerate and maintain elevated nitrate uptake rates has been proposed to explain the dominance of diatoms over all other phytoplankton groups. Moreover, the observed delay in biomass accumulation following nitrate supply after initiation of upwelling events has been hypothesised to result from changes in the diatom community structure or from physiological acclimation. To in- vestigate these mechanisms, we used both numerical modelling and experimental incubations that reproduced the characteristic succession from small to large species in phytoplankton community composition and size structure. Using the Tracers Of Phytoplankton with Allometric Zooplankton (TOPAZ) ecosystem model as a framework, we nd that variations in functional group-specic traits must be taken into account, through adjust- ments of group-dependent maximum production rates (P Cmax ,s 1 ), in order to accurately reproduce the observed patterns and timescales of size-partitioned new production in a non-steady state environment. Repre- sentation of neither nutrient acclimation, nor diatom diversity in the model was necessary as long as lower than theoretical maximum production rates were implemented. We conclude that this physiological feature, P Cmax , is critical in representing the early, relatively higher specic nitrate uptake rate of large diatoms, and explains the differential success of small and large phytoplankton communities in response to nitrate supply during upwelling. © 2015 Elsevier B.V. All rights reserved. 1. Introduction A disproportionate fraction of global primary production (1030%) and carbon sequestration (4085%) relative to their ocean area (519%) occurs on continental shelves, especially in areas of coastal upwelling (Dunne et al., 2007; Longhurst et al., 1995; Muller-Karger et al., 2005). Wind-driven upwelling of nutrient rich water into the eu- photic zone along western continental margins drives the development of phytoplankton blooms, which are often dominated by large and chain-forming diatoms (Estrada and Blasco, 1985; Kudela et al., 2008). Because of their relatively large size and biomineral content, these bloom-forming diatoms efciently transfer newly produced biomass to higher trophic levels and to the deep ocean and seaoor, where car- bon sequestration occurs (Stock and Dunne, 2010; Thunell et al., 2007). Because of this interconnection between phytoplankton community composition and biogeochemical cycles, numerical models used to understand elemental cycles in the context of climate change can be improved by incorporating explicit representation of microbial commu- nity structure. Hydrodynamic turbulence, and hence the light and nutrient regime, exerts strong controls on the phytoplankton assemblage (Margalef, 1978). Some phytoplankton species, notably larger diatoms, are espe- cially adept at exploiting the higher nutrient conditions that character- ise upwelling events compared to other, often smaller phytoplankton species, whose growth rates saturate at lower nutrient and light levels (Barber and Hiscock, 2006; Finkel, 2001; Key et al., 2010; Litchman et al., 2007). Field and modelling studies have shown that as the total phytoplankton biomass increases, the biomass of larger phytoplankton increases relatively more than that of the smaller species. Across the resource concentration gradient, this means that disproportionally more biomass is added to the larger phytoplankton at higher resource concentrations (Goericke, 2011; Irigoien et al., 2004; Li, 2002; Poulin and Franks, 2010). Phytoplankton cell size is often used as a key functional characteris- tic. Size distribution of phytoplankton communities can be modelled without invoking grazing control via bottom-up control of phytoplank- ton growth through maximum growth rates that increase with cell size (Irwin et al., 2006). However, this approach contradicts observed allometric relationships, which nd decreasing maximum growth Journal of Marine Systems 148 (2015) 1425 Corresponding author at: Princeton University, Guyot Hall, Department of Geosciences, Princeton NJ 08544, USA. Tel.: +1 609 258 1052. E-mail address: [email protected] (N. Van Oostende). http://dx.doi.org/10.1016/j.jmarsys.2015.01.009 0924-7963/© 2015 Elsevier B.V. All rights reserved. Contents lists available at ScienceDirect Journal of Marine Systems journal homepage: www.elsevier.com/locate/jmarsys
Transcript
Page 1: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

Journal of Marine Systems 148 (2015) 14–25

Contents lists available at ScienceDirect

Journal of Marine Systems

j ourna l homepage: www.e lsev ie r .com/ locate / jmarsys

Phytoplankton succession explains size-partitioning of new productionfollowing upwelling-induced blooms

N. Van Oostende a,⁎, J.P. Dunne b, S.E. Fawcett a, B.B. Ward a

a Department of Geosciences, Princeton University, USAb NOAA Geophysical Fluid Dynamics Laboratory, USA

⁎ Corresponding author at: Princeton University,Geosciences, Princeton NJ 08544, USA. Tel.: +1 609 258 1

E-mail address: [email protected] (N. Van Oost

http://dx.doi.org/10.1016/j.jmarsys.2015.01.0090924-7963/© 2015 Elsevier B.V. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 1 October 2014Received in revised form 23 January 2015Accepted 30 January 2015Available online 7 February 2015

Keywords:PhytoplanktonCell sizeCommunity successionNew productionCoastal upwellingNitrate uptakeBiogeochemical model

Large and chain-forming diatoms typically dominate the phytoplankton biomass after initiation of coastalupwelling. The ability of these diatoms to accelerate and maintain elevated nitrate uptake rates has beenproposed to explain the dominance of diatoms over all other phytoplankton groups. Moreover, the observeddelay in biomass accumulation following nitrate supply after initiation of upwelling events has beenhypothesised to result from changes in the diatom community structure or from physiological acclimation. To in-vestigate these mechanisms, we used both numerical modelling and experimental incubations that reproducedthe characteristic succession from small to large species in phytoplankton community composition and sizestructure. Using the Tracers Of Phytoplankton with Allometric Zooplankton (TOPAZ) ecosystem model as aframework, wefind that variations in functional group-specific traitsmust be taken into account, through adjust-ments of group-dependent maximum production rates (PCmax, s

−1), in order to accurately reproduce theobserved patterns and timescales of size-partitioned new production in a non-steady state environment. Repre-sentation of neither nutrient acclimation, nor diatom diversity in the model was necessary as long as lower thantheoretical maximum production rates were implemented. We conclude that this physiological feature, PCmax, iscritical in representing the early, relatively higher specific nitrate uptake rate of large diatoms, and explains thedifferential success of small and large phytoplankton communities in response to nitrate supply duringupwelling.

© 2015 Elsevier B.V. All rights reserved.

1. Introduction

A disproportionate fraction of global primary production (10–30%)and carbon sequestration (40–85%) – relative to their ocean area(5–19%) – occurs on continental shelves, especially in areas of coastalupwelling (Dunne et al., 2007; Longhurst et al., 1995; Muller-Kargeret al., 2005). Wind-driven upwelling of nutrient rich water into the eu-photic zone along western continental margins drives the developmentof phytoplankton blooms, which are often dominated by large andchain-forming diatoms (Estrada and Blasco, 1985; Kudela et al., 2008).Because of their relatively large size and biomineral content, thesebloom-forming diatoms efficiently transfer newly produced biomassto higher trophic levels and to the deep ocean and seafloor, where car-bon sequestration occurs (Stock and Dunne, 2010; Thunell et al., 2007).Because of this interconnection between phytoplankton communitycomposition and biogeochemical cycles, numerical models used tounderstand elemental cycles in the context of climate change can be

Guyot Hall, Department of052.ende).

improved by incorporating explicit representation ofmicrobial commu-nity structure.

Hydrodynamic turbulence, and hence the light and nutrient regime,exerts strong controls on the phytoplankton assemblage (Margalef,1978). Some phytoplankton species, notably larger diatoms, are espe-cially adept at exploiting the higher nutrient conditions that character-ise upwelling events compared to other, often smaller phytoplanktonspecies, whose growth rates saturate at lower nutrient and light levels(Barber and Hiscock, 2006; Finkel, 2001; Key et al., 2010; Litchmanet al., 2007). Field and modelling studies have shown that as the totalphytoplankton biomass increases, the biomass of larger phytoplanktonincreases relatively more than that of the smaller species. Across theresource concentration gradient, this means that disproportionallymore biomass is added to the larger phytoplankton at higher resourceconcentrations (Goericke, 2011; Irigoien et al., 2004; Li, 2002; Poulinand Franks, 2010).

Phytoplankton cell size is often used as a key functional characteris-tic. Size distribution of phytoplankton communities can be modelledwithout invoking grazing control via bottom-up control of phytoplank-ton growth through maximum growth rates that increase with cell size(Irwin et al., 2006). However, this approach contradicts observedallometric relationships, which find decreasing maximum growth

Page 2: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

15N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

rates with increasing cell size (Edwards et al., 2012), but see Marañónet al. (2013). To resolve this contradiction, it has been proposed thatfurther phylogenetic constraints be included in models. For example,maximum potential growth rates can be assigned to broad taxonomicgroups to impose controls on phytoplankton community structure(e.g., diatoms vs. green algae cf. Litchman et al. (2010)).

The differentiation between growth traits of functional groups incurrent generation global models is still poorly constrained (Finkelet al., 2010). Coastal ecosystem models using at least two, oftensize-based, functional phytoplankton groups have been used to improveour understanding of patterns of primary production (Goebel et al.,2010; Klein, 2002; Li et al., 2010;Moloney et al., 1991). The dynamic na-ture of coastal ecosystems creates transient environmental conditionsto which individual phytoplankton cells rapidly adjust in terms ofresource allocation to nutrient uptake and cell growth (Morel, 1987).Some of the earlier coastal ecosystem models invoked physiologicalmechanisms, such as acclimation to changing light and/or nutrientconditions, or ecological succession to explain the apparent accelerationof nitrate uptake rate by the whole phytoplankton community follow-ing an upwelling event (Dugdale et al., 1990; Wilkerson and Dugdale,1987; Zimmerman et al., 1987), but see Garside (1991). In addition,many instances of acclimation of nitrate uptake or assimilation andthe uncoupling of nutrient uptake and cell growth have been docu-mented experimentally (Collos, 1986; Collos et al., 2005; Jochem et al.,2000; Smith et al., 1992). However, it has not been possible to quantita-tively link these ecophysiological processes to the observed timing andpartitioning of new production into different phytoplankton size andtaxonomic groups following an upwelling event.

