Modelling the effect of vertical mixing on bottle incubations for...

19
MARINE ECOLOGY PROGRESS SERIES Mar Ecol Prog Ser Vol. 435: 13–31, 2011 doi: 10.3354/meps09193 Published August 22 INTRODUCTION The relevance of marine phytoplankton as a food for the pelagos and as a potential sink of atmospheric carbon is widely recognised. Marine phytoplankton stands at the base of the pelagic food webs and may act as a potential sink of atmospheric carbon. The ability to determine reliably in situ phytoplankton primary pro- ductivity and growth is crucial for understanding the dynamics of aquatic ecosystems, for climate predic- tions, and for the management of water quality. While primary production measurements try to quantify the conversion of sunlight and inorganic substances into organic compounds in units of grams carbon per unit time and unit area or volume, the growth rates express the increase in the phytoplankton population in units of inverse time. Typical approaches for productivity measurements include the 14 C and O 2 methods (Will- iams et al. 1983), while growth rates are sometimes inferred from the changes in total cell numbers (Rhee & Gotham 1981, Kagami & Urabe 2001), or — more commonly — total chlorophyll a (Chl) (Landry & Has- sett 1982, Bienfang & Takahashi 1983, Calbet & Landry 2004) or taxon-specific pigments (Latasa et al. 1997). © Inter-Research 2011 · www.int-res.com *Email: [email protected] Modelling the effect of vertical mixing on bottle incubations for determining in situ phytoplankton dynamics. I. Growth rates Oliver N. Ross 1, *, Richard J. Geider 2 , Elisa Berdalet 3 , Mireia L. Artigas 3 , Jaume Piera 1 1 Mediterranean Centre for Marine & Environmental Research, Unidad de Tecnología Marina (UTM, CSIC), Passeig Marítim de la Barceloneta 37-49, 08003 Barcelona, Spain 2 Dept. Biological Sciences, University of Essex, Wivenhoe Park, Colchester CO4 3SQ, UK 3 Instituto de Ciencias del Mar (ICM, CSIC), Passeig Marítim de la Barceloneta 37-49, 08003 Barcelona, Spain ABSTRACT: Reliable estimates of in situ phytoplankton growth rates are central to understanding the dynamics of aquatic ecosystems. A common approach for estimating in situ growth rates is to incu- bate natural phytoplankton assemblages in clear bottles at fixed depths or irradiance levels and mea- sure the change in chlorophyll a (Chl) over the incubation period (typically 24 h). Using a modelling approach, we investigate the accuracy of these Chl-based methods focussing on 2 aspects: (1) in a freely mixing surface layer, the cells are typically not in balanced growth, and with photoacclimation, changes in Chl may yield different growth rates than changes in carbon; and (2) the in vitro methods neglect any vertical movement due to turbulence and its effect on the cells’ light history. The growth rates thus strongly depend on the incubation depth and are not necessarily representative of the depth-integrated in situ growth rate in the freely mixing surface layer. We employ an individual based turbulence and photosynthesis model, which also accounts for photoacclimation and photo- inhibition, to show that the in vitro Chl-based growth rate can differ both from its carbon-based in vitro equivalent and from the in situ value by up to 100%, depending on turbulence intensity, optical depth of the mixing layer, and incubation depth within the layer. We make recommendations for choosing the best depth for single-depth incubations. Furthermore we demonstrate that, if incubation bottles are being oscillated up and down through the water column, these systematic errors can be significantly reduced. In the present study, we focus on Chl-based methods only, while productivity measurements using carbon-based techniques (e.g. 14 C) are discussed in Ross et al. (2011; Mar Ecol Prog Ser 435:33–45). KEY WORDS: Chlorophyll · Photoacclimation · Photoinhibition · Lagrangian modelling · Individual based modelling · Turbulence · Surface mixing layer · Yo-yo technique Resale or republication not permitted without written consent of the publisher OPEN PEN ACCESS CCESS

Transcript of Modelling the effect of vertical mixing on bottle incubations for...

Page 1: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

MARINE ECOLOGY PROGRESS SERIESMar Ecol Prog Ser

Vol. 435: 13–31, 2011doi: 10.3354/meps09193

Published August 22

INTRODUCTION

The relevance of marine phytoplankton as a food forthe pelagos and as a potential sink of atmospheric carbon is widely recognised. Marine phytoplanktonstands at the base of the pelagic food webs and may actas a potential sink of atmospheric carbon. The ability todetermine reliably in situ phytoplankton primary pro-ductivity and growth is crucial for understanding thedynamics of aquatic ecosystems, for climate predic-tions, and for the management of water quality. Whileprimary production measurements try to quantify the

conversion of sunlight and inorganic substances intoorganic compounds in units of grams carbon per unittime and unit area or volume, the growth rates expressthe increase in the phytoplankton population in unitsof inverse time. Typical approaches for productivitymeasurements include the 14C and O2 methods (Will -iams et al. 1983), while growth rates are sometimesinferred from the changes in total cell numbers (Rhee& Gotham 1981, Kagami & Urabe 2001), or — morecommonly — total chlorophyll a (Chl) (Landry & Has-sett 1982, Bienfang & Takahashi 1983, Calbet & Landry2004) or taxon-specific pigments (Latasa et al. 1997).

© Inter-Research 2011 · www.int-res.com*Email: [email protected]

Modelling the effect of vertical mixing on bottleincubations for determining in situ phytoplankton

dynamics. I. Growth rates

Oliver N. Ross1,*, Richard J. Geider2, Elisa Berdalet3, Mireia L. Artigas3, Jaume Piera1

1Mediterranean Centre for Marine & Environmental Research, Unidad de Tecnología Marina (UTM, CSIC), Passeig Marítim de la Barceloneta 37-49, 08003 Barcelona, Spain

2Dept. Biological Sciences, University of Essex, Wivenhoe Park, Colchester CO4 3SQ, UK 3Instituto de Ciencias del Mar (ICM, CSIC), Passeig Marítim de la Barceloneta 37-49, 08003 Barcelona, Spain

ABSTRACT: Reliable estimates of in situ phytoplankton growth rates are central to understanding thedynamics of aquatic ecosystems. A common approach for estimating in situ growth rates is to incu-bate natural phytoplankton assemblages in clear bottles at fixed depths or irradiance levels and mea-sure the change in chlorophyll a (Chl) over the incubation period (typically 24 h). Using a modellingapproach, we investigate the accuracy of these Chl-based methods focussing on 2 aspects: (1) in afreely mixing surface layer, the cells are typically not in balanced growth, and with photoacclimation,changes in Chl may yield different growth rates than changes in carbon; and (2) the in vitro methodsneglect any vertical movement due to turbulence and its effect on the cells’ light history. The growthrates thus strongly depend on the incubation depth and are not necessarily representative of thedepth-integrated in situ growth rate in the freely mixing surface layer. We employ an individualbased turbulence and photosynthesis model, which also accounts for photoacclimation and photo -inhibition, to show that the in vitro Chl-based growth rate can differ both from its carbon-based invitro equivalent and from the in situ value by up to 100%, depending on turbulence intensity, opticaldepth of the mixing layer, and incubation depth within the layer. We make recommendations forchoosing the best depth for single-depth incubations. Furthermore we demonstrate that, if incubationbottles are being oscillated up and down through the water column, these systematic errors can besignificantly reduced. In the present study, we focus on Chl-based methods only, while productivitymeasurements using carbon-based techniques (e.g. 14C) are discussed in Ross et al. (2011; Mar EcolProg Ser 435:33–45).

KEY WORDS: Chlorophyll · Photoacclimation · Photoinhibition · Lagrangian modelling · Individualbased modelling · Turbulence · Surface mixing layer · Yo-yo technique

Resale or republication not permitted without written consent of the publisher

OPENPEN ACCESSCCESS

Page 2: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 201114

Current knowledge and technology do not permit usto accurately determine the real in situ growth in nat-ural cells that move freely through the water column. Awide range of sophisticated in vitro methods have beendeveloped to measure phytoplankton growth rates at aspecies or group level (Stolte & Garcés 2006). A com-mon approach is to place a natural population into aclosed container (polycarbonate bottle or carboy) andto determine the increase in phytoplankton Chl duringthe incubation period (e.g. Bienfang & Takahashi 1983,Furuya et al. 1986, Furnas 1990). Since this methodonly delivers the net growth of the population, it wasfurther refined to the so-called dilution method(Landry & Hassett 1982, Landry et al. 1995) to provideadditional information on the gross growth rate andthus the grazing pressure (see Calbet & Landry 2004,Behrenfeld 2010, Gutiérrez-Rodríguez et al. 2010, andreferences therein).

The sampling bottles are often incubated at a singlefixed depth (light level), typically the same depthwhere the samples were taken. Fixed-depth bottleincubations do not necessarily provide a good repre-sentation of the temporal and spatial variability in thephysical underwater environment, however, as phyto-plankton is constantly being moved through the sur-face mixing layer (SML) (Lewis et al. 1984b) and cellsbecome exposed to a wide range of light in tensities(see Fig.1 in Ross et al. 2008). Phytoplankton organ-isms possess a number of mechanisms to cope with thestresses imposed by fluctuating light (Mac Intyre et al.2000). Amongst the best understood of these mecha-nisms are photoacclimation, which in volves changes incellular pigment content, and photoprotection, whichinvolves reducing the efficiency of excitation energytransfer from pigments to reaction centres (Falkowski1983, Lewis et al. 1984a, Geider et al. 1998). Failure toalleviate the light stress by photoacclimation and/orphotoprotection may result in photo inhibition of photo-synthesis and, in extreme cases, cell death (Eilers &Peeters 1988, Long et al. 1994, Behrenfeld et al. 1998,Ross et al. 2008). These processes are relevant to as -sessing phytoplankton growth from incubation ex peri -ments, because the overall performance (growth) of acell, as well as the relationship between chloro phyllsynthesis, photosynthesis, and cell growth will varymarkedly with the cell’s light history. One problem thatmust be faced by the experimentalist is to determine atwhich depth in the water column (e.g. at which lightlevel) to conduct an incubation in order to obtain accu-rate measurements of the in situ growth rate.

The problem in using fixed-depth/irradiance bottleincubation experiments to estimate the in situ growthrate is thus 2-fold.

(1) Differences between growth rates estimated fromchanges in chlorophyll concentration and carbon: A

sample of a natural phytoplankton population that istaken from a single depth will contain cells with differ-ent light histories and thus different photoacclimationstates (e.g. different Chl:C and different degrees ofphotoinhibition). Very rarely will a particular cell befully adapted to the light intensity at the depth atwhich it is found at a given instant in time (see Fig. 3).By incubating cells at a fixed depth, their residencetime at the incubation depth is artificially increased,which provides the opportunity to (1) photo acclimateto the ambient light level, resulting in uncoupling ofchlorophyll synthesis from CO2 fixation, (2) accumu-late photoinhibitory damage if cells are incubated athigh light intensities too close to the surface, or (3)repair photoinhibitory damage when incubated at lowlight deeper in the water column. In general, it is there-fore not justified to assume that phytoplankton growthis balanced, i.e. that the increase in chlorophyll ismatched by the increase in carbon so that the formermay be used as proxy for the determination of the carbon-based growth rate μC. Although most experi-mentalists are aware of this problem (Brown et al.2002, Latasa et al. 2005, Gutiérrez-Rodríguez et al.2010) it remains difficult to provide an accurate quan-tification of the associated error.

