<<

PHOTOPHYSIOLOGICAL PROPERTIES OF THE MARINE PICOEUKARYOTE

PICOCHLORUM RCC 237 (TREBOUXIOPHYCEAE, CHLOROPHYTA)1

Céline Dimier, Federico Corato, Giovanni Saviello and Christophe Brunet2

Stazione Zoologica "Anton Dohrn", Villa Comunale, 80121 Naples, Italy

Running title: Photophysiology of Picochlorum

1Received Accepted

2Author for correspondence: email [email protected]

1 Key index words: Fluorescence quantum yield; Non-Photochemical Quenching of fluorescence;

Photoacclimation; ; Xanthophyll Cycle; Zeaxanthin.

Abbreviations: Ax, antheraxanthin; chl a, a; DPS, de-epoxidation state

[=(Ax+Zx)/(Vx+Ax+Zx)]; DTT, Dithiotreitol; ETR, electron transport rate; FALS, Forward

Angle Light Scatter; Fq'/Fm', operating efficiency of PSII; Fv/Fm, maximum photochemical

efficiency of PSII; HL, High Light; LHC, Light-Harvesting Complex; LL, Low Light; ML,

Moderate Light; NF, Norflurazon; NPQ, Non-photochemical quenching; PAM, Pulse Amplitude

Modulation; PFD, Photon flux density; PSII, Photosystem II; RALS, Right Angle Light Scatter;

RLC, Rapid light curve; Vx, violaxanthin; Zx, zeaxanthin.

2 The photophysiological properties of strain RCC 237 belonging to the marine picoplanktonic Picochlorum (Henley et al.) genus were investigated under different photon flux densities (PFD, ranging from 40 to 400 µmol photons·m-2·s-1), mainly focusing on the development of the xanthophyll cycle and its relationship with the non-photochemical quenching of fluorescence (NPQ). The xanthophyll cycle functioning and its photoprotective role was investigated by applying a progressive increase of PFD and using Dithiotreitol and Norflurazon to block specific enzymatic reactions in order to deeply study the relationship between xanthophyll cycle and NPQ. These two processes were significantly related only during the gradually increasing light periods and not during stable light periods, where NPQ and zeaxanthin were decoupled. This result reveals that NPQ is a photoprotective process developed by algae only when cells are experiencing increasing PFD or in response to stressful light variations, for instance after a sudden light shift. Results showed that the photobiological properties of

Picochlorum strain RCC 237 seem to be well related to the surface water characteristics, since it is able to maintain its photosynthetic characteristics under different PFDs and to quickly activate the xanthophyll cycle under high light.

3 INTRODUCTION

In natural environments, cells continuously experience light variations due to the diel cycle and the vertical displacement of cells within the mixed layer due to hydrodynamics. Since light is crucial for survival and success of algae in aquatic , they have to continuously photoacclimate, through biochemical, physiological and/or morphological variations. Photophysiological properties exhibited by algae seemed to be related to the environment where they grow (Stomp et al. 2004, Strzepek and Harrison 2004), converging toward the hypothesis that light is a relevant factor driving competition, exclusion and/or diversity in the algal (Litchman and Klausmeir 2001, Floder and Burns 2005).

An important trait of phytoplankton functional diversity is cell size that is in part responsible for the variety of biological and/or ecological behaviors with respect to the environment, e.g. sinking rate, nutrient consumption, light utilization and packaging effect (Fogg 1991, Raven 1998).

Picoplankton (size < 3 µm) are known to be important contributors to communities in many coastal and oceanic ecosystems (e.g. Raven 1998, Not et al. 2005). The high level of taxonomic biodiversity in the picoeukaryotic fraction (Moon-van der Staay et al. 2001, Diez et al. 2004) reinforces the key role of this group in marine structure and functioning.

Nevertheless, very few data exist on their photophysiology (e.g. Wilhem et al. 1989;

Timmermans et al. 2005), in contrast to (e.g. Bibby et al. 2003 and references therein, Stomp et al. 2004) and to the larger cell size eukaryotic community (e.g. Casper- Lindley and Bjorkman 1998, Lavaud et al. 2004).