This study had two main objectives: First, to further evaluate thephytoplankton community composition and succession in a previouslydescribed mesocosm experiment (Fawcett and Ward, 2011) and sec-ond, to determine which ecophysiological processes in phytoplanktoncommunities regulate the observed patterns in the timing and size-partitioning of new production following the initiation of an upwellingevent. The mesocosm experiment simulated growth conditions aftercoastal upwelling initiation and measured production and nutrientuptake into three size fractions (small (PS) = 0.7–5 μm; medium(PM) = 5–20 μm; large (PL) = N20 μm) during the subsequent devel-opment of a phytoplankton bloom. We hypothesised that, to explainthe observed patterns in new production and shifts in communitycomposition, i) the phytoplankton size groups had different intrinsicmaximum production rates, and that the timing of new production insuch a transient environment could be explained alternatively by ii)the physiologically-limited ability of phytoplankton groups to acclimateto the improved nutrient conditions, or iii) by allowing for greater reso-lution in taxonomic and subsequently physiological diversity. To testthose hypotheseswe calibrated the ecosystemmodel “Tracers Of Phyto-plankton and Allometric Zooplankton” (TOPAZ) (Dunne et al., 2005,2013) using a coastal upwelling species assemblage, and used it as aframework for evaluating the results of the mesocosm experiment.

2. Methods

2.1. Field experiment overview

Phytoplankton blooms were initiated by simulating coastal post-upwelling conditions in three 200 l mesocosms (hereafter, B1, B2, B3)as reported by Fawcett and Ward (2011). Briefly, water was collectedduring the upwelling season in central Monterey Bay, California(36.85°N, 121.97°W; bottom depth = 250 m) from 70 m depth,where the nitrate concentration ([NO3]) was sufficiently high(~20 μmol l−1) to initiate a bloom. The mesocosms were inoculatedwith 2 l of surface water and incubated for 8 days at the light and tem-perature conditions of surface water. The water in the mesocosms wasmixed three times each day tominimise particle settling. The initial am-monium (NH4) and NO3 concentrations in the mesocosms were typical

of seawater upwelled from 70 m depth in Monterey Bay. See Fawcettand Ward (2011) for nutrient concentration data and mass balancevalidation.

Ambient uptake rates of NH4 and NO3 by the PS, PM and PL size frac-tions were determined daily starting on the second day using isotopictracer (15N) incubations (3 h) of 1.5 l subsamples from the mesocosmsand size-fractionation of particulate nitrogen (PN) by filtration. PN and15N content were measured using an elemental analyser coupled to aEuropa Scientific 20/20 mass spectrometer, as described in Fawcettand Ward (2011) and in the supplementary methods section. The up-take rates (ρi, μmol l−1 h−1) were calculated according to the equationof Dugdale and Goering (1967). Specific rates of uptake of NH4 (VNH4)

and NO3 (VNO3) were calculated by normalising ρi to [PN] and averaging

over the length of the photoperiod during the experiment (13.9±0.1 h,based on continuous measurements of photosynthetic active radiation(PAR) performed by the Moss Landing Marine Laboratories WeatherStation; http://weathernew.mlml.calstate.edu).

The relative specific uptake rates of NO3 (Ni) in each PN size fraction(Pi) were calculated by normalising the specific nutrient uptake rate in aparticular size fraction (VNiPi) to the combined specific uptake in all sizefractions, according to:

RelVPiNi ¼ VPi

Ni=Xn

i¼1ρPiNi =Xn

i¼1PN½ �PiÞ:

�ð1Þ

The equivalent model-derived relative specific uptake rates werecalculated, using daily-averaged values, as:

RelVPiNi ¼ JuptakePiNi = PN½ �Pi=

Xni¼1

JuptakePiNi =Xn

i¼1PN½ �PiÞ :

�ð2Þ

Total and size-fractionated f-ratios were then calculated accordingto the formulation of Eppley and Peterson (1979):

f Pi ¼ VPiNO3

= VPiNH4

þ VPiNO3

� �: ð3Þ

2.2. Plankton community composition

Plankton cells were identified and counted in preserved samples (1%paraformaldehyde) using light microscopy (Garrison et al., 2005).Smaller cells (b5 μm) were classified as autotrophic or heterotrophicflagellates based on the presence or absence of a chloroplast;picocyanobacterial cells were not counted. At least 270 smaller cellsand 730 larger cells were counted each day by microscopy, yielding aminimum counting accuracy of 50% at the 95% confidence level for thespecies that becamemost abundant throughout the experiment. Largercells were identified to the species level, except for ciliates, which wereclassified as Strombidinium sp., Oligotrichous, Tintinnid or aloricate(Hasle and Syvertsen, 1997; Steidinger and Tangen, 1997; Throndsen,1997). Biovolumes were estimated using cellular geometrical shapesand average cellular dimensions (Olenina et al., 2006), and biovolumewas then converted to carbon using the relationships of Menden-Deuer and Lessard (2000).

Growth rates of the plankton species or groups thatweremost abun-dant at the start or end of the experiment were estimated using anexponential growth model where net specific growth rate (μ, day−1)equals the slope of the linear regression of the natural logarithm ofcell abundance with time (Wood et al., 2005). Differences betweenmesocosms in the net growth rates and initial summed abundance ofthe main small or large phytoplankton species or groups were deter-mined by analysis of covariance (ANCOVA) and post hoc Tukey's honestsignificant difference (HSD) test.

Page 3: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

Box 1Nitrogen sources and sinks.

The main nitrogen sources and sinks as depicted in the diagram inFig. 1 are represented by the equations below. See Dunne et al.(2013, Supplementary material section) for more details.

∂NPS

∂t ¼ JuptakePSNO3þ JuptakePSNH4

−JgrazPSN ð4Þ

∂NPM

∂t ¼ JuptakePMNO3þ JuptakePMNH4

−JgrazPMN ð5Þ

∂NPLi

∂t ¼ JuptakePLiNO3þ JuptakePLiNH4

ð6Þ

∂NZμ

∂t ¼ JgrazPSN þ JgrazPMN −JprodDetN −JprodDONiN −JlossZμN ð7Þ

∂NDet

∂t ¼ JprodDetN −JreminDetN ð8Þ

∂NDONi

∂t ¼ JprodDONiN −JreminDONiN ð9Þ

∂NH4

∂t ¼ JexcretZμNH4þ JlossZμN þ JreminDONi

N

−JuptakePSN −JuptakePMN −JuptakePLN

ð10Þ

∂NO3

∂t ¼ −JuptakePSN −JuptakePMN −JuptakePLN ð11Þ

16 N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

2.3. Overview of the model's structure and assumptions

The TOPAZ biogeochemical model (Dunne et al., 2005, 2013) wasused as a framework to reproduce and evaluate the nitrogen dynamicsin the experimental simulation of a coastal upwelling event as describedabove. Nitrogen is the biomass currency in the model, which uses 30state variables to resolve the cycling of nitrogen, carbon, phosphorus,silicate, iron, calcium carbonate and oxygen. The differential equationsencompassing the sources and sinks for each state variable are outlinedin Dunne et al. (2013, Supplementary material section). The planktonicfood web comprises 10 nitrogen state variables, whose interactions areillustrated in Fig. 1 and described in Box 1. In brief, light (Irrmem) regu-lates phytoplankton production (Geider et al., 1997), the rate of whichis determined by the most limiting nutrient (nitrogen, phosphorus, oriron (Fe)) sensu Liebig (NutLim) and the metabolic cost of biosynthesis(ζ) (Geider, 1992) (Box 2). Nitrogen and phosphorus limitation ofphytoplankton (Lim) are hyperbolic functions of their ambient concen-tration and the nutrient uptake half-saturation constants of a particularphytoplankton size group (Edwards et al., 2012).

Uptake of NO3 and NH4 by phytoplankton is calculated as a fractionof total nitrogen uptake (Box 2, E12 and E13). Ammonium and NO3 lim-itation are calculated following Frost and Franzen (1992), where NO3

uptake is inhibited when NH4 concentrations are high relative to thehalf-saturation constant for NH4. By contrast, iron limitation (defFe) de-pends on the internal cellular iron quota (Sunda and Huntsman, 1995).Phosphate uptake and limitation (defPO4

) are patterned after the optimalallocation theory of Klausmeier et al. (2004) inwhichN:P is determinedby the allocation of resources to photosynthetic machinery, nutrientuptake, assembly, and other biomass constituents. The P:N ratio thatbecomes limiting for phytoplankton growth is taken from the fractionof the cell dedicated to assembly after Klausmeier et al. (2004). SeeDunne et al. (2013, Supplementary material section) for further details.

The growth rate of phytoplankton cells is a function of themaximumcarbon-specific rate of photosynthesis (PCm) at a given temperature andlevel of nutrient and light limitation, and the cost of biosynthesis (Box 2,E16, E19 and E20). PCm, in turn, is set by the maximum carbon-specificrate of photosynthesis at a reference temperature (0 °C) (PCmax) andis a function of temperature (i.e., PCm = PCmax ⋅ ekT, with k =0.0631 °C−1 and T in °C) according to Eppley (1972). The PCmax valuesin the initial model were set to yield a theoretical maximum growthrate (0.94 ⋅ 10−5s−1) following Bissinger et al. (2008).