(2) The growth rate varies with incubation depth andmay not be representative of the entire SML: It is notclear a priori, that cells incubated at a single fixeddepth (and thus at relatively stable light conditions)should grow at a rate that is representative of theentire layer. Particularly in coastal areas or estuaries,where strong turbulent mixing in combination withhigh light attenuation coefficients may subject the cellsto large and rapid changes in light intensity (e.g. Lizonet al. 1998, Sharples et al. 2001, Moore et al. 2006, Ross& Sharples 2007), or during deep mixing events in theopen ocean in winter and spring (e.g. Fig. 1 in Ross etal. 2008, Woods & Onken 1982, Nagai et al. 2003), theassociated error can become quite large. Alternativeapproaches have been devised which aim to alleviatethis problem by simulating the large-scale vertical dis-placements due to turbulence by moving the incuba-tion bottles up and down through the water column orusing on-deck incubators that expose the samples tovariations in light intensity (Marra 1978, Köhler 1997,Köhler et al. 2001, Gocke & Lenz 2004). While both theyo-yo technique and the linear incubator appear to bepromising approaches to overcome the limitationsassociated with fixed-depth incubations, they bothcome with the caveat that the choice of the amplitudeand period of the vertical oscillations/light variations isan ad hoc decision which is often taken without muchprior knowledge of the in situ mixing conditions.

In the present study we employ a state-of-the-artindividual-based model of turbulence and phytoplank-

Page 3: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 15

ton growth, acclimation, and inhibition to compare theChl-based growth rates obtained from fixed-depth bot-tle incubations with the carbon equivalent and thegrowth rates obtained for a population that was freelymixing in the SML. As it is currently not possible tomeasure the in situ growth rate of a freely mixingphytoplankton population non-intrusively, the use of amodel seems the best approach. The model consists of2 components: (1) a Lagrangian turbulence module,which provides the individual particle tracks and asso-ciated light histories in response to turbulence mixing,and (2) a biological module, which calculates thegrowth rates based on the individual light histories,taking into account photoacclimation and photoinhibi-tion. The Lagrangian technique (Woods & Onken 1982,Wolf & Woods 1988, Lizon et al. 1998, Broekhuizen1999, Ross & Sharples 2007) seems a natural choice toapproach this problem, as its Eulerian counterpart can-not provide the necessary information on the individ-ual light histories. We simulate shallow and deep, tur-bid and clear, turbulent and quiescent water columnsin order to quantify the differences between thegrowth rates for a range of natural environments. Wepropose remedies and recommendations for the exper-imental protocol that will help to minimise the errorsassociated with fixed-depth incubations. We alsoexamine the extent to which a vertically oscillatingbottle incubation can deliver more accurate estimatesof the in situ growth rates. In a companion paper (Rosset al. 2011, this volume), we investigate how fixeddepth incubations affect estimates of primary produc-tivity, which are usually based on 14C in a series ofincubations that are suspended at fixed depthsthroughout the mixing layer.

METHOD

With the vertical being the key dimension for ourproblem it is sufficient to employ a 1D vertical physical-biological model. Each of the 2 main components (theLagrangian model and the biological module) isdescribed in more detail in the following sections.

All simulations run for a total of 5 d (4 d of spin-upplus 1 d for measuring) and the particles are mixedusing an eddy diffusivity in combination with a ran-dom walk approach (see below). In the runs whichsimulate the fixed-depth bottle incubations, the turbu-lent mixing is switched off at the beginning of the 5thday and all particles remain ‘frozen’ in place for 24 h.We then compare the growth rates, both within eachdepth bin and integrated over the entire layer, duringthe incubation period to the growth rates from theruns where particles were freely mixing on the 5thday. Both sets of experiments that were run for each

scenario thus share the same 4 d spin-up and only dif-fer on the last day.

Unless stated otherwise, all experiments used thesame peak mid-day incident irradiance of 650 W m−2,45% of which is considered to be photosyntheticallyactive radiation (PAR). This leaves a surface mid-dayPAR of about Im = 1200 μmol photons m−2 s−1 thatdecreases exponentially with depth as

I(z) = I(z=0) ekz = I(z=0) e−ξ (1)

where k is the PAR attenuation coefficient and z is thedepth (z < 0) in metres. The light availability thusalways depends on the product of k and −z, which isthe optical depth ξ. We therefore express the verticaldimension in our figures in terms of ξ. Over the courseof the light period (15 h per 24 h) the surface irradianceI(z=0) was varied using a semi-sinusoidal modulation.

Note that the term surface mixing layer (SML)denotes the isothermal or isopycnal layer which isactively mixing, i.e. where the eddy diffusivity is con-sistently higher than some background value. It is thusa purely physical definition and does not imply that thelayer is fully mixed in a biological sense such that thephysiology would be homogeneous throughout thelayer. This would depend on the ratio of the physicalmixing time-scale versus the biological time-scales ofphotoacclimation, -inhibition, and growth (see the dis-cussion in Ross et al. 2011, Lewis et al. 1984b). Whilemost mixing layers will have mixing time scales (τ) ofthe order of hours to few days, we also found reports ofmixing layers with mixing time scales (τm) of the orderof thousands of hours. In Table 1 of Gallegos & Platt(1985), for instance, the diffusivity at a station in theoligotrophic ocean was reported as 10−5 m2 s−1. Whilenothing is said about the depth of the SML, we wouldobtain a mixing time scale of 2500 h for a shallow SMLof 10 m. Such a layer can only remain mixed if it isoverturned at night due to convective instabilities. Forscenarios with τm > 100 h, we therefore simulated theconvective overturning by temporarily increasing theeddy diffusivity Km from its ‘normal’ daytime value to10−2 m2 s−1 for 5 h each night. This was not necessaryfor scenarios with τm ≤ 100 h, as the mixing was suffi-cient to keep the biomass homogeneously distributed.Table 1 contains a summary of the main model para-meters and their standard values.

Lagrangian particle tracking. We represent phyto-plankton using an individual based approach, i.e.phytoplankton are not modelled as a concentration ofcarbon or chlorophyll per unit volume (Eulerianapproach) but as individual cells that are moved freelythrough the water column by turbulence. The advan-tage of using a Lagrangian approach lies in its ability torecord the individual light histories of the cells, whichallows for a more accurate simulation of the photo-

Page 4: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 2011

physiological response. The total number of Lagran -gian particles in our simulations varied depending onthe total water depth. For simulations of shallow waterscenarios (H = 6 m), we used 8000 Lagrangian parti-cles, while in deeper water (H ≥ 60 m) we used 80 000.This high number was necessary in order to obtain reli-able statistics at each 1 m depth bin. We distinguishbetween Lagrangian particles (the number of whichstays fixed) and phytoplankton cells (which areallowed to divide and multiply). At the start of anexperiment, the Lagrangian particles are homoge-neously distributed with depth, and each particle rep-

resents the same number of phytoplankton cells(Lagrangian Ensemble approach). Over the course ofthe experiment, the distribution of Lagrangian parti-cles always remains homogeneous as we assume neu-tral buoyancy.

Given that most diatoms have sustained sinkingspeeds of <3 m d−1 (Eppley et al. 1967, Bienfang 1981,Waite et al. 1997, Ptacnik et al. 2003), we decided toneglect diatom sinking altogether, as tests with suchsinking velocities did not significantly alter our results.Motility on the other hand, i.e. the ability for directedswimming by dinoflagellates or coccolithophores,

16

Symbol Description Value Unit

aChl Chl a specific absorption coefficient 6.56 m2 gChl−1

CF Amount of carbon in functional pool Variable gC cell−1

C Total amount of carbon in cell Variable gC cell−1

Chl Amount of chlorophyll a in cell Variable gChl cell−1

D Depth of water column (model domain) Variable mH SML thickness Variable mI Light intensity Variable μmol photons m−2 s−1

Ik Light saturation parameter Variable μmol photons m−2 s−1

k PAR attenuation coefficient Variable m−1

kr Repair rate from photoinhibition 4.5 × 10−5 s−1

kd Damage probability for photoinhibition 1.4 × 10−7 –K Turbulent diffusivity Variable m2 s−1

Kbg Background turbulent diffusivity below SML 10−5 m2 s−1

Km Mid SML turbulent diffusivity Variable m2 s−1

Ksurf Turbulent diffusivity at surface boundary 0.1Km m2 s−1

P cellmax Max. cell-based C production Variable gC cell−1 s−1

Pchlm Max. Chl-based C production (with inhibition) Pchl

max ϑ gC gChl −1 s−1

Pchlmax Max. Chl-based C production (without inhibition) 3.3 × 10–3 gC gChl−1 s−1

PFmax Max. carbon-based production at a given N:C Variable s−1

z Particle position < 0 mαChl

max Initial slope of the P-I curve (without inhibition) 6.3 × 10−6 gC m2 (gChl μmol photons)−1

Δt Model time step 1–6a sε1 Error between μincub

Chl and μCincub Variable %

ε2 Error between μincubChl and 3μC

mixed Variable %φmax Max. quantum yield of photosynthesis 0.96 × 10−6 gC (μmol photons)−1

φm Achieved quantum yield when inhibited Variable gC (μmol photons)−1

μxincub x-derived growth rates from incubations Variable d−1

3μxmixed x-derived depth-integrated growth rates from turbulent water Variable d−1

ΠLmax Max. prop. of CF allocated to light harvesting 0.33 –

θC Chl:C ratio of cell Variable gChl gC−1

θCF Chl:C ratio for functional pool only Variable gChl gC−1

(θCL )max Max Chl:C ratio in light-harvesting pool 0.28 gChl gC−1

ϑ Proportion of functional PSII reaction centres 0 ≤ ϑ ≤ 1 –σPSII Absorption cross-section of PSII 1.5 m2 (μmol photons)−1

ξ Optical depth within water column Variable –ξSML Optical depth of the surface mixing layer Variable –τm Mixing time scale of surface mixing layer Variable h

aDepending on the value of Km

Table 1. Model parameters and meaning of symbols used throughout the text. A list of the remaining model parameters for the biological model can be found in Appendix 1 and Ross & Geider (2009, their Table 1). SML: surface mixing layer; PAR:

photosynthetically active radiation

Page 5: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 17

could potentially alter our results as different swim-ming strategies may lead to highly inhomogeneousparticle concentrations (turbulence permitting) andthus growth rates different from those obtained for ahomogeneously distributed population. As any choiceof swimming strategy (phototactic, geotactic, etc.)would have to be made arbitrarily, we omitted motilityfrom our analysis. Because the main goal of this workwas to evaluate whether vertical mixing by turbulencecan be neglected for the determination of growth rates,i.e. whether a fixed depth bottle incubation deliversresults that are representative of a freely mixing watercolumn, we could also neglect factors such as nutrientsor grazing, as their absence or presence would havethe same effect on both the stranded and freely mixingpopulations. We also did not concern ourselves withother sources of error, such as bottle effects, whichhave already been addressed elsewhere (see Furnas2002, for a review).