An important photoprotective mechanism active at short temporal scales is the dissipation of excess energy, through the non-photochemical quenching of chl a fluorescence that is catalyzed

4 by the xanthophyll cycle (Finazzi et al. 2006). This process involves enzymatic conversion from violaxanthin to antheraxanthin and zeaxanthin in plants and (Gilmore et al. 1994,

Masojidek et al. 1999) and from diadinoxanthin to diatoxanthin in chromophyte alge (Lavaud et al. 2002). Accumulation of zeaxanthin or diatoxanthin is triggered by the formation of a pH gradient across the thylakoid membrane (Demmig-Adams and Adams 2000). Much attention has been paid to the photoprotection mechanism in chromophyte algae (e.g Lavaud et al. 2004 and references therein, Harris et al. 2005), while few studies have dealt with marine green micro- algae (e.g. Gilmore and Yamamoto 2001, Garcia-Mendoza et al. 2002). The presumed minor contribution of green algae into the phytoplankton community, as well as the similarity of their xanthophyll cycle to the well-documented one found in terrestrial plants, (e.g. Havaux and

Niyogi 1999, Demmig-Adams and Adams 2000) make the study of photoregulation in this group of low ecological interest. Recently, it has been shown that these algae are among the main picoeukaryote contributors in many ecosystems (Diez et al. 2004, Not et al. 2005), highlighting their capacity to adapt to different ecosystem characteristics and their probably relevant role in ecosystem structure and functioning.

This study investigates the photophysiological properties of the strain RCC 237 belonging to the picoeukaryotic genus Picochlorum (Chlorophyta, Trebouxiophyceae). The short-term photoacclimative process is dissected in order to test the presence of an efficient acclimation to high photon flux density, through rapid xanthophyll cycle activation. Our assumption is based on the biological peculiarities of picoeukaryotes, e.g. low sinking rate and packaging effect (Raven 1998), that may lead this group to be more sensitive to light than large cells. This could be the case of the strain RCC 237 that was isolated from Mediterranean surface waters. Pigments, quenching of fluorescence and PSII efficiency were used as photoacclimative

5 indices during different light treatments and in presence or absence of enzymatic inhibitors

linked to xanthophyll cycle activity.

MATERIALS AND METHODS

Algal model and culture conditions. The strain RCC 237 of Picochlorum (Chlorophyta,

Trebouxiophyceae), isolated by Dr. D. Vaulot from the surface layer (20 m depth) of the

Mediterranean Sea, was provided by the Roscoff Culture Collection (France; Vaulot et al. 2004).

This small non-flagellated coccoid strain (diameter 1.5 µm) was cultivated non-axenically at 20°

-2 -1 C under 40 µmol photons·m ·s (measured by a 4 π QSL-2100, Biospherical instruments INC,

San Diego, USA) provided by one lamp (OSRAM, Decostar 51, 50 W; Munich, Germany) with

a 12:12 light:dark photoperiod. Algae were grown in Keller medium (Keller et al. 1987) in 3-L

glass cylinders and the cultures were continuously aerated and maintained in exponential phase

by daily dilution of half of the culture with fresh medium during more than 10 days before the

experiments. The pH and temperature were estimated daily with a HI- 9214-Stick pHmeter

(Hanna Instruments, Woonsocket, USA) while cell concentration was controlled by flow

cytometry (see below).

Experimental design and sampling. The four experiments were conducted in triplicate

aerated cultures at 20° C. Each culture flask was illuminated by one lamp (OSRAM, Decostar

51, 50 W; Munich, Germany), and the three lamps were mounted on a prototype device, called

“PLIS” (“Progressive Light Increase System”) allowing gradual changes of PFD on the culture

flasks, controlled by a BASIC program. A 3-L glass culture flask was maintained under the

6 initial conditions as an experimental control. During the four experiments, 30 mL of culture was

sampled (see below) for HPLC-analyzed pigments, absorption spectrum on a filter by a

spectrophotometer, cellular parameters by , quantum yield of fluorescence and

rapid light curves (RLCs) by Phyto-Pam.

Low light to high light experiment. The “PLIS” was setup as follows: after 30 minutes at

40 µmol photons·m-2·s-1, light was gradually increased to 200 µmol photons·m-2·s-1, over 60 min.

-2 -1 For the following three hours, light remained constant at 200 µmol photons·m ·s and then

-2 -1 increased to 400 µmol photons·m ·s over 30 min, remaining constant at this high light value for

2 hours. Frequency of sampling was high, especially during the two increasing light periods, with a total of 23 sampling points (Fig. 1a)

High light to low light experiment. This experiment was conducted on cells acclimated at

-2 -1 400 µmol photons·m ·s during 6 hours (i.e. the duration of the previous experiment) in order to

study the relaxation of the photosynthetic apparatus under low light (40 µmol photons·m-2·s-1) during 3 hours.