Several assumptions were made to compare the modelled data withthe measurements from the mesocosm study: 1) mesozooplanktongrazers were ignored in the model because the water was collected by

Fig. 1. Summary of nitrogen (N) state variables and their interactions in the modifiedTOPAZ model (Table 1, configuration d). P: phytoplankton; Zμ: microzooplankton; Det:detritus; S: small; M: medium; L: large, r: fast-growing, K: slow-growing; NO3: nitrate;NH4: ammonium; DON: dissolved organic N, l: labile, sl: semi-labile.

Niskin bottles, which most likely excluded large grazers (feeding onlarge or chain-forming diatoms) but included protozoan grazers(feeding on smaller phytoplankton); 2) the mesocosm water waswell-mixed (because it was stirred); 3) iron was non-limiting sincethe mesocosm water was collected near-shore and the experimentwas not performed in an iron-clean environment; 4) fluxes of nitrogenfixation and denitrification were not significant during the incubations,because inorganic nitrogen concentrations were almost always high, anitrogen mass balance between nutrient loss and biomass gain wasobserved, thewaterwaswell-oxygenated, and no sediment layer devel-oped. Because of these assumptions, and for better comparisonwith theexperimental results of Fawcett and Ward (2011), the diazotrophs inthe original TOPAZ formulation were replaced by an additional large(N20 μm) phytoplankton size group (PL). This size group encompassedmainly large diatoms such as Thalassiosira spp. and smaller chain-forming diatoms such as Chaetoceros spp.. Thus, in analogy to themesocosm experiment 3 phytoplankton size groups were consideredin the model. The two smaller phytoplankton size groups (PS and PM)were subject to loss through grazing by microzooplankton (Zμ) bymeans of a Type II Holling functional response (Holling, 1959). Theclearance rate (λ, [mol N m−3 s]−1) was estimated based upon agross growth efficiency (gge) of 0.3 (Hansen et al., 1997; Straile,1997), microscopy measurements of microzooplankton growth rate(μZμ) (see above), and small and medium phytoplankton biomass (NPS

and NPM): λ = μZμ / gge / (NPS + NPM). Microzooplankton loss wasgoverned by quadratic loss representing cannibalism (Steele andHenderson, 1992).

Page 4: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

Box 2Nitrogen uptake, nutrient and light limitation, growth, grazing, produc-tion of DON and detritus, and remineralisation.

Processes controlling the main nitrogen sources and sinks asdepicted in the diagram in Fig. 1 are represented by the equationsbelow. See Dunne et al. (2013, Supplementary material section)for an overview of the parameter values used in the TOPAZmodel.

JuptakePiNO3¼ μPi � NPi � LimPi

NO3

LimPiNO3

þ LimPiNH4

ð12Þ

JuptakePiNH4¼ μPi � NPi � LimPi

NH4

LimPiNO3

þ LimPiNH4

ð13Þ

NutLimPi ¼ min LimPiNO3

þ LimPiNH4

; min def PiPO4;def PiFe

� �� �ð14Þ

NutLimPimem tþ 1ð Þ ¼ NutLimPi

mem tð Þ

þ NutLimPi tð Þ−NutLimmem tð Þ� �

� γNutmem

ð15Þ

PPiCm ¼ PPi

Cmax � min NutLimPi;NutLimPi

mem

� �� ekT ð16Þ

PPiCmθ ¼ PPi

Cmax � NutLimPimem � ekT ð17Þ

θPi ¼ θPimax−θPimin

� �= 1þ θPimax−θPimin

� �� αPi � Irrmem:

0:5PPiCmθ

!

þ min NutLimPi;NutLimPi

mem

� �� θPiNutLim þ θPimin

ð18Þ

IrrLimPi ¼ 1−e−αPi �Irr�θPi

PPiCm ð19Þ

μPi ¼ PPiCm

1þ ζð Þ � IrrLimPi ð20Þ

JgrazPiN ¼ min1Δt

;λ � ekT� �

� NPi

NPi þ NPimin

!� NPi � NZμ ð21Þ

JlossZμN ¼ γNZμ � ekT � NZμ � NZμ ð22Þ

JprodDetN ¼ f PSDet � JgrazPSN � 1−ϕDONl−ϕDONslð Þ � ekreminT ð23Þ

JprodDONslN ¼ ϕDONsl � JgrazPSN ð24Þ

JprodDONlN ¼ ϕDONl � JgrazPSN � LimZμP ð25Þ

JreminDetN ¼ γNDet � NDet ð26Þ

JreminDONiN ¼ γNDONi � NDONi ð27Þ

17N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

2.4. Model initialization and forcing

The model was initialised using the initial concentrations measuredin eachmesocosm for NO3 andNH4, PN in each size group, and dissolvedorganic nitrogen (DON). The initial concentration of other nutrientssuch as phosphate and silicate were based on local climatologies(Garcia et al., 2010). The initial dissolved [Fe] was set in excess as16.45 nmol l−1, which corresponds to 5 times the upwelling valuemeasured by Johnson et al. (2001), to account for the non iron-cleanenvironment of the experimental mesocosms. Because of the very lowsurface [NO3] at the sampling site prior to the upwelling experiment(K. Johnson, personal communication), we set the initial values of thenutrient limitation memory (NutLimmem) to be low, corresponding toa Fe:N molar ratio of 10−2, which is equivalent to ambient concentra-tions of dissolved Fe of 1 nmol l−1 and NO3 of 100 nmol l−1. Themodel was forced using the water temperature measured in themesocosms and the time-series of local photosynthetic active radiation(PAR) measurements fromMoss Landing Marine Laboratories WeatherStation, scaled to measured PAR intensity inside the mesocosms.

Allocation of initial PN into non-living detritus, heterotrophic andautotrophic biomasswas constrained bymicroscopy estimates of plank-ton biomass. We estimated the initial detritus content of the PN byassuming that the exponential rate of biomass increase derived fromthe microscopy data (Nmicroscopy) should equal that of PN minus aninitial amount of non-proliferating detrital PN (PNdetritus), according to:

dNmicroscopy

dt¼ Nmicroscopy � et ð28Þ

dPNbiomass

dt¼ PNbiomass � et ð29Þ

PN tð Þ ¼ PNbiomass � et þ PNdetritus: ð30Þ

Detritus initially constituted 38%, 68%, and 39% of the largest PN sizefraction, and 54%, 71%, and 64% of the summed PNb5 + PN5–20 sizefractions in B1, B2, and B3, respectively.

In accordance with the community succession patterns thatemerged from themicroscopy data (Fig. 2), we further divided the larg-est phytoplankton subgroup (PL) into an early-dominant, low growthrate group (PLK), and a later-dominant, high growth rate group (PLr).Based on the microscopy data, members of PL were operationallydefined as large autotrophic cells (equivalent spherical diameter(ESD) N20 μm or known to bear spines and/or to be colony-forming)contributing at least 5% to total plankton biomass on at least onesampling time point during the experiment. An exponential growthrate cut-off of 1.0 day−1 was then used to allocate the PL species tothe PLr and PLK subpopulations. To constrain the differential growthcharacteristics of these two PL subpopulations in the model (3 PCmax &2 PL groups), we used the ratio of the exponential growth rates of thebiomass of PLr species to the biomass of PLK species derived from themicroscopy counts. It was thus only necessary to optimise for PCmax inone of the two PL subpopulations.

2.5. Model configurations for hypothesis testing

To test our hypotheses about the ecophysiological mechanismsgoverning the timing and partitioning of new production into differentphytoplankton size groups, we adapted alternative configurations of themodel. In addition to adding the largest phytoplankton size group to thebaseline TOPAZ model, the following parameters and tracers weremodified in the different model configurations (Table 1): the number

Page 5: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

102

103

104

105

106

cell

abun

danc

e (c

ells

l1 )

Chaetoceros debilisChaetoceros curvisetusThalassiosira anguste lineataEucampia zodiacusPseudonitzschia sp.

0 1 2 3 4 5 6 7 810

3

104

105

106

days

cell

abun

danc

e (c

ells

l1 )

small pennate diatomauto. Gymnodinoidauto. flagellatehetero. flagellate

A

B

Fig. 2. Abundance (cells l−1) of large (A) and small (B) plankton species or groups that weremost numerous at the start or end of the field experiment inmesocosmB3 (see Fig. S1 for theresults of B1 and B2). auto.: autotrophic, hetero.: heterotrophic. Note the difference in the y-axis scale between large and small groups at the start of the incubation.

18 N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

of different optimised PCmax values (a. 1 PCmax or b. 3 PCmax), the rate ofnutrient acclimation (γNutmem) (c. 3 PCmax & nutrient acclimation) andthe number of PL groups (from one to two; d. 3 PCmax & 2 PL groups).The choice of parameters (PCmax, γNutmem) to be tested was based ontheir influence on NO3 consumption from results of sensitivity analyses(see Supplementary methods). Nutrient acclimation (Box 2, E15) wasintroduced into the model by including a simple temporal dependenceof the maximum photosynthetic carbon fixation rate (PCm) on themulti-nutrient limitation status of the cell (NutLimmem).