The one-dimensional (vertical) particle-trackingmodel is based on Ross & Sharples (2004). Vertical pro-files of the eddy diffusivity, K(z) (m2 s−1) (see below), incombination with a random walk approach were usedto provide the random, turbulent mixing of Lagrangianparticles. At each time step, Δt, a particle was movedfrom its present position, zn, to its new position, zn+1,using:

(2)where K’ = dK/dz. R is a random process of zero meanand variance r (e.g. r = 1/3 for R ∈ [−1, 1]). The top andbottom boundaries were reflecting according to

(3)

where H is the depth of the SML and D is the totaldepth of the water column.

We created a physical environment based both onobservational (e.g. Sharples et al. 2001) and modellingresults (e.g. Ross & Sharples 2007) of a turbulent SML,which fulfils all necessary requirements for use withthe random walk in terms of differentiability, steadi-ness, and reflecting boundary conditions (cf. Ross &Sharples 2004). The resulting diffusivity profile is con-stant throughout most of the mixing layer. The transi-tional areas where the diffusivity is allowed to de -crease in a continuous and differentiable fashiontowards the top and bottom boundaries are designed todeliver a time step greater than 1s in the random walkwhile keeping the error below 1% (cf. Ross & Sharples2004). The profile is constructed as follows:

(4)

where Kbg is the background diffusivity in the stratifiedarea below the SML, and Ksurf is the near surface diffu-sivity, which was set to between 20 and 90% of themid-SML value Km, depending on the profile. TheRoman numerals in Eq. (4) denote the 3 different seg-ments of the profile: [I] top half of SML, [II] lower halfof SML, and [III] below the SML (cf. Fig. 1).

We tested a wide range of environmental scenariosusing different combinations of attenuation coeffi-cients, mixing layer depths, and turbulent eddy diffu-sivities. The physical environment can be charac-terised using 2 important parameters, namely themixing time scale, τm, and the optical depth of the SML,ξSML, which are given by:

(5)

where H is the SML depth and Km the maximum (mid-SML) eddy diffusivity. Fig. 2 provides a visual repre-sentation of the effect of different mixing time scaleson the change of a particle distribution with time.

Biological model. The biological model used in thesimulations is based on the RGI model in Ross & Geider(2009). Rather than being solely carbon-based, the bio-logical model employs the cell as the basic unit, whichis a more natural choice for use in individual-basedmodelling approaches such as this one. The model dif-ferentiates between a nitrogen-containing functionalcarbon pool and an energy reserve pool, which does

z z K z t RK z

n n'

n

deterministic term

n

+ + ++

1 = ( )2

Δ

112

( )1/2

K z t t

r

'n

random term

Δ Δ( )⎡

⎢⎢⎢

⎥⎥⎥

zz z

H z z Dnn n

n n+

+ +

+ +→

−+ −

⎧⎨⎪

⎩⎪1

1 1

1 1

, > 0

2 , <

if

if

K z

K K K zmz

( ) =

( ) 1 1.5 0| |2.3surf surf for+ − −⎡⎣ ⎤⎦ ≥− >> /2

( ) 1 1.5 ( )2.3

+ − −⎡⎣ ⎤⎦ −− +

H

K K KmH z

[I]

forbg bg HH z H

K z H

/2 ( )

<

≥ ≥ −

⎨⎪⎪

⎩⎪⎪

[II]

for [III]bg

τ ξmm

HK

Hk= =2

and SML

−D

−H

−H/2

0

Dep

th

Kbg Ksurf Km

Eddy diffusivity

Segment I

Segment II

Segment III

SML

Fig. 1. Example diffusivity profile where the Roman numerals denote the 3 different segments from Eq. (4)

Page 6: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 2011

not contain nitrogen. The cell-specific light-saturatedphotosynthetic rate is assumed to scale with the size ofthe functional pool, and the light- limited photosynthe-sis rate with the cellular chlorophyll content. Themodel has been validated against data from light shiftexperiments on Skeletonema costatum (Anning et al.2000, Ross & Geider 2009) and the values of the growthand acclimation parameters we use in this study arerepresentative of this diatom. As we are not concernedabout absolute growth rates but content ourselves withcomparative analyses, the results presented in thisstudy are insensitive to small changes in these parame-ter values and are therefore deemed to be valid for awider range of phytoplankton species. Although wemade some modifications to the RGI model and addednew equations to account for photo inhibition, most ofthe model equations are un changed. We therefore donot re-iterate the entire model description but limit our

description to the new and modified parts of the model.A complete set of the biological model equations,including the unchanged parts from the original RGI

model, is contained in Appendix 1 (see Ross & Geider2009 for additional information).

We extend the RGI model from Ross & Geider (2009)to account for photoinhibition using a similar approachas in Cianelli et al. (2004) except that we apply theinhibition primarily to the initial slope of the photo -synthesis-light curve, αChl, and only once the damagebecomes more substantial also to the maximum cell-based carbon production, P cell

max. The proportion of stillfunctional reaction centres, ϑ, depends on the damageprobability, kd, and the recovery rate, kr, of the D1 pro-tein as (cf. Han 2002, Cianelli et al. 2004):

(6)

where σPSII is the absorption cross-section ofPhotosystem II (PSII), aChl the Chl specificabsorption coefficient and φmax the maximumquantum yield of photosynthesis (Table 1).The values chosen for kd and kr from Table 1are based on Cianelli et al. (2004), while thevalues for σPSII and φmax are based on Suggettet al. (2009) and MacIntyre et al. (2002),respectively. The value of aChl was deter-mined experimentally for Thalassiosirapseudonana (Ross & Geider 2009). The para-meter ϑ describes the proportion of still func-tional PSII reaction centres (RCII) (0 ≤ ϑ ≤ 1)with the extremes of ϑ = 1 corresponding tozero inhibition (i.e. all centres are functional)and the theoretical value of ϑ = 0 corre-sponding to 100% inhibition (with no func-tional centres remaining). We expect theprobability of photo inhibition (per photonabsorbed) to be greater when photosynthesisis light saturated and almost all of the reac-tion centres are closed than when photosyn-thesis is light limited and almost all of thereaction centres are open. This is expressedby the term

in Eq. (6) which is a proxy for reaction centreclosure. The denominator is the rate of lightabsorption by the cell (photons per cell perunit time) and the numerator is the rate ofuse of photons for photosynthesis.

When the maximum quantum efficiency ofphotosynthesis, φmax, is determined by the

dd

=1 (1 ) , if > 0

(1 ) , if = 0

PSII

cellmax

1

t

k IP

I a Chlk I

k I

d Chl r

r

( )ϑ − σ ϑ − φ⎛

⎝⎞⎠ + − ϑ

− ϑ

⎨⎪

⎩⎪

11

−⎛⎝⎜

⎞⎠⎟

−PI a ChlChl

cellmax( )φ

18

H

0

Dep

th

(a) τm=1h

H

0

Dep

th

H

0

Dep

th

H

0

(b) τm=10h

H

0

H

0

H

0

(c) τm=100h

H

0

H

0

H

0

(d) τm=1000h

H

0

H

0

+5.5h

+30min

t=0

Fig. 2. Effect of the mixing time scale τm on the vertical particle distribu-tions over the course of time. The top row shows the initial conditions attime t = 0, the middle row is after 30 min into the simulation, and the bot-tom row shows the particles after another 5.5 h, i.e. 6 h after the start of theexperiment. The particles were assigned a certain marker and colour

depending on their initial position in the water column

Page 7: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth

maximum quantum efficiency of PSII, we expect theinitial slope of the photosynthesis-light curve, αChl, tobe given by:

(7)

Hence instead of Eq. (14) from Ross & Geider (2009) weobtain the following modified equation for the cellbased production:

(8)

Previous work suggested that photoinhibition notonly affects the initial slope αChl but also the light satu-rated maximum production P cell

max (Behrenfeld et al.1998). We implemented this by changing Eq. (15) fromRoss & Geider (2009) to the following minimum condi-tion:

(9)

where CF and Chl are the amounts of functional carbonand chlorophyll per cell, respectively. PF

max is the maxi-mum carbon-based production at a given nitrogen-to-carbon (N:C) ratio and Pchl

m is the maximum Chl-basedcarbon production at a given inhibition level:

(10)

Implicit in Eq. (9) is the assumption that control of thelight saturated photosynthesis rate may lie either in thelight-reactions (e.g. in the photosynthetic electrontransfer chain) or the dark-reactions (e.g. in the Calvincycle) of photosynthesis. Light saturated photosyn the -sis is often limited by the dark-reactions (Sukenik et al.1987). However, when Pchl

m Chl < PFmaxCF, the light reac-

tions will be limiting. Control of the light-saturatedphotosynthesis is likely to be transferred from thedark-reactions to the light-reactions as a consequenceof photoinhibition. Falkowski (1981) calculated a theo-retical maximum value for Pchl

max of 24 gC (gChl)−1 h−1 forambient water temperatures of 25°C. Given that photo-synthetic electron transfer reactions have a Q10 ofabout 2 (Raven & Geider 1988), the value can be scaledfor environmental temperature: so at 15°C it would be12, whereas at 5°C it would be 6 gC (gChl)−1 h−1. In ourmodel we used a value of Pchl

max = 12 gC (gChl)−1 h−1.The parametrisation of the effect of photoinhibition

of the light saturated photosynthesis in Eq. (9) is basedon the observation that phytoplankton possess an ‘excess’ PSII capacity (Behrenfeld et al. 1998). The firstpart in Eq. (9) is identical to Eq. (15) from Ross & Gei-der (2009) and describes the light saturated photo -synthesis in the absence of photoinhibition. Hence, ini-tially, P cell

max will not be affected as the product PFmaxCF

will be lower than Pchlm Chl. Once photo inhibition

increases, however, the proportion of functional RCII,ϑ, will decrease, as will also Pchl

m (Eq. 10). This para-metrisation is consistent with the observation that the

quantum efficiency is decreased by damage to PS II(Öquist et al. 1992), and Eqs. (8) & (9) are loosely basedon Fig. 1 of Behrenfeld et al. (1998).