DTT addition experiment. After one hour of illumination at 40 µmol photons·m-2·s-1, the

-1 DTT was injected in aqueous solution into the culture to a final concentration of 500 µmol·L

(Lohr and Wilhelm 2001). Cells were incubated for 10 minutes with the inhibitor before the shift

to high light (400 µmol photons·m-2·s-1). The control culture with no-addition of DTT followed

the same light change as the three other flasks.

7 NF addition experiment. A methanolic solution of NF was injected into the culture flask

-1 to a final concentration of 10 µmol·L methanol (2.5% v/v methanol; Garcia-Plazaola et al.

2002). Cells were incubated for 10 minutes with the inhibitor before being shifted to 400 µmol photons·m-2·s-1. It was previously verified that methanol (without NF) at this concentration has

no effect on the algal photophysiology (data not shown).

Pigment analysis. Samples of 10 mL were filtered onto GF/F glass-fiber filters

(Whatman, Maidstone, UK) and immediately stored in liquid nitrogen until analysis. Using the method outlined in Casotti et al. (2005), the pigment solution extract was injected in a Hewlett

Packard series 1100 HPLC (Hewlett-Packard, Kennett Square, PA, USA) with a C8 BDS 3 µm

Hypersil, IP column (Thermo Electron corporation, USA). The mobile phase was composed of two solvents: A, methanol, aqueous ammonium acetate (70:30) and B, methanol. Pigments were

detected spectrophotometrically at 440 nm using a Hewlett Packard photodiode array detector

model DAD series 1100 while fluorescent pigments were detected using a Hewlett Packard

standard FLD cell series 1100 with excitation and emission wavelengths setup at 407 nm and

665 nm, respectively. Determination and quantification of pigments were realized using pigment standards from the D.H.I. Water & Environment (Hørsholm, Denmark).

Absorption spectrum. Absorption was measured on 10 ml culture samples filtered onto

GF/F filters (Whatman, 25 mm), following the procedure of Tassan and Ferrari (1995), between

400 and 800 nm with 1-nm increment on a spectrophotometer (Hewlett Packard HP-8453E)

equipped with an integrating sphere (Labsphere RSA-HP-53).

8 Flow cytometry analysis. Duplicate 1 ml samples were fixed in 10% paraformaldehyde

and immediately stored in liquid nitrogen. Analysis was done with a FACScalibur (Becton

Dickinson, USA) using 6 µm beads (Flow set, Coulter) as internal standard. The method and apparatus was the same as in Casotti et al. (2005).

Fluorescence measurement. Photochemical efficiency of PSII was estimated with a

Phyto-PAM fluorometer (Heinz Walz GmbH, Effeltrich, Germany). Maximum photochemical efficiency (Fv/Fm, with Fv = Fm-Fo) was measured on 15 min dark-adapted sample while the operating photochemical efficiency (Fq’/Fm’, with Fq’ = Fm’- F’) was measured on actinic light- exposed sample (Harris et al. 2005). The effective non-photochemical quenching of fluorescence was quantified by the Stern-Volmer expression:

effNPQ = (Fm/Fm’) - 1 (1)

where, Fm and Fm’ are the maximum fluorescence values from dark- and actinic light-exposed

samples, respectively. Fm and Fm’ were measured after a saturation pulse of bright red light

(intensity of 2400 µmol photons·m-2·s-1) applied during 450 ms, causing a complete reduction of

the PSII acceptor pool.

The RLCs (rapid light curves) were determined at 5 sampling points (0, 70, 210, 310 and

390 min) on 15 min dark-adapted samples by applying ten increasing actinic irradiances (from 8 to 1500 µmol photons·m-2·s-1, two minutes at each PFD level). The absolute electron transport rate was calculated as:

9 absETR = (Fq’/Fm’) x PFD x (a*ph/2) (2)

-2 -1 where, PFD was the incident irradiance (µmol photons·m ·s ) and (a*ph/2) the mean absorption

value of phytoplankton normalized by chl a (m2·mg chl a-1) and divided by 2 under the

assumption that half of the absorbed light is distributed to PSII. The RLCs were fitted with the equation of Eilers and Peeters (1988) to estimate the photosynthetic parameters αetr, Ek and

absETRmax.

In the RLC, operating photochemical efficiency decreases as NPQ develops (Villareal

2004). This NPQ capacity was calculated after each actinic irradiance level according to equation

(1), where Fm was the maximum fluorescence measured after the 15 min dark-adaptation and

Fm’ was the maximum fluorescence measured after each actinic light level. The maximum NPQ

(maxNPQ) was determined as the highest value reached on the NPQ coefficient vs PFD curve.