The ability of the different model configurations to reproduce themesocosm data was used as ameasure of theirmechanistic explanatorypotential. We employed an optimisation routine to calibrate the phyto-plankton group parameter values of PCmax from the original TOPAZmodel and γNutmem in the configuration that included nutrient acclima-tion. In effect, the model was fitted to measured values of [NH4], [NO3],[PN] and chlorophyll a concentration [Chla] for each size fraction,

Table 1Normalised cumulative error (ΣNRMSE) and optimised parameter values of PCmax and γNutmem

95% confidence interval of bootstrapping results (N=500) ofmesocosmB3.ΣNRMSE values repPN concentrations, and f-ratios within 1 SD (see Table S1 for ΣNRMSE and optimised paramet

Model configuration ΣNRMSE 105 PCmax at 0

PS (b5 μm)

a. 1 PCmax 3.49 ± 0.47 0.45 ± 0.04b. 3 PCmax 2.75 ± 0.23 0.39 ± 0.07c. 3 PCmax & nutrient acclimation 2.66 ± 0.23 0.49 ± 0.06d. 3 PCmax & 2 PL groups 2.77 ± 0.29 0.39 ± 0.07

γNutmem (day−

PS (b5 μm)

c. 3 PCmax & nutrient acclimation 2.66 ± 0.23 1.5 ± 0.6

[DON], and the f-ratios taking measurement error into account. The f-ratio values for the first 3 days only were used in the optimisationsince uptake rates of ammonium were prone to overestimation oncethe ambient NH4 pool was depleted (Fawcett and Ward, 2011). Forthe optimisation of γNutmem (see below), the minimum rate was set to0.5 day−1 in accordance with the findings of Sakshaug and Holm-Hansen (1977) regarding the time lag of PN accumulation bySkeletonema costatum in response to a phosphate pulse. The maximumacclimation rate was set to 24.0 day−1, which is essentially instanta-neous and based on the time necessary for bacterial DNA synthesisand chromosome replication induced by a nutritional shift-up(Ingraham et al., 1983).

The degree of mismatch between the model and the measurementsof the above-mentioned variables was represented by the root meansquared error normalised to 1 standard deviation of the measurements(NRMSE; table S1), where a NRMSE value of 1 or less represents a

for each phytoplankton size group (±1 SD) in the differentmodel configurations from theresent the relative deviation ofmodelled values frommeasurements of nutrients, Chla ander values of PCmax and γNutmem from single optimizations for each mesocosm).

°C (s−1)

PM (5–20 μm) PL(r) (N20 μm) PLK (N20 μm)

0.45 ± 0.04 0.45 ± 0.04 –

0.46 ± 0.11 0.54 ± 0.08 –

0.53 ± 0.10 0.67 ± 0.06 –

0.46 ± 0.11 0.83 ± 0.08 0.35 ± 0.03

1)

PM (5–20 μm) PL (N20 μm)

1.8 ± 0.8 1.0 ± 0.6 –

Page 6: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

102

101

100

101

102

PN

(µm

ol N

l1 )

total >20 µm 5 20 µm 0.7 5 µm

0

5

10

15

20

25

NO

3 (µm

ol l

1 )0

0.2

0.4

0.6

0.8

1

frac

tion

of to

tal P

N

>20 µm 5 20 µm 0.7 5 µm

0

0.2

0.4

0.6

0.8

1

NH

4 (µm

ol l

1 )

0 1 2 3 4 5 6 7 80

2

4

6

8

10

12

14

Chl

a (µ

g l

1 )

days

total >20 µm 5 20 µm 0.7 5 µm

0 1 2 3 4 5 6 7 80

0.2

0.4

0.6

0.8

1

days

fra

tio

A B

C D

E F

Fig. 3. Comparison of modelled nitrogen dynamics and mesocosm measurements using the same theoretical PCmax value (0.94 ⋅ 10−5s−1 at 0 °C) for each phytoplankton size group(original TOPAZ configuration; ΣNRMSE = 6.36). Measurements from mesocosm B3 are denoted by dots with error bars representing 1 SD, while lines represent modelled values.(A) [PN] in each size fraction, (B) [NO3], (C) proportion of total PN in each size fraction, (D) [NH4], (E) [Chla] in each size fraction, (F) f-ratio based onNO3 andNH4 uptake by the total PN pool.

19N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

perfectmatchwithinmeasurement error. A nonlinear optimisation rou-tinewas used to search the parameter space, using a bounded version ofMATLAB's Nelder–Mead simplex direct search algorithm fminsearch(Lagarias et al., 1998), by minimising the sum of NRMSE for each vari-able. Because small errors in the initial model values propagate in thismesocosm experimental design, and to allow for robust comparisonbetween the results from different model configurations we repeatedthe optimisation routines 500 times using a bootstrapping procedurewhere themeasured values of nutrients, particulate and dissolved nitro-gen, Chla concentrations and estimates of initial detritus content wererandomly perturbed around their mean using a normal distribution oftheir uncertainties.

3. Results

3.1. Species and size structure succession in the phytoplankton communityduring experimental upwelling-induced bloom development

The phytoplankton community was initially composed mainly ofsmall autotrophic and heterotrophic flagellates (b5 μm), Gymnodinoiddinoflagellates (b20 μm), small pennate diatoms (b20 μm), along with

larger cells such as the diatoms species Eucampia zodiacus andPseudonitzschia sp., and was very similar among mesocosms (Fig. 2and Fig. S1). However, there were inter-mesocosm differences in theinitial relative abundance of certain species, including: higher abun-dance of autotrophic Gymnodinoid dinoflagellates in B1 and B2, hetero-trophic flagellates in B2 and B3, Pseudonitzschia sp. in B1 and B3, thedinoflagellate Cochlodinium sp. in B1 and B3, and the diatom genusDactyliosolen in B1 (data not shown). Microzooplankton comprisedmainly heterotrophic flagellates and ciliates (e.g., Strombidinium sp.),both of which were more abundant in B2 than in B1 and B3 at the endof the incubation (data not shown). The starting [PN] was initiallyvery low in all mesocosms (1.00 ± 0.07 μmol N l−1; Figs. 3 to 5,Figs. S5 and S6), and increased considerably more in B3 (18.4 fold)than in B1 and B2 (12.8 and 13.4 fold) during the experiment. Most ofthe accumulated PN derived from NO3 assimilation, since the initial[NH4] was low and rapidly consumed, [NO2] was negligible, [DON]remained roughly constant at ~3 μmol l−1, and a nitrogenmass balancewas maintained in all mesocosms (Fawcett and Ward, 2011).

Community succession was most obvious among diatoms whereChaetoceros spp. and Thalassiosira spp. (N20 μm, in B3) came to domi-nate the assemblage (Fig. 2). An increase in the abundance of large

Page 7: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

102

101

100

101

102

PN

(µm

ol N

l1 )

total >20 µm 5 20 µm 0.7 5 µm

0

5

10

15

20

25

NO

3 (µm

ol l

1 )

0

0.2

0.4

0.6

0.8

1

frac

tion

of to

tal P

N

>20 µm 5 20 µm 0.7 5 µm

0

0.2

0.4

0.6

0.8

1

NH

4 (µm

ol l

1 )

0 1 2 3 4 5 6 7 80

2

4

6

8

10

12

14

Chl

a (µ

g l

1 )

days

total >20 µm 5 20 µm 0.7 5 µm

0 1 2 3 4 5 6 7 80

0.2

0.4

0.6

0.8

1

days

fra

tio

A B

C D

E F

Fig. 4. Comparison of modelled nitrogen dynamics and mesocosmmeasurements using a single optimised PCmax value for all of the 3 phytoplankton size groups (Tables 1 and S1, config-uration a). Measurements frommesocosmB3 are denoted by symbolswith error bars representing 1 standard deviation, while lines represent modelled values. (A) [PN] in each size frac-tion, (B) [NO3], (C) proportion of total PN in each size fraction, (D) [NH4], (E) [Chla] in each size fraction, (F) f-ratio based on NO3 and NH4 uptake by the total PN pool.

20 N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

cells (N20 μm) was apparent after the first day of incubation fore.g., Chaetoceros curvisetus, while an early increase in cell abundanceof small flagellates was not observed. It is noteworthy that no changesin NH4 and NO3 concentrations were detected during at least the first2 days in all mesocosms, probably because of the low initial cell abun-dance. Although [PN] increased in each size fraction over the course ofthe experiment, the highest increase in [PN] was invariably observedin the largest size fraction (31.0, 37.4, 48.9 fold in B1, B2, and B3). As aconsequence, a gradual shift in proportional contribution to PN fromthe smallest size fraction to the largest size fraction occurred in allmesocosms, and was most pronounced in B3 (Fig. 5). The difference intiming of bloom development between mesocosms was likely due to ahigher initial abundance of the main large phytoplankton species inB3 (ANCOVA and Tukey's HSD test: F = 6.51, p = 0.011 and p b 0.05).In subsequent analyses we will mainly focus on B3 because of its morecomplete succession sequence and the lack of significant differences be-tween the growth rates of the dominant small and large phytoplanktonspecies in the different mesocosms (ANCOVA: large: F = 0.33, p =0.725; small: F = 0.1, p = 0.908).