Quantification of errors. For the analyses we focuson comparing the Chl-derived growth rate from fixed-depth bottle incubations, μincub

Chl , with its carbon-derivedequivalent, μC

incub, and the carbon-based growth rate inthe freely mixed water column, 3μC

mixed, which we willrefer to as the in situ growth rate. In accordance withthe interests of experimentalists, we consider thedepth-averaged carbon growth rate in the SML, whichis denoted by the overbar in 3μC

mixed. The 2 differenterrors to be considered are thus ε1, the error due tousing Chl as a proxy for measuring growth, and ε2, theerror due to arresting the cells at a single fixed depthduring the incubation:

(11)

where the growth rates are calculated from:

(12)

with T = t1 − t0 being the incubation period (24 h).Simulating the yo-yo approach. One question which

may be interesting for experimentalists is whetherchanges in the experimental protocol, e.g. by oscillat-ing the incubation bottles vertically through the SML,can mitigate some of errors observed in fixed depthincubations. In order to address this question we usedthe same model setup as before, with the exceptionthat the cells are not arrested at fixed depths duringthe 24 h incubation period, but the cells within a partic-ular depth bin are collectively moved up and downthrough the water column at a pre-determined velocitywyo-yo. The choice of wyo-yo is to some extent arbitraryand was based on the standard deviation (SD) of therandom walk which increases with time as (Ross &Sharples 2004):

(13)

To obtain some estimate of Σ at longer time scales weevaluated Eq. (13) for t = 1 h. This yields the SD inmetres of a distribution of particles that was originallylocated at a point source z = 0. In order to convert thisinto a velocity we used

(14)

For the example of 10−2 m2 s−1 this yields wyo-yo ≈4.6 m h−1. The scaling factor of 0.55 in Eq. (14) resulted

= 1 ecellmaxcell max

cellP P I Chl PChl

−⎡⎣ ⎤⎦− α( )

PP C

P Chl

F F

mchlmax

cell maxMIN=⎧⎨⎪

⎩⎪

P Pmchl

axchl= m ϑ

α φ φ ϑ α ϑChlm

Chl Chl Chla a= = =max max

ε μ μμ1( ) =( ) ( )

| ( ) |100%z

z zz

Chl C

C

incub incub

incub

a

− ⋅

nndincub mixed

mixedε μ μ

μ2( ) =

( )100%z

zChl C

C

− ⋅

μ

μ

X

Xz

zX z tX z t T

X z t

( ) =( , )( , )

1

=( , )

1

0

1=0

ln

ln

⎛⎝

⎞⎠

−and

HH

z

HX z t T

∑∑−

⎝⎜⎜

⎠⎟⎟( , )

1

0=0

Σ( ) = 2t K tm

wt

yo-yoh

1 h= 0.55

( =1 )Σ

19

Page 8: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

from tests with a range of environmental conditions.Particularly in experiments where the diffusivity waslow, we noticed that without the scaling factor, Eq. (14)would overestimate the vertical velocities, yielding avertically too homogeneous growth rate compared tothe mixed conditions. The choice of this scaling factoris based on a best fit analysis for a range of experi-ments to minimise the errors in the shape and magni-tude of the growth rates. For water columns where theturbulence is high, an overestimation of wyo−yo is nottoo critical, as the growth rate is already homogeneouswith depth and the magnitude of the growth rate is notaffected by an overestimation.

RESULTS

While an Eulerian model always has uniform physi-ologies in each depth bin as it implicitly averages overthe individuals, the Lagrangian counterpart can pro-vide much more detailed information, including theamount of heterogeneity within a 1 m depth elementdue to cells with different light histories having beenmixed together by turbulence. When the cells arefreely mixing, weak turbulence (Fig. 3a) leads to a dis-tribution of Chl:C that is rather heterogeneous, both

vertically and also within a depth bin. At high turbu-lence intensities (Fig. 3b), Chl:C is more uniform, bothwithin and between depth bins, as the cells do not haveenough time to acclimate to any particular light level.While Fig. 3a provides a good example of a surfacemixing layer that is not mixed in the biological sense,Fig. 3b shows a surface layer that is not only mixing butalso mixed, with little vertical heterogeneity in thephysiology.

When the cells are kept at fixed depths to simulatethe bottle incubation, their residence time at each lightlevel is greatly increased, allowing them to acclimate tothe ambient light. As a result, the vertical heterogeneityin Chl:C is increased, while only a small heterogeneityis maintained within each depth bin (Fig. 3c,d).

Growth rates

Throughout the SML, but particularly near the sur-face, the 2 growth rates, μincub

Chl and μCincub, differ con -

siderably from one another (Fig. 4). Near the surfaceμincub

Chl is lower than μCincub, due to photoacclimation, and

ε1 is negative there (Fig. 4b,d,f,h). In samples incu-bated at high light, Chl synthesis is reduced while carbon uptake remains high, which causes a shift in

Mar Ecol Prog Ser 435: 13–31, 201120

Fig. 3. Relative abundance of different Chl:C ratios, θChl [gChl (gC)−1] for different depths in an optically shallow mixing layer(ξSML = 2.7), using 2 different mixing time scales, τm (cf. Eq. 5). The presence of many colours at a particular depth indicates a greatphysiological heterogeneity of the phytoplankton population in terms of their Chl:C ratio, while differences in the colour distrib-ution between the surface and bottom indicate vertical heterogeneity (due to insufficient mixing). (a,b) distribution at 17:30 h onthe day prior to the 24 h incubation, i.e. where the cells are still allowed to move freely; (c,d) distribution at 17:30 h on the day of

the 24 h incubation during which the cells had been kept at fixed depths

Pre-incubation τm = 100 h

Dep

th (m

)

0 20 40 60 80 100−6

−5

−4

−3

−2

−1

0

θChl

0.020.0250.030.0350.040.0450.050.0550.06

2.7

2.25

1.8

1.35

0.9

0.45

0

Op

tical dep

th ξ

(b) τm= 1 h

τm = 100 h τm= 1 h

Dep

th (m

)

0 20 40 60 80 100−6

−5

−4

−3

−2

−1

0

2.7

2.25

1.8

1.35

0.9

0.45

0

Op

tical dep

th ξ

Post-incubation

Dep

th (m

)

Cumulative % cells0 20 40 60 80 100

−6

−5

−4

−3

−2

−1

0

2.7

2.25

1.8

1.35

0.9

0.45

0

Op

tical dep

th ξ

(d)

Dep

th (m

)

Cumulative % cells0 20 40 60 80 100

−6

−5

−4

−3

−2

−1

0

2.7

2.25

1.8

1.35

0.9

0.45

0

Op

tical dep

th ξ

(c)

(a)

Page 9: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth

the distribution of Chl:C ratios (Fig. 5). The same argu-ment can be used to explain why ε1 turns positive atgreater depths where low light conditions lead to anincreased Chl synthesis in combination with a reducedcarbon uptake. For ε2 we find the opposite trend,with μincub

Chl overestimating in situ growth (3μCmixed) near

the surface and underestimating it at depth. This is

trivial, as it reflects the different light availabilities atdifferent depths.

Although the errors at the individual depths can beconsiderable, reaching 50 to 100% (Fig. 4d), the differ-ences between the SML-integrated growth rates 3μremain small (Table 2). Particularly for ξ 4, the invitro growth rate, μincub

Chl , significantly underestimates

21

0 0.5 1

0

1

2

3

4

5

(a) H = 6 m, τm= 100 h, ξSML

= 5.4

(e) H = 60 m, τm= 10000h, ξSML

= 5.4

0(c) H = 6 m, τm = 1 h, ξ

SML= 5.4

(g) H = 6 m, τm = 100 h, ξSML

= 2.7

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(b) Errors ε1 and ε2

ε1

ε2

0 0.5 1

1

2

3

4

5

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(d) Errors ε1 and ε2

ε1

ε2

0 0.5 1

0

1

2

3

4

5

Op

tical

dep

th ξ

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(f) Errors ε1 and ε2

ε1

ε2

0 0.5 1

0

0.5

1

1.5

2

2.5

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

0.5

1

1.5

2

2.5

(h) Errors ε1 and ε2

ε1

ε2

0 0.5 1

0

1

2

3

4

5

μ (d–1)

(i) as (a) but triple kd and Im = 850 W m

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(%)

(j) Errors ε1 and ε–2 –2

ε1

ε2

0 0.5 1

0

1

2

3

4

5

μ (d )

(k) as (c) but triple kd and Im = 850 W m

μincubC

μincubChl

μincubcells

μmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(%)

(l) Errors ε1 and ε2

ε1

–1

Fig. 4. Growth rates and associated errors for different scenarios. The vertical resolutions for the output were 1 m and 3 m for theshallow (H = 6 m) and deep (H = 60 m) cases, respectively. Panels (i) to (l) show the effect of maximising photoinhibition bytripling the damage probability kd and increasing the noon irradiance to 850 W m−2. The dashed vertical lines show the value for3μC

mixed which we use as representative in situ SML growth rate, i.e. the ‘truth’ in respect to which we calculate ε2. The numeric values for 3μC

mixed and the other depth-integrated growth rates are shown in Table 2

Page 10: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 2011

the in situ carbon based growth rate 3μCmixed. By compar-

ing the different growth rates for a range of environ-mental scenarios (Fig. 4) we generally find that 3μC

mixed ≈3μChl

mixed (Table 2).For completeness, we also show the growth rate cal-

culated from the increase in cell numbers, μcell. In gen-

eral, μcell(z) differs from the carbon-based equivalent,and also shows more variability in the vertical. Due toits dependence on the cell cycle it is a slightly coarsermeasure of growth rate, as the timing of cell divisionrelative to the end of the incubation period could sig-nificantly alter the results. Nevertheless, the depth-

22

Dep

th (m

)

0 20 40 60 80 100−6

−5

−4

−3

−2

−1

0

5.4

4.5

3.6

2.7

1.8

0.9

0

Op

tical dep

th ξ

Km = 10−2 m2 s−1

Km = 10−2 m2 s−1

Km = 10−4 m2 s−1

Km = 10−4 m2 s−1

Dep

th (m

)

0 20 40 60 80 100−60

−50

−40

−30

−20

−10

0

θChl

0.020.0250.030.0350.040.0450.050.0550.060.0650.070.0750.08

5.4

4.5

3.6

2.7

1.8

0.9

0

Op

tical dep

th ξ

Dep

th (m

)

% cells0 20 40 60 80 100

−6

−5

−4

−3

−2

−1

0

5.4

4.5

3.6

2.7

1.8

0.9

0

Op

tical dep

th ξ

Dep

th (m

)

% cells0 20 40 60 80 100

−60

−50

−40

−30

−20

−10

0

5.4

4.5

3.6

2.7

1.8

0.9

0

Op

tical dep

th ξ

(a) (b)

(c) (d)