RESULTS

The low light acclimated cells were characterized by a dominance of violaxanthin (Vx,

72%) among the three-xanthophyll cycle pigments (violaxanthin-antheraxanthin-zeaxanthin, Vx-

Ax-Zx, Fig. 1b) and a high value of maximum photochemical efficiency (Fv/Fm = 0.69). These

-1 values as well as the sustained growth rate (0.39 d ) and the high values of absETRmax and

- -1 -1 - -1 -1 - photosynthetic efficiency (0.85 mol e ·g chl a ·h and 0.013 mol e ·g chl a ·h ·(µmol photons·m

2·s-1)-1, respectively) revealed a good acclimative state to low light. The strong decrease of

Fq’/Fm’ for PFDs > 250 µmol photons·m-2·s-1 along the RLC (Fig. 2a) showed a high capacity

for energy dissipation, consistent with the NPQ capacity evolution and its maximum value (1.94,

Fig. 2b).

10

Gradual increasing light experiment. The de-epoxidation of Vx into Ax and Zx mainly

occurred during the two increasing light periods (Fig. 1, b and c) and was negatively correlated with the maximum photochemical efficiency (p<0.001), which slightly decreased with time (-

8.5%, data not shown). Activation of the Vx de-epoxidase already occurred at low light since the

synthesis of Ax was initiated after 12 min (i.e. 72 µmol photons·m-2·s-1, Fig. 1b). Significant

linear relationships were found between Zx/chl a and effNPQ only during the two increasing light

periods:

2 effNPQ = 6.0 (Zx/chl a) – 0.31 (r = 0.75, n=9) (3)

for the 40 to 200 µmol photons·m-2·s-1 light shift, and

2 effNPQ = 0.67 (Zx/chl a) – 0.07 (r = 0.85, n=9) (4)

for the 200 to 400 µmol photons·m-2·s-1 light shift.

The relationships (3) and (4) revealed that not all the intracellular zeaxanthin was linked

to energy dissipation. Indeed, effNPQ reached approximately the same value at the end of the two

increasing light periods (around 0.10, Fig. 1c) despite a Zx/chl a two fold higher under 400 with

respect to 200 µmol photons·m-2·s-1 (0.025 vs 0.012 mol Zx·mol chl a-1, respectively; Fig. 1b).

During the two stable light periods, the effNPQ decreased while Ax/chl a and Zx/chl a continued

to increase and Vx/chl a reached a plateau, showing a decoupling with the xanthophyll cycle

11 (Fig. 1, b and c). The effNPQ decreased faster at high light than under moderate light (< 30 min

and 2 hours after light stability, respectively). On the contrary, no difference was revealed in the

NPQ capacity evolution along the RLC for these two sampling points (Fig. 2b). The higher zeaxanthin content under moderate light with respect to low light allows a faster photoprotective response when cells experience increasing values of PFD, with little need to dissipate excess energy through NPQ. This increase in zeaxanthin in relation to the difference in PFD (200 or 400

µmol photons·m-2·s-1) was probably related to the different conformation of the light harvesting

complexes (Pascal et al. 2005). The amount of other accessory pigments, such as chl b/chl a and

ß-carotene/chl a, remained constant during the time course of the experiment, whereas lutein/chl a increased by 20% (data not shown). Photoregulation process developed by cells during this experiment seemed to be efficient since the photosynthetic characteristics (e.g. Fig. 2a) did not vary with increasing PFD. Also, the effNPQ remained low and the capacity to develop NPQ did

not significantly change with time (Fig. 2b). Algal growth was enhanced under high light, as revealed by the stronger increase of FALS and RALS (Fig. 3, a and b; +150.4 % and +67%,

respectively) with respect to low light (+74% and +38.8%, respectively). Increase of red chl a

-1 fluorescence and similar mean of chl a·cell in both the light conditions (Fig. 3c, p > 0.05, mean

± SD = 0.052 ± 0.005 fmole chl a·cell-1) revealed that cells were not negatively affected by high

light.

High light to low light experiment. The xanthophyll pool of the high-light-acclimated

cells (6 hours at 400 µmol photons·m-2·s-1) was dominated by Ax (51% vs 26.1% (Vx) and

22.7% (Zx), initial values in Fig. 4a). As at the end of the previous experiment (PFD = 400 µmol

-2 -1 photons·m ·s ), the effNPQ was very low (0.02) and the maximum photochemical efficiency was

12 high (Fv/Fm = 0.65). Under LL, the Zx/chl a started to decrease after 2 min and Vx/chl a

strongly increased (Fig. 4a). In less than two hours, the xanthophyll pigments reached a plateau

(Fig. 4a), with Vx, Ax and Zx contributing for 64%, 18% and 18%, respectively, which was

similar to the values found for the low light acclimated cells. The PSII reaction centers quickly

opened, as suggested by the rapid increase of the Fq'/Fm', which almost reached the value of

Fv/Fm (Fig. 4b). The maxNPQ decreased during this experiment (from 2.36 to 1, data not shown) in agreement with the lowering of the photoprotective pigment pool.