3.2. Mechanisms governing the timing and partitioning of new productioninto different phytoplankton size groups

The original model (Dunne et al., 2013), which used the theoreticalmaximum PCmax value of 0.94 ⋅ 10−5s−1 sensu Bissinger et al. (2008),strongly mischaracterised the observed rates of phytoplankton biomassaccumulation, nutrient consumption and community structure ob-served in the mesocosms (Fig. 3). Optimising the value of a singlePCmax for all phytoplankton size groups resulted in a downward revisionof the theoretical maximum PCmax value by 50%, yielding an improvedfit between the modelled and measured values of biomass accumula-tion and nutrient consumption (Table 1, configuration a). Althoughthe overall rate of NO3 andNH4 consumption and total PN accumulationcould be reasonably well represented when all phytoplankton groupshad the same PCmax value, community succession could not (Fig. 4 andTable 1, Tables S1 and S2). The shift in community size structure andthe associated transfer of NO3 into each phytoplankton size group weresignificantly better explained by the model output when optimisedphytoplankton group-specific values of PCmax were implemented

Page 8: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

102

101

100

101

102

PN

(µm

ol N

l1 )

total >20 µm 5 20 µm 0.7 5 µm

0

5

10

15

20

25

NO

3 (µm

ol l

1 )

0

0.2

0.4

0.6

0.8

1

frac

tion

of to

tal P

N

>20 µm 5 20 µm 0.7 5 µm

0

0.2

0.4

0.6

0.8

1

NH

4 (µm

ol l

1 )

0 1 2 3 4 5 6 7 80

2

4

6

8

10

12

14

Chl

a (µ

g l

1 )

days

total >20 µm 5 20 µm 0.7 5 µm

0 1 2 3 4 5 6 7 80

0.2

0.4

0.6

0.8

1

days

fra

tio

A B

C D

E F

Fig. 5. Comparison ofmodelled nitrogen dynamics andmesocosmmeasurements using optimised PCmax values for each of the 3 phytoplankton size groups (Tables 1 and S1, configurationb). Measurements from mesocosm B3 are denoted by symbols with error bars representing 1 standard deviation, while lines represent modelled values. (A) [PN] in each size fraction,(B) [NO3], (C) proportion of total PN in each size fraction, (D) [NH4], (E) [Chla] in each size fraction, (F) f-ratio based on uptake NO3 and NH4 uptake by the total PN pool.

21N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

(Fig. 5 and Figs. S3 and S4; ΣNRMSE in Table 1, Table S1 and Table S2,configuration b). Different phytoplankton groups had significantlydifferent PCmax values, with the PL group having the highest PCmax valuecompared to the other groups and to the model configuration using asingle uniform PCmax value (i.e., configuration a; comparison of thebootstrapping results of B3 by ANOVA and Tukey's HSD test: F =308.53, p b 0.001 and p b 0.001).

Table 2Measured and modelled maximum VNO3 for each particulate matter size fraction or phy-toplankton size group from B3, averaged over the length of the photoperiod. Modelledvalues are derived from the model configuration using optimised, group-specific PCmax

values (Table S1, configuration b).

PN size fractionor phytoplanktongroup

Size cut-offor ESD (μm)

Measuredmax. VNO3

(h−1)

Modelledmax. VNO3

(h−1)

Time atmax. VNO3

(d)

small, PS b5 0.11 ± 0.01 0.08 5medium, PM 5–20 0.11 ± 0.04 0.09 5large, PL N20 0.11 ± 0.10 0.09 4

To explain the observed delay in PN accumulation (Fig. 4) and lack ofgrowth in the initially abundant population of small cells (cell abun-dance of autotrophic flagellates increased only after day 3; Fig. 2) duringthe first two to three days of themesocosmexperiment, we allowed thepossibility for thephytoplankton groups to acclimate to improvednutri-ent conditions following upwelling simulation (configuration c). Ineffect, this introduced a damped growth response with a timescale de-pendent on the optimised parameter γNutmem. The inclusion of nutrientacclimation improved the overall fit of themodel with the observations(ΣNRMSE) compared to the model configuration that had phytoplank-ton group-specific PCmax values only (Table 1 and S1, and Fig. S2;ANOVA and Tukey's HSD test: F = 430.89, p b 0.001 and p = 0.015).Also, the [PN] in the smaller plankton size groups was better matchedto the observations when nutrient acclimation was included(Table S2). These improvements were minor, however, because of therelated effect of PCmax and γNutmem on the rate of nitrate consumption,and were not always apparent in mesocosms B1 and B2 (Table S1 andTable S2).

To address the discrepancy between the immediate growth of cer-tain large phytoplankton species as observed by microscopy (Fig. 2)

Page 9: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

0

0.5

1

1.5

2

2.5

3

mea

sure

d re

lativ

e V

NO

3

>20 µm 5 20 µm 0.7 5 µm

0

0.5

1

1.5

2

mod

elle

d re

lativ

e V

NO

3

PL

PM

PS

0 1 2 3 4 5 6 7 80

0.5

1

1.5

2

mod

elle

d re

lativ

e V

NO

3

days

1 PCmax

3 PCmax

A

B

C

Fig. 6. Relative specific NO3 uptake rates, (A) measured and corrected for initial detrituscontent in each PN size fraction in B3 and modelled for each phytoplankton group inmodel configurations (B) using a single optimised PCmax value or (C) group-specificoptimised PCmax values. A relative VNO3 value of 1 (dashed grey line) indicates that therelative specific uptake rate of a particular group is equal to that of the total assemblage.

22 N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

and the apparent need for a growth lag period to match the observed[PN] changes in B3 (Fig. 4), we divided the largest phytoplankton sizegroup into two subgroups representing initially abundant, slower grow-ing species (PLK; e.g., E. zodiacus) and initially rare, faster growingspecies (PLr; e.g., Chaetoceros spp. and Thalassiosira spp.). Adding thisbiological diversity to the model (configuration d) did not improve theoverall fit to the observations compared to the model configurationwith only phytoplankton group-specific PCmax values (i.e., configurationb; Fig. S3 and ΣNRMSE in Table 1; ANOVA and Tukey's HSD test: F =430.89, p b 0.001 and p N 0.05). This model configuration, however,was able to better explain the accumulation of PN in the largest sizefraction (Table S2) and demonstrated that cells of the initially rare PLr

population could indeed achieve near theoretical maximum growthrates (Table 1 and Table S1).

3.3. Specific nitrate uptake rates and community structure shift

Themeasured andmodelledmaximumNO3uptake rates for each PNsize fraction during the development of the bloom in B3 are shown inTable 2 as a specific uptake rate (i.e., nutrient transport rate normalisedto [PN] for each size fraction or phytoplankton group and averaged over

the daily photoperiod). Focusing on B3, and using the most parsimoni-ous model configuration that yielded an improved fit (3 PCmax), maxi-mum VNO3

values for the modelled bloom were of similar magnitudefor all size groups, but generally lower than the VNO3

values derivedfrom the isotopic tracer incubations (Table 2). It is noteworthy thatthemaximumVNO3

of PL was reached one day earlier than in the smallergroups, PS and PM.

Phytoplankton groups with a higher than average relative specific

nitrate uptake rate (RelVPiNO3

N 1) will have a competitive advantage.The measured and modelled RelVNO3

for each PN size fraction and phy-toplankton group, respectively, reflects the nitrate uptake velocity of aparticular size group normalised to the nitrate uptake velocity of the en-tire phytoplankton assemblage. The measured RelVNO3

for each PN sizefraction (Fig. 6A), corrected for the initial detritus content, did notshow an obvious temporal trend, nor was there a clear difference inRelVNO3

between the different PN size fractions, except for the highRelVNO3

of PL after the second day. In contrast to the measured uptakerates, the modelled specific nutrient uptake rates are internally consis-tent (i.e., values at a given time point are dependent on previous valuessuch that mass balance is achieved by definition). Moreover, themodelled values are not confounded by inclusion of initial detritusinto the biomass pool or violation of the tracer level assumptions duringexperimental incubations (Lipschultz, 2008), as likely happened for theNH4 uptake ratemeasurements beyond day 3. The RelVNO3

of PL was ini-tially higher than that of the smaller groups in themodel configurationswith a single optimised PCmax value (configuration a) or group-specificoptimised PCmax values (configuration b) (Fig. 6B and C), and evenmore pronounced in the model configuration including nutrient accli-mation or resolving higher PL diversity (Fig. S6). This early potentialdominance is caused in part by the nutrient uptake characteristics andthe potential maximum production rate of each phytoplankton group.The effect of the latter is reflected in the extension of the period of

high RelVPLNO3

when the PL group is attributed a higher PCmax in themodel configurations using optimised group-specific PCmax values(Fig. 6C). Even though the RelVNO3

PL decreased to the level of the totalassemblage during bloom development and concomitant nutrientexhaustion (Fig. 6 and Fig. S6), the PL group sustained its competitiveadvantage longer than the smaller phytoplankton size groups whilenitrate was plentiful.

4. Discussion

The mesocosm experiment (Fawcett and Ward, 2011) reproducedthe conditions of a coastal upwelling event and the ensuing phytoplank-ton bloom, as evidenced by the timescale of nutrient drawdown and thesuccession in phytoplankton community size structure as a whole andthe diatom assemblage in particular (Dugdale and Wilkerson, 1989;Dugdale et al., 2006;Wilkerson et al., 2006). Phytoplankton species dom-inating the community at the end of the experiment (e.g., Chaetocerosdebilis, C. curvisetus and Thalassiosira anguste-lineata) were typical ofthose that dominate the phytoplankton bloom assemblage followingcoastal upwelling (Lassiter et al., 2006; Venrick, 2009). To simulatethese characteristic patterns of nutrient drawdown and partitioning ofnitrate uptake into different phytoplankton size groups, it was neces-sary to differentiate phytoplankton group-specific maximum produc-tion rates, through optimisation of PCmax values. The inclusion ofneither nutrient acclimation (through γNutmem), nor resolution ofhigher diversity within the largest model size group, representingdiatoms, further improved the overall model fit significantly beyondutilisation of appropriate group-specific maximum production ratevalues. While this implies that inclusion of additional diversity withinfunctionally homogeneous groups was not necessary to explain therates of new production during coastal upwelling conditions, we alsodemonstrated that an initially very small fraction of the PL populationcould indeed achieve near theoretical maximum growth rates (Table 1

Page 10: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

23N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

and Table S1). Furthermore, acclimation of phytoplankton cells' physio-logical condition to higher nutrient conditions, although experimentallydocumented, was not essential to simulate the evolution of size-partitioned new production during the experiment. This relativelyparsimonious mechanistic physiological functioning and resultingbiogeochemical signature provides a tantalizingly simple communitylevel description. In the next sections we reconcile the seeming contra-diction between the relatively simple biogeochemical outcome and theunderlying complex competitive interactions and resulting composi-tional changes at the group level observed through biodiversityindicators.