Pre-incubation

Post-incubation

Fig. 5. Distributions of Chl:C for 2 different mixed layers which represent (a) a coastal setting (from Fig. 4a) and (b) an open oceansetting (cf. Fig. 3 for explanation of the colours and bars). The vertical resolution in (a) and (c) is 1 m and in (b) and (d) it is 3 m.Although both SMLs differ in depth, attenuation coefficient, and turbulent mixing intensity, they have the same values for the

mixing time scale (τm = 100 h) and optical depth (ξSML = 5.4). As a result, both SMLs show equivalent distributions of Chl:C

Fig. 4 Environmental parameters Growth rates (d–1)

H Km k τm ξSML3μChl

incub 3μChlmixed 3μC

incub 3μCmixed 3μcell

incub 3μcellmixed

a 6 10−4 0.9 100 5.4 0.62 0.60 0.68 0.59 0.61 0.60c 6 10−2 0.9 1 5.4 0.49 0.53 0.62 0.54 0.58 0.44e 60 10−4 0.09 104 5.4 0.73 0.70 0.78 0.70 0.70 0.68g 6 10−4 0.45 100 2.7 0.84 0.83 0.86 0.83 0.77 0.78

Growth rates for a deep SML with same values of τm and ξSML as in Fig. 4a– 60 10−2 0.09 100 5.4 0.61 0.60 0.68 0.59 0.61 0.60

Using triple kd and Im = 850 W m−2

i 6 10−4 0.9 100 5.4 0.51 0.48 0.51 0.48 0.50 0.48k 6 10−2 0.9 1 5.4 0.49 0.52 0.52 0.52 0.51 0.69

Table 2. Depth-averaged growth rates for the scenarios shown in Fig. 4. The last 2 rows have triple kd and Im = 850 W m−2. τm is in hours; units for the remaining environmental parameters can be found in Table 1

Page 11: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 23

integrated rates typically give an acceptable approxi-mation of the in situ values (Table 2).

Sensitivity to optical depth and mixing time scale

A reduction in the mixing time scale τm produces a more homogeneous water column (Figs. 3b & 6b),which in turn amplifies the differences between thefreely mixing and static incubation, yielding slightlylarger errors ε1 and ε2. An increase in the mixing timescale, on the other hand, means that the cells have ahigher residence time at each depth, which allowsthem to adapt better to the ambient light level. In sucha low turbulence scenario, arresting the cells in placeduring the fixed-depth bottle incubation thus repre-sents only a small change to the freely mixing condi-tions (compare the mixed and incubated conditions inFig. 6c) and the error ε1 is reduced (Fig. 4e,f). However,

ε2 remains large, which is to be expected given the ver-tical heterogeneity in μincub

Chl . The fact that the cells arebetter adapted to their light environment and thus lessinhibited at greater depths (Fig. 6c vs b) also results inhigher depth-integrated growth rates (3μChl

mixed and3μC

mixed, which reach 0.7 d−1 for τm = 104 h, compared to0.6 d−1 for τm = 100 h, and 0.54 for τm = 1 h (Table 2).With everything else left unchanged, a reduction inturbulence thus results in a water column that is moreproductive.

If we use the same mixing time scale as in Fig. 4(a)but halve the attenuation coefficient, thereby halvingξSML, the result is a significant reduction in the errors,in particular, ε2 (Fig. 4g,h). Having made the SML moretransparent to light means that the base of the SMLreceives relatively high light intensities (the location ofthe 1% light level for the simulations in Fig. 4(a–f) isnow the location of the 10% light level). The weakerlight gradient produces a weaker vertical gradient in

0.9

0.90.8

0.7

0.6

0.50.4

0 6 12 18 24

4

3

2

1

0.9

0.9

0.8

0.7 0.60.5

0 6 12 18 24

3.5

2.5

1.5

4.5

0.9

0.90.8

0.70.6

0.50.4

0 6 12 18 24

4

3

2

1

4

3

2

1

Incubated Mixed Incubated Mixed

0.9

0.90.8

0.7

0.7

0.8

0 6 12 18 24

3.5

2.5

1.5

4.5

Op

tical

dep

th ξ

0.9

0.90.9

0.80.7

0.6

0.5

0.40.3

0 6 12 18 245

4

3

2

1

0.9

0.9

0.80.7 0.6

0.5 0.4

0 6 12 18 245

4

3

2

1

0.9

0.9

0.80.7

0.6

0.5

0.4

0.3

0 6 12 18 24

0.9 0.9

0.7

0.8

0.6

0.5

0.4

0.3

0 6 12 18 24

2

1.5

1

0.5

0.8

0.9

0.9

0.8

0.7

0.60.5

0.4

0.3

0.20.1

0 6 12 18 24

4

3

2

1

0.8

0.9

0.9

0.80.7

0.6

0.5

0.4

0.30.2

0 6 12 18 24

3.5

2.5

1.5

4.5

0.8

0.9

0.8

0.7

0.6

0.5

0.40.3

0.20.1

0 6 12 18 24

4

3

2

1

Time of day (h)

0.8

0.80.6

0.70.5

0.40.3

0.3

0.4

0.50.6

0.7

0 6 12 18 24

3.5

2.5

1.5

4.5

(e) as in (a) but triple kd, Im= 850 W m−2 (f) as in (b) but triple kd, Im= 850 W m−2

(d) τm = 100 h, ξSML

= 2.7(c) τm = 10000 h, ξSML

= 5.4

(a) τm =100 h, ξSML

= 5.4 (b) τm =1 h, ξSML

= 5.4

Fig. 6. Value of the photoinhibition parameter ϑ for the simulations from Fig. 4

Page 12: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 2011

Chl:C and a smaller heterogeneity within a particulardepth bin (cf. Fig. 3a). As a result, al though the error ε1

is still of the order of O(20–30%), the more important ε2

has almost disappeared, which means that μincubChl is now

a much better approximation of the in situ growth rate3μC

mixed, irrespective of the depth at which the incubationtakes place. Despite photoinhibition extending to thebase of the SML (Fig. 6d), the high light availabilitythroughout the SML results in higher overall growthrates (Table 2).

The results in Fig. 4(a,b) were obtained for a shal-low (H = 6 m) and relatively turbid (k = 0.9 m−1) SMLrepresentative of an estuary or coastal lagoon. Theoptical depth of this SML was ξSML = 5.4 and the mix-ing time scale τm = 100 h (Eq. 5), which renders thewater column vertically heterogeneous, both in termsof Chl:C (Fig. 5a) and also the inhibition ϑ (Fig. 6a).We found that experiments with identical values ofξSML and τm delivered the same trends and magni-tudes in the vertical Chl:C distribution, the resultinggrowth rates and associated errors. E.g. moving fromthe estuarine setting in Fig. 4(a,b) to an open oceansetting with H = 60 m, the same values for ξSML andτm can be maintained by reducing the PAR attenua-tion to k = 0.09 m−1 and increasing the eddy diffusiv-ity to Km = 10−2 m2 s−1 (see Table 2). The result is anidentical distribution of Chl:C (Fig. 5) and also identi-cal distributions and magnitudes for the growth ratesand associated errors (not shown). Compared to thediagrams in Fig. 3, the cells achieve higher values ofChl:C in Fig. 5, as the water column was more turbidin this experiment and thus the lower part of theSML was more light limited.

Photoinhibition

The simulations in Fig. 4a–h used a moderate valuefor the photodamage probability, kd, which yields min-imum values for ϑ of about 0.4 in a freely mixing watercolumn with τm = 100 h (Fig. 6a). The amount of inhibi-tion generally decreases with increasing turbulence(Fig. 6a,b) but extends deeper into the SML. In aneffort to maximise the photodamage, we repeated thesimulations from Fig. 4a,c with a 3 times higher dam-age probability kd and an increased noon irradiance Im.The result is a higher photoinhibition (Fig. 6e,f) whichleads to a depression in the incubated growth rates μC

incub (Fig. 4i,k). In the freely mixing water column, theeffect on 3μC

mixed is only noticeable in a moderately tur-bulent water column (Fig. 4a,i) while in a highly turbu-lent water column, 3μC

mixed is almost un affected(Fig. 4c,k). Because the carbon based growth rate ismore affected by the higher inhibition and due to thelower depth-averaged growth rates, both ε1 and ε2 arereduced near the surface and at depth except wherethe growth rate becomes negative (Fig. 4j,l).

To generalise our findings, we repeated the aboveexperiments for a wide range of physical environmentsand found that the vertical distribution and magnitudeof errors depends on the 2 parameters ξSML and τm

(Fig. 7a). ε1 and ε2 were generally small if the SML waswell lit, i.e. ξSML 3. In these cases μincub

Chl delivered agood approximation of the growth in the freely mixingwater column 3μC

mixed. ε1 was also small if the water col-umn was very stable with low turbulence intensities(τm is very large) (Region I in Fig. 7a). For ξ 4, wefound that ε2 ≈ −100% irrespective of τm. As the SML

24

1001011021031040

2

4

6

8

10

12

14

16

18

τm (h)

ξ SM

L

(a)

−100 −75 −50 −25 0 25 50 75 1001

0.5

0(b)

Typical error (%)

Frac

tion

of S

ML

dep

th

ε2

ε1

ε2 minimal

ε1 minimal

Region II: transitional

Region III:large errors >100% possible.In the upper SML:ε1<0, O(30−50%)

ε2>0, O(20−50%)

In the lower SML:ε1>0, O(50−100%)

ε2<0, O(50−100%)

ε2(ξ≥4) ≈ −100%

Region I: Small or negligible ε1 (typically < 15%)but ε2 can be large (up to 100%) if ξSML ≥ 3

Fig. 7. Summary of the general trends observed in the numerical experiments. The double-ended arrows in (b) indicate theranges where the respective errors are minimal. If ξSML 6, ε2 is minimal near the mid-SML and ε1 slightly above. For more turbid

mixing layers these locations are shifted upward

Page 13: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 25

becomes more turbid and also more turbulent (RegionIII in Fig. 7a) both errors can become very large. Nearthe surface μincub

Chl typically underestimates the carbonequivalent μC

incub (unless the inhibition is very high)while overestimating the in situ growth rate 3μC

mixed by10 to 50%, depending on the environmental condi-tions. ε1 is therefore typically negative near the surfacewhile ε2 is positive (Fig. 7b). The opposite is true for thelower SML, where the signs of ε1 and ε2 change and themagnitudes approximately double. The region wherewe found the errors to be the smallest is near the mid-SML (as long as ξmid-SML 4).