Use of dithiotreitol. Death or stress were not evident by flow cytometry (see Casotti et al.

2005; data not shown) during the experiment. The initial rise of effNPQ in the control and +DTT

cultures (Fig. 5a) was not related to Zx synthesis (p>0.05), revealing a rapid formation of a Zx-

independent NPQ due to the sudden light shift from 40 to 400 µmol photons·m-2·s-1. Thus the

effNPQ decreased only in +DTT, probably because Zx pool was not high enough to maintain the

thermal dissipation rate of excess of energy in the cells (Fig. 5, b and c). Indeed, the de- epoxidation was inhibited and zeaxanthin decreased in +DTT after 30 minutes (Fig. 5, b and c), in agreement with results from Gilmore et al. (1994) who showed that epoxidation can be activated during light-induced lumen acidification after inhibition of the de-epoxidase with DTT.

This hypothesis would also explain the decrease of Ax as already described by Gilmore et al.

(1998) in higher plants.

Use of norflurazon. The addition of NF did not obviously affect the amount of Vx, Ax

and Zx, at least in the first 60 min of the experiment, under high light compared to the control

(Fig. 6, a and b). Since NF inhibited the phytoene desaturase, cells apparently used the pre-

13 existing pool of ß-carotene to produce Vx, Ax and Zx. Indeed, the decrease of ß-carotene (-

0.007 mol·mol chl a-1) almost corresponded to the increase in VxAxZx pool (+0.010 mol·mol chl

a-1, Fig. 6, a and c). The maxNPQ estimated from the RLC decreased from 1.35 to 0.35 after 1

hour in +NF, whereas it remained constant in the control (data not shown), revealing a decrease

of the cell’s capacity for thermal energy dissipation. This suggests photoinhibition, consistent

with a ~30% reduction of Fv/Fm in the NF-treated culture with respect to the control (Fig. 6f), in

part due to an increase in Fo in +NF (+ 36%, data not shown). We have no explanation for the

peak of effNPQ in +NF cells 10 min after the shift to high light (Fig. 6e).

DISCUSSION

During this study the photoprotective processes developed at short temporal scale by strain RCC 237 of Picochlorum was investigated. Little physiological information is available on the picoeukaryotes (e.g. Sobrino et al. 2005, Timmermans et al. 2005, Veldhuis et al. 2005,

Finazzi et al. 2006), whereas their diversity in marine ecosystems is usually high (Diez et al.

2004, Not et al. 2005). Indeed, the suggested differentiation in ecotypes of some picoeukaryotes

(Rodriguez et al. 2005) makes it difficult to make any generalizations about the physiological results obtained from one strain. Meanwhile, their biological peculiarities (e.g. low sinking rate, small package effect, efficient resource acquisition and use for growth; Raven, 1998) make them a relevant model for comparison with larger cells, such as . Whereas this strain cannot be considered as a model for chl b-containing picoeukaryotes nor for the genus Picochlorum, results obtained in this study provide information about the photoacclimation properties of a green picoeukaryote growing in oceanic surface waters. Photo-physiological and photoprotective

14 properties differ among phytoplankton according to their phylogeny and ecology (Casper-

Lindley and Bjorkman 1998, Strzepek and Harrison 2004). The strain RCC 237 grows better

-2 -1 -2 -1 -1 -1 under 400 µmol photons·m ·s than under 40 µmol photons·m ·s (µ = 0.80 d and 0.40 d , respectively) and its photophysiological properties seem to be adapted to the high and variable irradiance characteristics of oceanic surface water. This strain is able to photoacclimate to a broad range of PFD by maintaining its photosynthetic characteristics, such as absETRmax (0.85

- -1 -1 - -1 -1 - mol e ·g chl a ·h ), photosynthetic efficiency (αetr = 0.013 mol e ·g chl a ·h ·(µmol photons·m

2·s-1)-1), and light saturation (Ek = 65 µmol photons·m-2·s-1). Assuming a transfer of four

-1 -1 electrons for 1 mol O2, a rate of 0.21 mol O2·g chl a ·h has been estimated from the absETRmax value, indicating similar photosynthetic performance as for instance grown under 50 µmol photons·m-2·s-1 (Harris et al. 2005). The fast recovery of PS II properties (< 1 hour, e.g. pigments or maximum photochemical efficiency) at low PFD reveals the high

plasticity of PSII for optimizing photochemistry, in agreement with the ability to sustain growth

under variable PFD.