4.1. Group-specific maximum production rates

All phytoplankton cells generally respond to a relative improvementin nutrient and light conditionswith an increase in growth rate, and thesaturation level of this growth response is specific to cell size andphylogenetic group (Barber and Hiscock, 2006; Litchman et al., 2010;Marañón et al., 2013). Such a size-specific growth response partlyexplains the additive nature of the phytoplankton size distribution asa function of resource concentration (Poulin and Franks, 2010) andthe accumulation of large to intermediate-sized diatoms during upwell-ing conditions. Thus, the use of multiple parameter values to describegrowth characteristics (e.g., PCmax and nutrient half-saturation coeffi-cients) of phytoplankton cells in ecosystem models is fundamental forthe accurate representation of different phytoplankton functionalgroups, which often differ in size (Follows et al., 2007).

In this study, the original parameterisation of the TOPAZ ecosystemmodel, which used a universal parameter value for maximal productionrate based on maximum growth rate as a function of temperature(Bissinger et al., 2008), could not reproduce the rate and size-partitioning of new production from experimental observations(Fawcett and Ward, 2011) (Fig. 3). Instead, lower, size group-specificmaximal production rates were necessary to achieve correspondencebetween the modelled output and observations. Compared to openocean systems to which the TOPAZ model is currently being applied,the lower PCmax values generated by this study likely result from thesub-optimal conditions that characterise transient environments suchas coastal upwelling systems, or from physiological characteristics ofthe species adapted to these environments. Within each mesocosm,the optimum PCmax value of the smaller phytoplankton size groupswas lower than that of the largest size group (Table 1). This was thecase even though grazing pressure, which would act to elevate PCmax

during the optimisation process, had been taken into account. Suchlow PCmax values for smaller size groups such as flagellates and smallGymnodinoids are consistent with data compiled by Geider et al.(1997). Diatoms typically dominated the largest size fraction in themesocosm experiment and generally have high PCmax and μmax in natu-ral environments or cultures (Furnas, 1990; Geider et al., 1997). Theoptimised PCmax value of the large, fast growing phytoplankton group(PLr) was more similar to the predicted values (Bissinger et al., 2008;Marañón et al., 2013) than that of the other size groups since it wasrooted in the growth rates of fast growing species determined bymicroscopy (Table 1 and Fig. 2). One example of a large, fast growingspecies is the chain-forming diatom C. debilis. Even though it has amodest cell size by diatom standards (ESD = ~13 μm) it was retainedin the N20 μm size fraction during the size filtration procedure in themesocosm experiment. Its cell size places C. debilis in the top part ofthe unimodal distribution of growth rate versus cell size in Fig. 1 ofMarañón et al. (2013), which, together with its ability to avoidmicrozooplankton grazing by bearing setae and forming cell chains,renders C. debilis a successful bloom species in a coastal upwellingenvironment.

Considering the growth of the community in the largest size fractionas awhole, the initial delay in PN accumulation in this fraction in B3wasbetter reproduced by themodel when the initial growth rate of this size

group was reduced, either by incorporating a slower community-levelnutrient acclimation timescale or by dividing the large-sized communityinto initially abundant, slower growing species (e.g., E. zodiacus andPseudonitzschia sp.) and initially rare, faster growing species(e.g., C. debilis and T. anguste-lineata). The apparent difference in growthstrategy of the two bloom-forming diatom species, T. anguste-lineataand the smaller C. debilis, is noteworthy. While T. anguste-lineata wasnot detected during the first 4 days of the experiment, after which itgrew exponentially, C. debilis was growing exponentially by the secondday (Fig. 2). This is consistent with the findings of Collos (1982, 1986),and suggests a decoupling of nitrate uptake and growth by the generallylarger T. anguste-lineata cells compared to the reduced capacity for ni-trate storage by C. debilis. Both growth strategies were successful inthis particular context.

4.2. Nutrient acclimation

The temporal lag between the metabolic stimulus and thebiogeochemically-relevant responses, such as nutrient uptake rate andcell growth, can be affected by both the prior physiological status ofthe organism and the magnitude of the environmental shift (Collos,1986; Collos et al., 2005; Smith et al., 1992). Nutrient-limited cellsexposed to a sudden increase in ambient nutrient concentration oftendisplay a time lag in growth or physiological acclimation to the newcondition (Collos, 1986; Collos et al., 2005). Given the lack of upwellingin the two weeks preceding the experiment (www.pfeg.noaa.gov)and the extremely low surface [NO3] at the time of the experiment(K. Johnson, personal communication), the phytoplankton cellsfrom the surface that were used to inoculate the nutrient-rich 70 mdepthwaterwere likely nutrient stressed. The growth lag of the initiallyabundant small flagellates and the lack of a detectable change in [PN]over the first three days of the experiment partially support this hy-pothesis. We acknowledge, however, that suchmesocosm experimentsentail certain known (e.g., the die-off of the Pseudonitzschia sp. popula-tions) and unknown artifacts that could not be coveredwhenmodellingsuch an idealised system. Nonetheless, a small percentage of the largerphytoplankton cells began growing as early as the secondday of incuba-tion (Fig. 2), suggesting a size group-dependent growth response to ashift fromnutrient-limited to nutrient-replete conditions. Alternatively,these cells may have originated from the deeper, nutrient-rich watersuch that they were light-limited but not nutrient-limited at the begin-ning of the experiment, andwere thus susceptible to a light acclimationresponse rather thannutrient acclimation (data not shown). Transitionsin species' abundances, community composition, and nitrate uptakerates in response to changes in nutrient supply (i.e., upwelling) havebeen documented in field and experimental studies (Berges et al.,2004; MacIsaac et al., 1985). Using a similar experimental set-up tothat described here (Fawcett and Ward, 2011), Berges et al. (2004)observed a slower response to nitrate enrichment of nitrate reductaseactivity by a phytoplankton assemblage dominated by small flagellates(10–24 h) compared to one dominated by chain-forming diatoms(5–6 h). Unfortunately, we cannot assess such early-stage growth ornutrient uptake acclimation of the larger diatom cells because of thelack of observations during this early time window (0–24 h). Consider-ation of functional group- or size-specific acclimation rates in transientenvironments merits further scrutiny in the context of the most adap-tive strategy in cellular resource allocation in agreement with therecommendations of Menge et al. (2011).

4.3. Dissolved inorganic nitrogen uptake rates

Early in the mesocosm experiment, the largest phytoplankton sizefraction showed significantly higher RelVNO3

, although all size fractionseventually achieved similar maximum specific uptake rates (Table 2and Fig. 6). This fast response coupledwith themaintenance of high up-take rates appears to be the mechanism by which diatoms exploit

Page 11: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

24 N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

upwelling conditions (Fawcett and Ward, 2011). Using the calibratedmodel, we derived nutrient uptake rates for the various phytoplanktonsize fractions that were internally consistent and not confounded by thepresence of detritus in the initial, newly upwelled PN pool (Garside,1991). The maximum modelled VNO3

of the large, medium, and smallphytoplankton size groups (Table 2) was in the range of the values re-ported by Dugdale et al. (1990, 2006) based on the empirical relation-ship between nitrate uptake and initial nitrate concentration in anupwelling region (0.06–0.09 h−1). Although the maximum values ofmodelled VNO3

of the different phytoplankton groups were not signifi-cantly different from each other, the large phytoplankton groupattained its higher VNO3

earlier than the other size groups (Table 2 andFig. 6). The general trend of VNO3

in themodel representationmainly re-flects group-specific differences in nitrate uptake affinity and PCmax

values, which drive the uptake rate. In coastal upwelling regions,enhanced capacity for nitrate uptake confers a competitive advantageupon diatoms for which several mechanisms have been proposed:i) uptake inhibition decoupled from cellular nutrient quota due to vac-uolar nutrient storage (Lomas and Glibert, 2000; Stolte and Riegman,1995), ii) lower temperature optima for nitrate and nitrite reductasesuggesting a possible physiological adaptation to thrive in cool, NO3-rich environments (Lomas and Glibert, 1999), or iii) a higher densityof transport sites inferred from the isometric scaling of maximumuptake rates (Aksnes and Cao, 2011; Marañón et al., 2013). The lowercellular surface area per biomass and the similar VNO3

of the large phy-toplankton group compared to the smaller groups in our model favourexplanation iii) above but do not exclude the other options. Overall,we found that when considering the full model of ecological function-ing, phytoplankton group-specific differences in PCmax and nutrient af-finities were sufficient to represent group-specific nutrient dynamics,avoiding the addition of an unnecessary layer of complexity derivedfrom the interpretational framework of VNO3

.