Sensitivity to the incubation time

The results shown in Fig. 4 & Table 2 were obtainedfor an incubation time of 24 h extending from midnighttill midnight on the last day of the simulations. Onemight intuitively assume that errors could be reducedby shortening the incubation time, i.e. the difference tothe in situ conditions would be smaller if the cells wereonly stranded at fixed depths for a few hours ratherthan the entire day. We found that the opposite holds

true, at least for low to moderate levels of inhibition. Ifthe bottles are incubated for only 4 h from 10:00 to14:00 h for instance (cf. Marra 1978), both the signand magnitude of the errors change significantly(Fig. 8a–d). Compared to the 24 h incubations, thedepth averaged Chl-based in vitro growth rate 3μChl

incub isnow only about half the in situ value 3μC

mixed. By shorten-ing the incubation time, and carrying out the incuba-tion symmetrically around noon, we have biased ourmeasurements be cause the time of maximal irradianceis also the time of maximal carbon uptake, while theChl:C ratio generally remains low. As a consequence,μincub

Chl significantly underestimates 3μCmixed throughout

the water column. While ε1 maintains a similar distrib-ution (negative near the surface and positive at depth),the more important ε2 has turned negative throughoutthe water column. If we induce a higher photoin -hibition in the cells by tripling the damage probabilitykd and raising the noon irradiance to Im = 850 W m−2

(Fig. 8e–h), the errors are reduced and μincubChl now pro-

vides a good estimate of 3μCmixed in mid SML (ε2 is very

small there). However, μChlincub still underestimates its

carbon equivalent by 10 to 15%.

0 0.05 0.1

0

1

2

3

4

5

(a) H = 6 m, τm= 100 h, ξSML= 5.4

µincubC = 0.058

µmixedC = 0.055

µincubChl = 0.028

µmixedChl = 0.030

µincubC

µincubChl

µmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(b) Errors ε1 and ε2

ε1

ε2

0 0.05 0.1

0

1

2

3

4

5

(c) H = 6 m, τm= 1 h, ξSML 5.4

µincubC = 0.043

µmixedC = 0.041

µincubChl = 0.018

µmixedChl = 0.022

µincubC

µincubChl

µmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(d) Errors ε1 and ε2

ε1

ε2

0 0.05 0.1

0

1

2

3

4

5

µ (h−1)

Op

tical

dep

th ξ

(e) as in (a) but triple kd and Im=850Wm−2

µincubC = 0.037

µmixedC = 0.037

µincubChl = 0.033

µmixedChl = 0.034

µincubC

µincubChl

µmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(%)

(f) Errors ε1 and ε2

ε1

ε2

0 0.05 0.1

0

1

2

3

4

5

µ (h−1)

(g) as in (c) but triple kd and Im=850Wm−2

µincubC = 0.033

µmixedC = 0.039

µincubChl = 0.032

µmixedChl = 0.038

µincubC

µincubChl

µmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(%)

(h) Errors ε1 and ε2

ε1

ε2

Fig. 8. Corresponding results to panels (a–d) and (i–l) in Fig. 4, except that the incubation time was not 24 h but only 4 h from10:00 to 14:00 h. The cell-based growth rate is not very meaningful for such short incubation times and has therefore been

omitted from this plot

Page 14: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 2011

Yo-yo techniques

To test whether a yo-yo approach, in which the incu-bation bottles are vertically moved up and downthrough the water column, could reduce the associatederrors, we applied this approach to the scenario fromFig. 4c, which falls into Region III in Fig. 7a and there-fore delivers relatively large errors. The result is a dra-matic reduction in both types of errors (Fig. 9). By oscil-lating the sampling bottles up and down we were ableto maintain them in balanced growth and obtain fairlyaccurate estimates of the growth rate in the freely mix-ing water column.

In order to effectively reduce the errors, it is crucialto choose an appropriate value for the oscillation veloc-ity wyo-yo. If the mixing time scale is low (in Fig. 9 τm =1 h) an overestimation of wyo-yo is not too critical as thehigh turbulence renders the SML fairly homogeneous(Figs. 3b & 6b). However, an underestimation of wyo-yo

can produce significant errors (not shown).

DISCUSSION

Accurate estimates of phytoplankton in situ growthrates are important for many applications, rangingfrom global climate predictions to local management ofwater quality. Using a modelling approach we ex -plored the magnitude of potential errors associatedwith the commonly used fixed-depth bottle incubationmethod. While there are many different sources oferror associated with fixed-depth bottle incubations(e.g. Furnas 2002), we focused our efforts on the quan-tification of 2 particular types of errors:

(1) The first error (ε1) accounts for the differencebetween the commonly used Chl-derived growth rate

μincubChl (z) and its carbon-derived equivalent μC

incub(z).The discrepancy is related to the fact that the cells con-tinuously photoacclimate and are typically not in bal-anced growth.

(2) The second error (ε2) accounts for the effect ofkeeping the incubations at fixed depths (in stable lightconditions) and assuming that the obtained growthrate corresponds to the entire mixing layer, i.e. itdescribes the discrepancy between 3μChl

incub(z) and thedepth integrated growth rate achieved in the freelymixing water column, 3μC

mixed which, for the purposes ofthis study, we considered to be the ‘true’ in situ growthrate.

We showed that depending on the depth at whichthe incubation takes place, both errors can be of thesame order of magnitude as the growth rate itself.Some methods exist which can help alleviate (Gutiér-rez-Rodríguez et al. 2010) or even completely avoid(Calbet & Landry 2004) ε1, but they cannot remedy thepotentially more important ε2. Even if we reduce ε1

completely, there is still a considerable error betweenμC

incub and 3μCmixed (Fig. 4). ε2 can be reduced by either

oscillating the bottle through the SML (Fig. 9) or byusing more than one depth for the incubation (Table 2).

Our results have shown that both ε1 and ε2 can berather large for a wide range of environmental condi-tions. In all tested scenarios (including the highly tur-bulent ones), the best depth to carry out the fixed-depth incubation (using only a single depth) was nearthe centre of the SML, if— and only if — the opticaldepth of the mid-SML, ξmid-SML, is no larger than 4. Ifξmid-SML > 4 the incubation should be carried out nearξ = 3.5 to 4. If most of the SML lies within the euphoticzone, and if the turbulence is low to moderate (i.e. ifthe physiology throughout the SML is vertically het-erogeneous, τm ≥ 100 h) we found that a good approachto reduce the errors is to carry out the incubationexperiment not only at a single depth but at a series ofdepths throughout the SML and then calculate thegrowth rate from this depth integrated in crease in Chl,which yields good agreement be tween 3μC

mixed and 3μChlincub

(Fig. 4 & Table 2). Our results were quite insensitive tosmall changes in the growth or photoacclimation para-meters used in the model. Only for very high photo-damage probabilities (kd) can the errors ε1 and ε2 bereduced near the surface, as uptake of carbon is sup-pressed due to lower photosynthetic efficiencies and ε2

can change sign (Fig. 4i–l). For the range of parametervalues we tested, we did not find any significant effecton ε2 at greater depths, however.

If the SML is more turbulent (e.g. if the distribution ofChl:C ratios is fairly constant with depth, τm ≤ 10 h), μincub

Chl can significantly differ from 3μCmixed (Fig. 4c,d) and

a superior approach to reduce ε2 is to oscillate the incu-bation bottle through the layer (Fig. 9). Although our

26

0 0.5 1

0

1

2

3

4

5

µ (d−1)

Op

tical

dep

th ξ

(a) H = 6 m, τm=1h, ξSML= 5.4

µincubC = 0.57

µmixedC = 0.54

µincubChl = 0.57

µmixedChl = 0.53

µincubcell = 0.46

µmixedcell = 0.44

µincubC

µincubChl

µincubcells

µmixedC

−100 −50 0 50 100

0

1

2

3

4

5

(%)

(b) Errors ε1 and ε2

ε1

ε2

Fig. 9. Applying the yo-yo-approach to the scenario from Fig.4c. Instead of incubating bottles at fixed depths, they werecycled vertically through the SML using wyo−yo = 4.6 m h−1

(cf. Eq. 14)

Page 15: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 27

modelling approach showed that this yo-yo methoddelivers far smaller errors, the problem remains thatwe have no reliable method of estimating the oscilla-tion velocity, the vertical amplitude, or the time courseof the oscillations (e.g. linear velocity vs. sinusoidal vs.a more random movement). Eq. (14) represents anempirical relationship which produced results thatwere quantitatively in good agreement with thoseobtained in normally mixed conditions but it requiresprior knowledge of the turbulence diffusivity in thewater column, which is often unknown. A wrongchoice may not increase accuracy.

In general, the conclusions are:(1) The magnitude of ε1 decreases with increasing

mixing time scale τm (Eq. 5) (Fig. 7a).(2) The magnitude of both errors increases with

increasing turbidity of the water column (ξSML, Eq. 5)(Fig. 7a).

(3) If the mixing layer is optically shallow and rela-tively stable (low turbulence), then fixed-depth incu-bations at optical depths 4 should yield good esti-mates (errors 20%) of the growth rate in the freelymixing water column (Fig. 7).

(4) If the water column is more turbid and/or moreturbulent, the best region to use fixed-depth bottleincubations (error 30%) is near the centre of the sur-face mixing layer, provided that this region has an opti-cal depth of 4 (Fig. 7b).

(5) Reducing the incubation time from 24 to 4 h doesnot necessarily lead to improved accuracy. For low tomoderate photoinhibition, a shorter incubation time,especially if located symmetrically about noon, willlead to a significant increase in the associated errors, inparticular ε2 (Fig. 8).

(6) One additional way to reduce the errors in highlyturbulent and turbid mixing layers is to vertically oscil-late the incubation bottles through the water column(yo-yo-method) which, by choosing an appropriate yo-yo velocity and amplitude, can reduce the errors to 15% throughout the water column (Fig. 9b).

(7) We often found that 3μChlincub ≈ 3μC

mixed (Fig. 4).Although we are aware that logistic constraints oftendo not permit this, we found that the best way toreduce ε2 to negligible levels is to carry out the incuba-tion experiments at several depths which cover the dif-ferent light environments within the mixing layer.

Accuracy of yo-yo incubations

The yo-yo method was a viable alternative only if thewater column was turbulent enough so that the cellsshowed a fairly homogeneous physiology throughoutthe layer, and most of the mixing layer was within theeuphotic zone. If the layer is not very turbulent, the

success of this method is very sensitive to the choice ofthe depth at which the sample is taken and at whichthe oscillation is started. E.g. if we were to use theyo-yo method in a shallow (H = 6 m) and turbid (k =0.9 m−1) water column with an eddy diffusivity of10−4 m2 s−1 (i.e. wyo-yo = 0.46 m h−1) and if we start theincubation at midnight, then cells that were originallynear the surface would already be at an optical depthof about 1.5 by the time the sun rises. By noon the cen-tre of the depth bin that was initially at the surfacewould be at the base of the SML and only by the timethe sun was setting would the cells have returned tothe surface. Hence in such a scenario the cells thatstarted near the surface would have spent much of thedaylight hours at great optical depths where the lightavailability is low (and vice versa for cells that start atgreater depths). If we start the yo-yo incubation at sun-rise or noon, instead of midnight, the effect on the cellsis quite different, hence the high sensitivity of theresults to the timing of the incubation. Since the errorassociated with fixed-depth incubations in watercolumns that are optically shallow is quite low(Fig. 4g,h), one could argue that in such environmentsit is not necessary to use the yo-yo technique as thefixed-depth incubations might already yield suffi-ciently accurate results. Even if the water column isslightly more turbid, it may be worthwhile consideringwhether the additional effort of installing an oscillationmechanism might not be better spent on suspending aseries of bottles at different depths that are representa-tive of the entire SML and which may deliver superiorresults (3μC

mixed ≈ 3μChlincub) without the added complication

of having to choose the parameters for the yo-yo oscil-lation.