Comparing the de-epoxidation state (DPS) values reached under high PFD, the strain RCC 237

seems able to activate the xanthophyll cycle more rapidly with respect to other green algae, such as Dunaliella tertiolecta (Casper-Lindley and Bjorkman 1998). More generally, the DPS value

reached under different PFD is higher in phytoplankton (> 0.25 for a range of light from 8 to 40

µmol photons·m-2·s-1, Schubert et al. 1994; Havelkova-Dousova et al. 2004; this study) than in

terrestrial plants (<0.12 for a range of light from 50 to 220 µmol photons·m-2·s-1, Havaux et al.

2000, Garcia-Plazaola et al. 2002). This feature could be linked to the more variable PFD

experienced by micro-algae that would probably require faster and/or stronger photoprotection

responses as compared to terrestrial plants. This hypothesis agrees with the results of Lavaud et

15 al. (2002) and Garcia-Mendoza et al. (2002), who showed that the xanthophyll content increased

under intermittent high light in diatoms and chlorophytes. The progressive increase of PFD

seems to enhance the synthesis of photoprotective pigment (reaching 0.025 mol Zx·mol chl a-1,

Fig. 1b) with respect to a sudden light shift (reaching 0.015 mol Zx·mol chl a-1, Fig. 6b). The

increase and maintenance of the Zx pool under moderate or high PFD allows a quicker effNPQ formation when cells experience a further increase of PFD. During stable light periods, the decrease of effNPQ with a constant DPS agrees with the fact that only ∆pH and not de-

epoxidation state is relaxed after light increase (Ruban and Horton 1999). The quicker decrease

-2 -1 -2 -1 of effNPQ under constant 400 µmol photons·m ·s than under 200 µmol photons·m ·s may relate to the higher content of Zx even though light is two fold higher. The ecological advantage may be a faster photoregulation under changing PFD, with little necessity to dissipate excess light. The overall lack of relationship between NPQ and zeaxanthin is due to the fact that not all the zeaxantin participates to the quenching of fluorescence and to the fast development of NPQ independently from zeaxanthin. NPQ is directly related to the synthesis of zeaxanthin only during gradual increasing light periods. The effNPQ per mole of Zx is much higher during the change from 40 to 200 µmol photons·m-2·s-1 than from 200 to 400 µmol photons·m-2·s-1 (see relationships 4 and 5) and the zeaxanthin synthesis was higher during the second light shift than during the first one (+ 115% and 40 %, respectively). Tardy and Havaux (1996) showed that a specific subset of Zx molecules could participate to NPQ formation. Since the effNPQ reaches the

-1 same value at the end of the two increasing light periods, the value of 0.012 mol Zx·mol chl a

(maximal value obtained after the first increasing light period) must correspond to the upper limit

of Zx contributing to NPQ in the strain RCC 237 of Picochlorum. The same quantity of Zx per

chl a must be devoted to another function under high PFD, such as a slow down of D1 protein

16 degradation rate (Jahns et al. 2000), or lipid peroxidation prevention (Havaux and Niyogi 1999).

The synthesis of zeaxanthin begins at the photosynthetic light saturation (Ek), i.e. around 65

µmol photons·m-2·s-1. On the contrary, a rapid quenching of fluorescence occurred without any

increase in zeaxanthin when cells experienced either sudden or gradual increase of PFD meaning

that is caused by an increasing light rather than excessive light per se. Indeed, the development

-2 -1 of effNPQ occurs at subsaturating PFD (significant increase at 50 µmol photons·m ·s ), revealing that energy dissipation begins before photosynthetic light saturation.

Zeaxanthin-independent NPQ can develop rapidly in parallel with the protonation of the PsbS protein in plants (Dall’Osto et al. 2005, Pascal et al. 2005). After formation, this NPQ is maintained by binding of Zx on PsbS (Niyogi et al. 2004). This two-step functioning is conserved in chl b-containing photoautotrophs from terrestrial plants (Niyogi et al. 2004) to phytoplankton (e.g. Casper-Lindley and Bjorkman 1998, Masojidek et al. 1999, this study).