5. Conclusions

Applying the TOPAZ ecosystem model to accurately reproduce theobserved patterns and timescales of size-partitioned new productionin a non-steady state environment identified the importance of group-specific functional traits, through adjustments of group-dependentmaximum production rates. This feature of phytoplankton physiologyis thus recognised as critical in determining species' response and rela-tive success in transient environments such as during coastal upwelling.It allows the diatom-dominated large size fraction in particular to re-spond rapidly to an injection of nitrate into surface water and accountsfor the ability of large cells to attain relatively higher specific growthrates on nitrate early on. The phytoplankton size group-dependentextent of downward revision from the theoretical maximum growthrates originally used in TOPAZ reflects the differences in taxonomiccomposition of each size group, the averaging effect of amalgamatingclosely related species adapted to thrive in different environments,and the sub-optimal growth of the phytoplankton community as awhole during transient environmental conditions. Potential limitationsof the parsimonious representation of physiology and biodiversitymay arise in, for example, post-upwelling conditions when largergrazers catch-up with the growth of large phytoplankton cells anddiverse ecological strategies (e.g., mechanical and chemical grazingdeterrence, mixotrophy or spore formation) may help in shaping thecommunity structure and export flux.While strong shifts in biodiversitywere observed within the largest, most competitive phytoplanktongroup, representation of only an averaged version of the most compet-itive of species in this group (diatoms) was sufficient to fit the observedbiogeochemical changes in this idealised, nutrient replete framework, inthe absence of large grazers. With the advent of physical models usinghigher spatial resolution more appropriate to representing coastalregions, inclusion of size group- or phylogenetic group-dependentmax-imum photosynthetic rates could benefit the ability of ecosystem

models to predict the dynamics of the substantial nutrient and organicmatter fluxes occurring on continental shelves.

Acknowledgements

We thank B. Song and P. Bhadury for the assistance with the samplecollection and incubation experiments. The staff of Moss LandingMarine Laboratory and Captain L. Bradford of the R/V ‘John Martin’were most cooperative and helpful. We are grateful to A. Gibson forperforming the phytoplankton microscopy counts, to A. Babbin and O.Coyle for the measurements of ammonium concentrations, to C. Stockand A. Babbin for the stimulating discussions and anonymous reviewersfor their comments and suggestions which helped to improve themanuscript. This research was supported by the NOAA Climate ProgramOffice grant n° NA110AR4310058.

Appendix A. Supplementary data

Supplementary data to this article can be found online at http://dx.doi.org/10.1016/j.jmarsys.2015.01.009.

References

Aksnes, D.L., Cao, F.J., 2011. Inherent and apparent traits inmicrobial nutrient uptake.Mar.Ecol. Prog. Ser. 440, 41–51.

Barber, R.T., Hiscock, M.R., 2006. A rising tide lifts all phytoplankton: growth response ofother phytoplankton taxa in diatom-dominated blooms. Glob. Biogeochem. Cycle 20.

Berges, J.A., Gibson, C.E., Stewart, B.M., 2004. Physiological responses of phytoplanktoncommunities in the Irish Sea to simulated upwelling. Hydrobiologia 517, 121–132.

Bissinger, J.E., Montagnes, D.J.S., Sharples, J., Atkinson, D., 2008. Predicting marine phyto-plankton maximum growth rates from temperature: improving on the Eppley curveusing quantile regression. Limnol. Oceanogr. 53, 487–493.

Collos, Y., 1982. Transient situations in nitrate assimilation by marine diatoms — short-term uncoupling of nitrate uptake and reduction. J. Exp. Mar. Biol. Ecol. 62, 285–295.

Collos, Y., 1986. Time-lag algal growth dynamics — biological constraints on primaryproduction in aquatic environments. Mar. Ecol. Prog. Ser. 33, 193–206.

Collos, Y., Vaquer, A., Souchu, P., 2005. Acclimation of nitrate uptake by phytoplankton tohigh substrate levels. J. Phycol. 41, 466–478.

Dugdale, R.C., Goering, J.J., 1967. Uptake of new and regenerated forms of nitrogen inprimary productivity. Limnol. Oceanogr. 12, 196–206.

Dugdale, R.C., Wilkerson, F.P., 1989. New production in the upwelling center at PointConception, California: temporal and spatial patterns. Deep Sea Res. A Oceanogr.Res. Pap. 36, 985–1007.

Dugdale, R.C., Wilkerson, F.P., Morel, A., 1990. Realization of new production in coastalupwelling areas — a means to compare relative performance. Limnol. Oceanogr. 35,822–829.

Dugdale, R.C., Wilkerson, F.P., Hogue, V.E., Marchi, A., 2006. Nutrient controls on newproduction in the Bodega Bay, California, coastal upwelling plume. Deep-Sea Res. IITop. Stud. Oceanogr. 53, 3049–3062.

Dunne, J.P., Armstrong, R.A., Gnanadesikan, A., Sarmiento, J.L., 2005. Empirical and mech-anistic models for the particle export ratio. Glob. Biogeochem. Cycles 19, 16.

Dunne, J.P., Sarmiento, J.L., Gnanadesikan, A., 2007. A synthesis of global particle exportfrom the surface ocean and cycling through the ocean interior and on the seafloor.Glob. Biogeochem. Cycles 21.

Dunne, J.P., John, J.G., Shevliakova, E., Stouffer, R.J., Krasting, J.P., Malyshev, S.L., Milly,P.C.D., Sentman, L.T., Adcroft, A.J., Cooke, W., Dunne, K.A., Griffies, S.M., Hallberg,R.W., Harrison, M.J., Levy, H., Wittenberg, A.T., Phillips, P.J., Zadeh, N., 2013. GFDL'sESM2 global coupled climate-carbon earth system models. Part II: carbon systemformulation and baseline simulation characteristics. J. Clim. 26, 2247–2267.

Edwards, K.F., Thomas, M.K., Klausmeier, C.A., Litchman, E., 2012. Allometric scalingand taxonomic variation in nutrient utilization traits and maximum growth rate ofphytoplankton. Limnol. Oceanogr. 57, 554–566.

Eppley, R.W., 1972. Temperature and phytoplankton growth in sea. Fish. Bull. 70,1063–1085.

Eppley, R.W., Peterson, B.J., 1979. Particulate organic-matter flux and planktonic newproduction in the deep ocean. Nature 282, 677–680.

Estrada, M., Blasco, D., 1985. Phytoplankton assemblages in coastal upwelling areas. In:Bas, C., Margalef, R., Rubias, P. (Eds.), International Symposium of Upwelling ofWestern Africa. Instituto de Investigaciones Pesqueras, Barcelona.

Fawcett, S.E., Ward, B.B., 2011. Phytoplankton succession and nitrogen utilization duringthe development of an upwelling bloom. Mar. Ecol. Prog. Ser. 428, 13–31.

Finkel, Z.V., 2001. Light absorption and size scaling of light-limited metabolism in marinediatoms. Limnol. Oceanogr. 46, 86–94.

Finkel, Z.V., Beardall, J., Flynn, K.J., Quigg, A., Rees, T.A.V., Raven, J.A., 2010. Phytoplanktonin a changing world: cell size and elemental stoichiometry. J. Plankton Res. 32,119–137.

Follows, M.J., Dutkiewicz, S., Grant, S., Chisholm, S.W., 2007. Emergent biogeography ofmicrobial communities in a model ocean. Science 315, 1843–1846.

Page 12: Journal of Marine Systems - The Fawcett Lab · Phytoplankton succession explains size-partitioning of new production ... For example, maximum potential ... Smaller cells (b5 μm)

25N. Van Oostende et al. / Journal of Marine Systems 148 (2015) 14–25

Frost, B.W., Franzen, N.C., 1992. Grazing and iron limitation in the control of phytoplank-ton stock and nutrient concentration — a chemostat analog of the pacific equatorialupwelling zone. Mar. Ecol. Prog. Ser. 83, 291–303.

Furnas, M.J., 1990. In situ growth-rates of marine-phytoplankton — approaches tomeasurement, community and species growth-rates. J. Plankton Res. 12, 1117–1151.

Garcia, H.E., Locarnini, R.A., Boyer, T.P., Antonov, J.I., Zweng, M.M., Baranova, O.K., Johnson,D.R., 2010. World Ocean Atlas 2009: Nutrients (Phosphate, Nitrate, Silicate) vol. 4.U.S. Government Printing Office, Washington, D.C.

Garrison, D.L., Gibson, A., Coale, S.L., Gowing, M.M., Okolodkov, Y.B., Fritsen, C.H., Jeffries,M.O., 2005. Sea-ice microbial communities in the Ross Sea: autumn and summerbiota. Mar. Ecol. Prog. Ser. 300, 39–52.

Garside, C., 1991. Shift-up and the nitrate kinetics of phytoplankton in upwelling systems.Limnol. Oceanogr. 36, 1239–1244.

Geider, R.J., 1992. Respiration — taxation without representation. In: Falkowski, P.G.,Woodhead, A.D. (Eds.), Primary Productivity and Biogeochemical Cycles in the Sea,Book 43.

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

Goebel, N.L., Edwards, C.A., Zehr, J.P., Follows, M.J., 2010. An emergent community ecosys-tem model applied to the California Current System. J. Mar. Syst. 83, 221–241.

Goericke, R., 2011. The structure of marine phytoplankton communities-patterns, rulesand mechanisms. Calif. Coop. Ocean. Fish. Invest. Rep. 52, 182–197.

Hansen, P.J., Bjørnsen, P.K., Hansen, B.W., 1997. Zooplankton grazing and growth: scalingwithin the 2–2,000 μm body size range. Limnol. Oceanogr. 42, 687–704.

Hasle, G.R., Syvertsen, E.E., 1997. Marine diatoms. In: Tomas, C.R. (Ed.), IdentifyingMarinePhytoplankton. Academic Press Inc., San Diego.

Holling, C.S., 1959. The components of predation as revealed by a study of small-mammalpredation of the European pine sawfly. Can. Entomol. 91, 293–320.

Ingraham, J.L., Maaløe, O., Neidhardt, F.C., 1983. Growth of the Bacterial Cell. SinauerAssociates.

Irigoien, X., Huisman, J., Harris, R.P., 2004. Global biodiversity patterns of marine phyto-plankton and zooplankton. Nature 429, 863–867.