All our results regarding the yo-yo technique shouldbe considered as preliminary. With our implementa-tion of the method, we were able to show that oscillat-ing the incubation bottles can have a positive effect incertain situations, but not in all. Further investigationsinto the yo-yo technique are necessary, however, sothat more detailed and quantitatively more sound rec-ommendations for an adequate sampling protocol canbe made.

Turbulence parameterisation

In the present study we chose a rather simple para-meterisation of turbulence, namely the eddy diffusivityapproach. As an additional simplification we decidedto prescribe the diffusivity profiles using analyticalequations (Eq. 4) rather than obtaining the values froma k-ε model that calculates the diffusivity dynamicallyfrom the external forcing and ambient water columnstability (e.g. Ross & Sharples 2007). As a result, our

Page 16: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 201128

diffusivity profiles are fairly constant throughout theSML (Fig. 1) and do not change with time. The discreteoutput of eddy diffusivities obtained from a k-ε modeldoes not fulfill the necessary criteria for use with therandom walk equation (Eq. 2). As a result, the discretedata need to be smoothed (e.g. using cubic splines) inorder to achieve the necessary continuity and differen-tiability in the profiles (Ross & Sharples 2004). As thediffusivity gradients are typically large near the SMLboundaries, the number of nodes for the spline alsoneeds to be quite large, yielding a high number ofcoefficients to evaluate a large number of polynomialsfor just 1 single profile which would change every 1 to2 min. In tests with several such profiles we found thatthe most important aspects of the diffusivity profile thatneed to be captured are the gradients (towards the sur-face and bottom) and the mid-SML amplitude. Anyother structure in the profiles is redundant, as it doesnot have any significant effect on the biology. Theoverall results presented in this study were thus notaffected by the simplified shape of the diffusivity pro-files used.

We also tested the effect of a time-varying diffusivityby examining 2 different scenarios: (1) a scenariowhere the mixing is dominated by periodic (M2) tidalturbulence, as would be the case in an estuary or shelfsea, and where the diffusivity varied between Km (dur-ing the ebb and flood flows) and 10% of Km (during thehigh and low water stand) (see for instance the bottommixed layer in Fig. 3c of Ross & Sharples 2007), and (2)a scenario where the intermittency was largely due tochanges in the surface wind forcing and where the dif-fusivity varied randomly between Km and 1% of Km. Inboth cases the impact on the errors was minimal anddid not warrant the inclusion of such effects in ouranalysis.

A third simplification in our experiments was that theSML depth was kept constant during the 5 d experi-ments and did not show any periodic deepening atnight, even for the scenarios where we used convectiveoverturning (cf. Woods & Onken 1982, Barkmann &Woods 1996). We therefore tested a scenario in whichthe SML depth oscillated between –75 m and –45 m (deepening at night associated with an increasein Km and shallowing during the day) with the 1% lightlevel being at about –51 m, and found that the overalleffect was a slight increase in ε2 in the top half of theSML [O(20%)] but no significant effect on ε1 or oneither error in the lower SML. We therefore expect ourresults to hold up also in a more dynamic physicalsetup.

While our results are obviously affected not only byour treatment of turbulence, but also by our somewhatidealised (and therefore tractable) treatment of bothphotoacclimation and photoinhibition, it is neverthe-

less the first study of its kind to employ a relativelysophisticated individual based model, both for thephysics and the biology, to address some of the prob-lems associated with fixed-depth bottle incubations.Due to our inability to measure in situ growth ratesnon-intrusively in the real world, it could be arguedthat a model is in fact the only viable approach avail-able to us at present in order to undertake such a com-parison. One of the challenges facing both modellersand experimentalists is the possibility of interactionsbetween the processes that have different characteris-tic time scales, so that a more fully integrated modelmay need to be developed.

Acknowledgements. This study received funding from the European Union Seventh Framework Programme(FP7/2007–2013) Marie Curie Intra-European Fellowship forCareer Development under grant agreement no. 255396 forproject AQUALIGHT (www. aqualight. info). R.J.G.’s researchon photoacclimation is supported by the UK Natural Environ-ment Research Council (NE/ G003688/ 1). M.L.A. benefitedfrom a FPU doctoral scholarship from the Spanish Ministry ofScience and Innovation (MICINN). The authors also acknowl-edge financial support from the Spanish funded projectsSUMMER (CTM2008-03309/MAR), ANERIS (PIF08-015), andECOALFACS (CTM2009-09581; a project endorsed by the‘Harmful Algal Blooms in Stratified Systems’ CRP of the GEO-HAB programme).

LITERATURE CITED

Anning T, MacIntyre HL, Pratt SM, Sammes PJ, Gibb S, Geider RJ (2000) Photoacclimation in the marine diatomSkeletonema costatum. Limnol Oceanogr 45:1807–1817

Barkmann W, Woods JD (1996) On using a Lagrangian modelto calibrate primary production determined from in vitroincubation measurements. J Plankton Res 18:767–788

Behrenfeld MJ (2010) Abandoning Sverdrup’s critical depthhypothesis on phytoplankton blooms. Ecology 91:977–989

Behrenfeld MJ, Prasil O, Kolber ZS, Babin M, Falkowski PG(1998) Compensatory changes in photosystem II electronturnover rates protect photosynthesis from photoinhibi-tion. Photosynth Res 58:259–268

Bienfang PK (1981) Sinking rates of heterogeneous, temper-ate phytoplankton populations. J Plankton Res 3:235–253

Bienfang PK, Takahashi M (1983) Ultraplankton growth ratesin a subtropical ecosystem. Mar Biol 76:213–218

Broekhuizen N (1999) Simulating motile algae using a mixedEulerian-Lagrangian approach: Does motility promotedinoflagellate persistence or co-existence with diatoms?J Plankton Res 21:1191–1216

Brown SL, Landry MR, Christensen S, Garrison D, GowingMM, Bidigare RR, Campbell L (2002) Microbial commu-nity dynamics and taxon-specific phytoplankton produc-tion in the Arabian Sea during the 1995 monsoon seasons.Deep Sea Res II 49:2345–2376

Calbet A, Landry MR (2004) Phytoplankton growth, micro-zooplankton grazing, and carbon cycling in marine sys-tems. Limnol Oceanogr 49:51–57

Cianelli D, D’Alcalá MR, Saggiomo V, Zambianchi E (2004)Coupling mixing and photophysiological response ofantarctic plankton: a Lagrangian approach. Antarct Sci 16:133–142

Page 17: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Eilers PHC, Peeters JCH (1988) A model for the relationshipbetween light intensity and the rate of photosynthesis inphytoplankton. Ecol Model 42:199–215

Eppley RW, Holmes RW, Strickland JDH (1967) Sinking ratesof marine phytoplankton measured with a fluorometer.J Exp Mar Biol Ecol 1:191–208

Falkowski PG (1981) Light-shade adaptation and assimilationnumbers. J Plankton Res 3:203–216

Falkowski PG (1983) Light-shade adaptation and verticalmixing of marine phytoplankton: a comparative fieldstudy. J Mar Res 41:215–237

Furnas M (1990) In situ growth rates of marine phytoplankton:Approaches to measurement, community and speciesgrowth rates. J Plankton Res 12:1117–1151

Furnas M (2002) Measuring the growth rates of phytoplank-ton in natural populations. In: Rao DVS, Rao S, GanapatiPN, Lafond EC, Humphrey GF (eds) Pelagic ecologymethodology. CRC Press, Boca Raton, FL, p 221–249

Furuya K, Takahashi M, Nemoto T (1986) Summer phytoplank-ton community structure and growth in a regional upwellingarea off Hachijo Island, Japan. J Exp Mar Biol Ecol 96:43–55

Gallegos CL, Platt T (1985) Vertical advection of phytoplank-ton and productivity estimates: a dimensional analysis.Mar Ecol Prog Ser 26:125–134

Geider RJ, MacIntyre HL, Kana TM (1998) A dynamic regula-tory model of phytoplanktonic acclimation to light, nutri-ents, and temperature. Limnol Oceanogr 43:679–694

Gocke K, Lenz J (2004) A new turbulence incubator for measuring primary production in non-stratified waters.J Plankton Res 26:357–369

Gutiérrez-Rodríguez A, Latasa M, Estrada M, Vidal M, Mar-rasé C (2010) Carbon fluxes through major phytoplanktongroups during the spring bloom and post-bloom in thenorthwestern Mediterranean Sea. Deep Sea Res I 57:486–500

Han BP (2002) A mechanistic model of algal photoinhibitioninduced by photodamage to photosystem-II. J Theor Biol214:519–527

Kagami M, Urabe J (2001) Phytoplankton growth rate as afunction of cell size: an experimental test in Lake Biwa.Limnology 2:111–117

Köhler J (1997) Measurement of in situ growth rates of phyto-plankton under conditions of simulated turbulence.J Plankton Res 19:849–862

Köhler J, Schmitt M, Krumbeck H, Kapfer M, Litchman E,Neale PJ (2001) Effects of UV on carbon assimilation ofphytoplankton in a mixed water column. Aquat Sci 63:294–304

Landry MR, Hassett RP (1982) Estimating the grazing impactof marine micro-zooplankton. Mar Biol 67:283–288

Landry MR, Kirshtein J, Constantinou J (1995) A refined dilu-tion technique for measuring the community grazingimpact of microzooplankton, with experimental tests inthe central equatorial Pacific. Mar Ecol Prog Ser 120:53–63

Latasa M, Landry MR, Schlüter L, Bidigare RR (1997) Pigmentspecific growth and grazing rates of phytoplankton in thecentral Equatorial Pacific. Limnol Oceanogr 42:289–298

Latasa M, Morn XAG, Scharek R, Estrada M (2005) Estimat-ing the carbon flux through main phytoplankton groups inthe northwestern Mediterranean. Limnol Oceanogr 50:1447–1458

Lewis MR, Cullen JJ, Platt T (1984a) Relationships betweenvertical mixing and photoadaptation of phytoplankton:similarity criteria. Mar Ecol Prog Ser 15:141–149

Lewis MR, Horne EPW, Cullen JJ, Oakey NS, Platt T (1984b)Turbulent motions may control phytoplankton photosyn-thesis in the upper ocean. Nature 311:49–50

Lizon F, Seuront L, Lagadeuc Y (1998) Photoadaptation andprimary production study in tidally mixed coastal watersusing a Lagrangian model. Mar Ecol Prog Ser 169:43–54