Nevertheless, the NPQ seems to be mainly zeaxanthin-dependent when this pigment is already present in relatively high amount in the cell as revealed during the second light increase (from

200 to 400 µmol photons·m-2·s-1, Fig. 1, b and c behavior could present ecological advantage,

since the protein PsbS enables fast and reversible NPQ, depending on zeaxanthin removal from

the binding sites and not on its epoxidation (Horton and Ruban 2005). The presence of this

mechanism in green algae and terrestrial plants could explain the lower pH needed for the

activation of the de-epoxidase enzyme (= 6.5, Emanuelson et al. 2003) with respect to

chromophyte algae, such as diatoms (= 7.2, Jakob et al. 2001), in which the PsbS protein has not

yet been found (Horton and Ruban 2005). State transition, an additional photoprotective

mechanism responsible for the uncoupling between zeaxanthin and NPQ in some green algae

17 (Masojidek 2004), has still to be investigated in green marine picoeukaryotes such as the

Picochlorum genus.

Acknowledgements

The authors thank the Roscoff Culture Collection (Roscoff, France) for sending us the strain of

Picochlorum sp. and R. Casotti for her flow cytometry experience. We also thank S. McDonald

for the English revision of the manuscript. Two anonymous referees and D. Vaulot are gratefully acknowledged for the numerous and helpful comments on the manuscript. C.D. is supported by a

Ph.D grant from SZN. This publication represents a contribution to the aims of the MarBef

Network of Excellence “Marine Biodiversity and Ecosystem Functioning” which is funded in the

Community’s Sixth Framework Programme (contract no. GOCE-CT-2003-505446). This publication is contribution number MPS07002 of MarBef.

18 List of references

Bibby, T.S., Mary, I., Nield, J., Partensky, F. & Barber, J. 2003. Low-light adapted

Prochloroccus species possess specific antennae for each photosystem. Nature 424: 1051-1054.

Casotti, R., Mazza, S., Brunet, C., Vantrepotte, V., Ianora, A. & Miralto, A. 2005. Growth

inhibition and toxicity of the aldehyde2-trans, 4-trans-decadienal on Thalassiosira

weissflogii (Bacillariophyceae). J. Phycol. 41: 7-20.

Casper-Lindley, C. & Bjorkman, O. 1998. Fluorescence quenching in four unicellular algae with

different light-harvesting and xanthophyll-cycle pigments. Photosynth. Res. 56: 277-289.

Dall'Osto, L., Caffari, S. & Bassi, R. 2005. A mechanism of non-photochemical energy

dissipation, independent from PsbS, revealed by a conformational change in the antenna protein

CP26. Plant Cell 17: 1-16.

Demming-Adams, B. & Adams III, W.W. 2000. Harvesting sunlight safely. Nature 403: 371-

374.

Díez, B., Massana, R., Estrada, M. & Pedrós-Alió, C. 2004. Distribution of eukaryotic

picoplankton assemblages across hydrographic fronts in the Southern , studied by

denaturating gradient gel electrophoresis. Limnol. Oceanog. 49: 1022-1034.

Eilers, P.H.C. & Peeters, J.C.H. 1988. A model for the relationship between light intensity and the rate of photosynthesis in phytoplankton. Ecol. Model. 42: 199-215.

Emanuelson, A., Eskling, M. & Akerlund, H.E. 2003. Chemical and mutational modifications of histidines in violaxanthin de-epoxidase from Spinacia oleracea. Physiol. Plant. 119: 97-104.

19 Finazzi, G., Johnson, G.N., Dall’osto, L., Zito, F., Bonente, G., Bassi, R. & Wollman, F.A. 2006.

Nonphotochemical quenching of chlorophyll fluorescence in Chlamydomonas reinhardtii.

Biochemistry, 45: 1490-1498.

Fisher, T., Berner, T., Iluz, D. & Dubinsky, Z. 1998. The kinetics of the photoacclimation

response of sp. (Eustigmatophyceae): A study of changes in ultrastructure and

PSU density. J. Phycol. 34: 818-824.

Floder, S. & Burns, C.W. 2005. The influence of fluctuating light on diversity and species

number of nutrient-limited phytoplankton. J. Phycol. 41: 950-956.

Fogg, G.E. 1991. The phytoplanktonic ways of life. New phytol. 118 : 191-232.

Garcia-Mendoza, E., Matthijs, H.C.P., Schubert H. & Mur, L.R. 2002. Non-photochemical

quenching of chlorophyll fluorescence in Chlorella fusca acclimated to constant and dynamic

light conditions. Photosynth. Res. 74: 303-315.