Irwin, A.J., Finkel, Z.V., Schofield, O.M.E., Falkowski, P.G., 2006. Scaling-up from nutrientphysiology to the size–structure of phytoplankton communities. J. Plankton Res. 28,459–471.

Jochem, F.J., Smith, G.J., Gao, Y., Zimmerman, R.C., Cabello-Pasini, A., Kohrs, D.G., Alberte,R.S., 2000. Cytometric quantification of nitrate reductase by immunolabeling in themarine diatom Skeletonema costatum. Cytometry 39, 173–178.

Johnson, K.S., Chavez, F.P., Elrod, V.A., Fitzwater, S.E., Pennington, J.T., Buck, K.R., Walz,P.M., 2001. The annual cycle of iron and the biological response in central Californiacoastal waters. Geophys. Res. Lett. 28, 1247–1250.

Key, T., McCarthy, A., Campbell, D.A., Six, C., Roy, S., Finkel, Z.V., 2010. Cell size trade-offsgovern light exploitation strategies in marine phytoplankton. Environ. Microbiol. 12,95–104.

Klausmeier, C.A., Litchman, E., Daufresne, T., Levin, S.A., 2004. Optimal nitrogen-to-phosphorus stoichiometry of phytoplankton. Nature 429, 171–174.

Klein, L., 2002. An Ecosystem Dynamics Model of Monterey Bay, California. MSc. Univer-sity of Maine.

Kudela, R.M., Banas, N.S., Barth, J.A., Frame, E.R., Jay, D.A., Largier, J.L., Lessard, E.J., Peterson,T.D., Woude, A.J.V., 2008. New insights into the controls and mechanisms of planktonproductivity in coastal upwelling waters of the Northern California Current System.Oceanography 21, 46–59.

Lagarias, J.C., Reeds, J.A., Wright, M.H., Wright, P.E., 1998. Convergence properties of theNelder–Mead simplex method in low dimensions. SIAM J. Optim. 9, 112–147.

Lassiter, A.M., Wilkerson, F.P., Dugdale, R.C., Hogue, V.E., 2006. Phytoplankton assem-blages in the CoOP-WEST coastal upwelling area. Deep-Sea Res. II Top. Stud.Oceanogr. 53, 3063–3077.

Li, W.K.W., 2002. Macroecological patterns of phytoplankton in the northwestern NorthAtlantic Ocean. Nature 419, 154–157.

Li, Q.P., Franks, P.J.S., Landry, M.R., Goericke, R., Taylor, A.G., 2010. Modeling phytoplank-ton growth rates and chlorophyll to carbon ratios in California coastal and pelagicecosystems. J. Geophys. Res. Biogeosci. 115.

Lipschultz, F., 2008. Isotope tracer methods for studies of the marine nitrogen cycle. In:Capone, D.G., Bronk, D.A., Mulholland, M.R., Carpenter, E.J. (Eds.), Nitrogen in theMarine Environment.

Litchman, E., Klausmeier, C.A., Schofield, O.M., Falkowski, P.G., 2007. The role of functionaltraits and trade-offs in structuring phytoplankton communities: scaling from cellularto ecosystem level. Ecol. Lett. 10, 1170–1181.

Litchman, E., Pinto, P.D., Klausmeier, C.A., Thomas, M.K., Yoshiyama, K., 2010. Linkingtraits to species diversity and community structure in phytoplankton. Hydrobiologia653, 15–28.

Lomas, M.W., Glibert, P.M., 1999. Temperature regulation of nitrate uptake: a novel hy-pothesis about nitrate uptake and reduction in cool-water diatoms. Limnol. Oceanogr.44, 556–572.

Lomas, M.W., Glibert, P.M., 2000. Comparisons of nitrate uptake, storage, and reduction inmarine diatoms and flagellates. J. Phycol. 36, 903–913.

Longhurst, A., Sathyendranath, S., Platt, T., Caverhill, C., 1995. An estimate of global prima-ry production in the ocean from satellite radiometer data. J. Plankton Res. 17,1245–1271.

MacIsaac, J.J., Dugdale, R.C., Barber, R.T., Blasco, D., Packard, T.T., 1985. Primary productioncycle in an upwelling center. Deep-Sea Res. A Oceanogr. Res. Pap. 32, 503–529.

Marañón, E., Cermeño, P., López-Sandoval, D.C., Rodriguez-Ramos, T., Sobrino, C., Huete-Ortega, M., Blanco, J.M., Rodriguez, J., 2013. Unimodal size scaling of phytoplanktongrowth and the size dependence of nutrient uptake and use. Ecol. Lett. 16, 371–379.

Margalef, R., 1978. Life-forms of phytoplankton as survival alternatives in an unstableenvironment. Oceanol. Acta 1, 493–509.

Menden-Deuer, S., Lessard, E.J., 2000. Carbon to volume relationships for dinoflagellates,diatoms, and other protist plankton. Limnol. Oceanogr. 45, 569–579.

Menge, D.N.L., Ballantyne, F., Weitz, J.S., 2011. Dynamics of nutrient uptake strategies:lessons from the tortoise and the hare. Theor. Ecol. 4, 163–177.

Moloney, C.L., Field, J.G., Lucas, M.I., 1991. The size-based dynamics of plankton foodwebs. II. Simulations of three contrasting southern Benguela food webs. J. PlanktonRes. 13, 1039–1092.

Morel, F.M.M., 1987. Kinetics of nutrient-uptake and growth in phytoplankton. J. Phycol.23, 137–150.

Muller-Karger, F.E., Varela, R., Thunell, R., Luerssen, R., Hu, C.M., Walsh, J.J., 2005. The im-portance of continental margins in the global carbon cycle. Geophys. Res. Lett. 32.

Olenina, I., Hajdu, S., Edler, L., Andersson, A., Wasmund, N., Busch, S., Göbel, J., Gromisz, S.,Huseby, S., Huttunen, M., Jaanus, A., Kokkonen, P., Ledaine, I., Niemkiewicz, E., 2006.Biovolumes and size-classes of phytoplankton in the Baltic Sea. HELCOM Baltic SeaEnvironment Proceedings 106, p. 144.

Poulin, F.J., Franks, P.J.S., 2010. Size-structured planktonic ecosystems: constraints,controls and assembly instructions. J. Plankton Res. 32, 1121–1130.

Sakshaug, E., Holm-Hansen, O., 1977. Chemical composition of Skeletonema costatum(grev) cleve and Pavlova (monochrysis) lutheri (droop) green as a function ofnitrate-limited, phosphate-limited, and iron-limited growth. J. Exp. Mar. Biol. Ecol.29, 1–34.

Smith, G.J., Zimmerman, R.C., Alberte, R.S., 1992. Molecular and physiological responses ofdiatoms to variable levels of irradiance and nitrogen availability — growth ofSkeletonema costatum in simulated upwelling conditions. Limnol. Oceanogr. 37,989–1007.

Steele, J.H., Henderson, E.W., 1992. The significance of interannual variability. In: Evans,G.T., Fasham, M.J.R. (Eds.), Towards a Model of Ocean Biogeochemical Processes.Springer-Verlag, Heidelberg, Germany.

Steidinger, K.A., Tangen, K., 1997. Dinoflagellates. In: Tomas, C.R. (Ed.), IdentifyingMarinePhytoplankton. Academic Press Inc., San Diego.

Stock, C., Dunne, J., 2010. Controls on the ratio of mesozooplankton production to primaryproduction in marine ecosystems. Deep-Sea Res. I Oceanogr. Res. Pap. 57, 95–112.

Stolte, W., Riegman, R., 1995. Effect of phytoplankton cell-size on transient-state nitrateand ammonium uptake kinetics. Microbiology-UK 141, 1221–1229.

Straile, D., 1997. Gross growth efficiencies of protozoan and metazoan zooplankton andtheir dependence on food concentration, predator–prey weight ratio, and taxonomicgroup. Limnol. Oceanogr. 42, 1375–1385.

Sunda, W.G., Huntsman, S.A., 1995. Iron uptake and growth limitation in oceanic andcoastal phytoplankton. Mar. Chem. 50, 189–206.

Throndsen, J., 1997. The planktonic marine flagellates. In: Tomas, C.R. (Ed.), IdentifyingMarine Phytoplankton. Academic Press Inc., San Diego.

Thunell, R., Benitez-Nelson, C., Varela, R., Astor, Y., Muller-Karger, F., 2007. Particulateorganic carbon fluxes along upwelling-dominated continental margins: rates andmechanisms. Glob. Biogeochem. Cycles 21.

Venrick, E.L., 2009. Floral patterns in the California Current: the coastal-offshore boundaryzone. J. Mar. Res. 67, 89–111.

Wilkerson, F.P., Dugdale, R.C., 1987. The use of large shipboard barrels and drifters tostudy the effects of coastal upwelling on phytoplankton dynamics. Limnol. Oceanogr.32, 368–382.

Wilkerson, F.P., Lassiter, A.M., Dugdale, R.C., Marchi, A., Hogue, V.E., 2006. The phyto-plankton bloom response to wind events and upwelled nutrients during the CoOPWEST study. Deep-Sea Res. II Top. Stud. Oceanogr. 53, 3023–3048.

Wood, A.M., Everroad, R.C., Wingard, L.M., 2005. Measuring growth rates in microalgalcultures. In: Andersen, R.A. (Ed.), Algal Culturing Techniques. Elsevier AcademicPress, London.

Zimmerman, R.C., Kremer, J.N., Dugdale, R.C., 1987. Acceleration of nutrient-uptake byphytoplankton in a coastal upwelling ecosystem — a modeling analysis. Limnol.Oceanogr. 32, 359–367.


Recommended