Long SP, Humphries S, Falkowski PG (1994) Photoinhibitionof photosynthesis in nature. Annu Rev Plant Physiol PlantMol Biol 45:633–662

MacIntyre HL, Kana TM, Geider RJ (2000) The effect of watermotion on short-term rates of photosynthesis by marinephytoplankton. Trends Plant Sci 5:12–17

MacIntyre HL, Kana TM, Anning T, Geider RJ (2002) Photo -acclimation of photosynthesis irradiance response curvesand photosynthetic pigments in microalgae and cyanobac-teria. J Phytol 38:17–38

Marra J (1978) Phytoplankton photosynthetic response to ver-tical movement in a mixed layer. Mar Biol 46:203–208

Moore CM, Suggett DJ, Hickman AE, Kim YN and others(2006) Phytoplankton photoacclimation and photoadapta-tion in response to environmental gradients in a shelf sea.Limnol Oceanogr 51:936–949

Nagai T, Yamazaki H, Kamykowski D (2003) A Lagrangianphotoresponse model coupled with 2nd-order turbulenceclosure. Mar Ecol Prog Ser 265:17–30

Öquist G, Chow WS, Anderson JM (1992) Photoinhibition ofphotosynthesis represents a mechanism for the long termregulation of photosystem II. Planta 186:450–460

Ptacnik R, Diehl S, Berger S (2003) Performance of sinkingand nonsinking phytoplankton taxa in a gradient of mix-ing depths. Limnol Oceanogr 48:1903–1912

Raven JA, Geider RJ (1988) Temperature and algal growth.New Phytol 110:441–461

Rhee GY, Gotham IJ (1981) The effect of environmental factors on phytoplankton growth: light and the interac-tions of light with nitrate limitation. Limnol Oceanogr 26:649–659

Ross ON, Geider RJ (2009) New cell-based model of photo-synthesis and photo-acclimation: accumulation and mobil-isation of energy reserves in phytoplankton. Mar EcolProg Ser 383:53–71

Ross ON, Sharples J (2004) Recipe for 1-d Lagrangian particletracking models in space-varying diffusivity. LimnolOceanogr Methods 2:289–302

Ross ON, Sharples J (2007) Phytoplankton motility and thecompetition for nutrients in the thermocline. Mar EcolProg Ser 347:21–38

Ross ON, Moore CM, Suggett DJ, MacIntyre HL, Geider RJ(2008) A model of photosynthesis and photo-protectionbased on reaction centre damage and repair. LimnolOceanogr 53:1835–1852

Ross ON, Geider RJ, Piera J (2011) Modelling the effect of ver-tical mixing on bottle incubations for determining in situphytoplankton dynamics. II. Primary production. Mar EcolProg Ser 435:33–45

Sharples J, Moore CM, Rippeth TR, Holligan PM, Hydes DJ,Fisher NR, Simpson J (2001) Phytoplankton distributionand survival in the thermocline. Limnol Oceanogr 46:486–496

Stolte W, Garcés E (2006) Ecological aspects of harmful algalin situ population growth rates. In: Granéli E, Turner JT(eds) Ecology of harmful algae, Vol. 189. Ecological Stud-ies. Springer-Verlag, Berlin, p 139–152

Suggett DJ, Moore CM, Hickman AE, Geider RJ (2009) Inter-pretation of fast repetition rate (FRR) fluorescence: signa-tures of phytoplankton community structure versus physi-ological state. Mar Ecol Prog Ser 376:1–19

Sukenik A, Bennett J, Falkowski P (1987) Light-saturatedphotosynthesis—limitation by electron transport or car-bon fixation? Biochim Biophys Acta 891:205–215

Ross et al.: Effect of vertical mixing on phytoplankton growth 29

Page 18: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Mar Ecol Prog Ser 435: 13–31, 201130

Waite A, Fisher A, Thompson PA, Harrison PJ (1997) Sinkingrate versus cell volume relationships illuminate sinkingrate control mechanisms in marine diatoms. Mar Ecol ProgSer 157:97–108

Williams PJB, Heinemann KR, Marra J, Purdie DA (1983)Comparison of 14C and O2 measurements of phytoplank-ton production in oligotrophic waters. Nature 305:49–50

Wolf KU, Woods JD (1988) Lagrangian simulation of primary

production in the physical environment — the deepchlorophyll maximum and nutricline. In: Rothschild BJ(ed) Toward a theory in biological-physical interactions inthe world ocean, NATO ASI Series C, Kluwer, Dordrecht,p 51–70

Woods JD, Onken R (1982) Diurnal variation and primary pro-duction in the ocean—preliminary results of a Lagrangianensemble model. J Plankton Res 4:735–756

Symbol Description Value Unit

a Factor by which dark N uptake rate is reduced 0.75 –aChl Chl a specific absorption coefficient 6.56 m2 gChl−1

CF Functional C content Eq. (24) pgC cell−1

CR Reserve C content Eq. (25) pgC cell−1

CT Total C content (functional plus reserve) CF + CR pgC cell−1

CFfis Functional C content required for cell fission 14 pgC cell−1

CFmin Subsistence level in functional pool 0.25 CF

fis pgC cell−1

CRthres C storage capacity to maintain cell through dark (24 − L)Rm

cell pgC cell−1

Chl Amount in chlorophyll a in cell Variable gChl cell−1

DIN Dissolved inorganic nitrogen concentration 500 mgN m−3

I Light intensity Variable μmol photons m−2 s−1

Ik Light saturation parameter Variable μmol photons m−2 s−1

kr Repair rate from photoinhibition 4.5 × 10−5 s−1

κN Half-saturation for N uptake 14 mgN m−3

L Light period 15 hN Cellular nitrogen Variable pgN cell−1

Pf Scaling factor 1.5 –P cell

max Max. cell-based C production Variable gC cell−1 s−1

Pmchl

Max. Chl-based C production (with inhibition) Pchlmax ϑ gC gChl−1 s−1

Pchlmax Max. Chl-based C production (without inhibition) 3.3 × 10−3 gC gChl−1 s−1

PFmax Max. CF -based production at a given Q Variable s−1

PCmax Max. CT-based production at a given Q Eq. (19) s−1

Q N:C ratio of entire cell N/CT gN gC−1

QF N:C ratio of functional pool (≡ Qmax) 0.19 gN gC−1

Qmin Min. N:C ratio of entire cell when CR is maximal 0.05 g N g C−1

Rmcell Cell-based maintenance respiration 0.05 pgC (cell d)−1

Tf Delay to complete cell division after reaching CFfis 2 h

VNmax Max. N uptake rate Eq. (23) gN (cell d)−1

αChlmax Initial slope of the P-I curve (without inhibition) 6.3 × 10−6 gC m2 (gChl μmol photons)−1

Δt Model time step 1–6 s φmax Max. quantum yield of photosynthesis 0.96 × 10−6 gC (μmol photons)−1

φm Achieved quantum yield when inhibited Variable gC (μmol photons)−1

μmax Maximum specific growth rate 1.15 d–1

ΠLmax Max. prop. of CF allocated to light harvesting 0.33 –

ϑ Proportion of functional PSII reaction centres 0 ≤ ϑ ≤ 1 – σPSII Absorption cross-section of PSII 1.5 m2 (μmol photons−1)θC Chl:C ratio of cell Chl/CT gChl gC−1

θCF Chl:C ratio for functional pool only Chl/CF gChl gC−1

(θCL )max Max. Chl:C ratio in light-harvesting pool 0.28 gChl gC−1

(θCF )max Max. Chl:C ratio in functional pool (θC

L )maxΠLmax gChl gC−1

ζ Cost of biosynthesis 3.0 gC gN−1

Table A1. Biological model parameters

Appendix 1. Biological model parameters (Table A1) and equations (Table A2). This section provides a summary of the main biological model equations, containing the equations from Tables 3 & 4 from Ross & Geider (2009) in their modified version usedin this study, plus the additions for photoinhibition. See the model description in Ross & Geider (2009) for a more detailed

explanation of each equation

Page 19: Modelling the effect of vertical mixing on bottle incubations for …repository.essex.ac.uk/1009/1/m435p013.pdf · 2011. 9. 23. · mixing in the SML. As it is currently not possible

Ross et al.: Effect of vertical mixing on phytoplankton growth 31

(15)

(16)

(17)

(18)

(19)

(20)

(21)

(22)

(23)

(24)

(25)

(26)

(27)

Cell division is triggered when CF > CFfis

after a delay tfis(28)

Cell death occurs when CF < CFmin (29)

P PI Chl

P

Chl

cellcell

maxcell

max

= 1− −⎛⎝⎜

⎞⎠⎟

⎡⎣⎢

⎤⎦

expα

⎥⎥

PP C

P Chlax

F F

mchlmax

cell mMIN=⎧⎨⎪

⎩⎪

P Pmchl

xchl= ma ϑ

P PL

QFf x

Fmax =

241μ ζma +( )

IP C

CP

k

F

chl C

F

T

C

chl C= =max max

α θ α θ

α φ φ ϑ α ϑChlm

Chlax

Chlax

Chla a= = =m m

dd

P

cellmaϑ σ ϑ φ

t

k IPI a Chl

kd SIIx

Chl r=1

/(1− −⎛

⎝⎜⎞⎠⎟ + − ϑϑ

ϑ

) , > 0

(1 ) , = 0

if

if

I

k Ir −

⎨⎪

⎩⎪

dd max

min

Nt

V VQ Q

Q QDIN

DINN N

F

F

n

N

≡ −−

⎡⎣⎢

⎤⎦⎥ +⎛⎝

⎞⎠=

κ

VQ C a

La I

a Qax

N axF F

Fm

m

max

if > 0=

124

μ

−( ) +⎡⎣⎢

⎤⎦⎥

CC F otherwise

⎧⎨⎪

⎩⎪

dd

if >

if = 0 and

thresCt

VQ

C C

R C

F

N

FR R

mcell R=

,

− CC F > 0

otherwise

,

0

⎨⎪⎪

⎩⎪⎪

dd

dd

cell cellCt

PCt

V RR F

Nm= − − −ζ

dd

if > andmax max thres

Chlt

VQ

C C IN

FL

LC R R

=

( )Π Ωθ ρ >> 0

if > C and = 0max max thres

,

( ) ,VQ

C IN

FL

LC R RΠ θ

−−

⎪⎪⎪⎪

⎪⎪⎪⎪

R CmC Rcell if = 0

otherwise

θ ,

0

ρ = − −⎛⎝⎞⎠

⎡⎣⎢

⎤⎦⎥

II

II

k

k

1 exp

Table A2. Biological equations

Editorial responsibility: Katherine Richardson, Copenhagen, Denmark

Submitted: September 30, 2010; Accepted: May 2, 2011Proofs received from author(s): August 12, 2011