Garcia-Plazaola, J.I., Hernandez, A., Artetxe, U. & Becerril, J.M. 2002. Regulation of the

Xanthophyll cycle pool size in duckweed (Lemna minor) plants. Physiol. Plant. 116: 121-126.

Gilmore, A.M., Mohanty, N. & Yamamoto, H.Y. 1994. Epoxidation of zeaxanthin and

antheraxanthin reverses non-photochemical quenching of photosystem II chlorophyll a

fluorescence in the presence of trans-thylakoid delta-pH. FEBS Lett. 350: 271-274.

Gilmore, A. M., Shinkarev, V.P., Hazlett, T. L. & Govindjee 1998. Quantitative analysis of the

effects of intrathylakoid pH and xanthophyll pigments on chlorophyll a fluorescence lifetime

distributions and intensity in thylakoids. Biochemistry 37: 13582-13593.

Gilmore, A. M. & Yamamoto, H.Y. 2001. Time-resolution of the antheraxanthin- and delta pH-

dependent chlorophyll a fluorescence components associated with photosystem II energy

dissipation in Mantoniella squamata. Photochem. Photobiol. 74: 291-302.

20 Harris, G.N., Scanlan, D.J. & Geider, R.J. 2005. Acclimation of Emiliania huxleyi

(Prymnesiophyceae) to photon flux density. J. Phycol. 41: 851-862.

Havaux, M. & Niyogi, K.K. 1999. The violaxanthin cycle protects plants from photooxidative damage by more than one mechanism. Proc. Natl. Acad. Sci. U.S.A. 96: 8762-8767.

Havaux, M., Bonfils, J.P., Lutz, C. & Niyogi, K.K. 2000. Photodamage of the photosynthetic apparatus and its dependence on the leaf developmental stage in the npq1 Arabidopsis mutant

deficient in the xanthophyll cycle enzyme violaxanthin de-epoxidase. Plant Physiol. 124: 273-

284.

Havelkova-Dousova, H., Prasil, O. & Behrenfeld, M.J. 2004. Photoacclimation of Dunaliella

tertiolecta (Chlorophyceae) under fluctuating irradiance. Photosynthetica 42: 273-281.

Henley, W.J., Hironaka, J.L. Guillou, L., Buchheim, M.A., Buchheim, J.A., Fawley, M.W. &

Fawley, K.P. 2004. Phylogenetic analysis of the “Nannochloris-like” algae and diagnoses of

Picochlorum oklahomensis gen. et sp. nov. (Trebouxiophyceae, Chlorophyta). Phycologia 43:

641-652.

Horton, P. & Ruban, A.V. 2005. Molecular design of the photosystem II light-harvesting

antenna: photosynthesis and photoprotection. J. Exp. Bot. 56: 365-373.

Jahns, P., Depka, B. & Trebst, A. 2000. Xanthophyll cycle mutants from Chlamydomonas

reinhardtii indicate a role for zeaxanthin in the D1 protein turnover. Plant Physiol. Biochem. 38:

371-376.

Jakob, T., Goss, R. & Wilhelm, C. 2001. Unusual pH-dependence of diadinoxanthin de- epoxidase activation causes chlororespiratory induced accumulation of diatoxanthin in the diatom Phaeodactylum tricornutum. J. Plant Physiol. 158: 383-390.

21 Keller, M.D., Selvin, R.C., Claus, W. & Guillard, R.R.L. 1987. Media for the culture of oceanic

ultraphytoplantkon. J. Phycol. 23: 633-638.

Lavaud, J., Rousseau, B., van Gorkom, H. J. & Etienne, A.L. 2002. Influence of the diadinoxanthin pool size on photoprotection in the marine planktonic diatom Phaeodactylum

tricornutum. Plant Physiol. 129: 1398-1406.

Lavaud, J., Rousseau, B. & Etienne, A.L. 2004. General features of photoprotection by energy

dissipation in planktonic diatoms (Bacillariophyceae). J. Phycol. 40: 130-137.

Litchman, E. & Klausmeier, C.A. 2001. Competition of phytoplankton under fluctuating light.

Am. Nat. 157: 170-187.

Lohr, M. & Wilhelm, C. 2001. Xanthophyll synthesis in diatoms: quantification of putative

intermediates and comparison of pigments conversion kinetics with rates constants derived from

a model. Planta 212: 382-391.

Masojidek, J., Kopecky, J., Koblizek, M. & Torzillo, G. 2004. The xanthophyll cycle in green

algae (Chlorophyta): Its role in the photosynthetic apparatus. Plant Biology 6: 342-349.

Masojidek, J., Torzillo, G., Koblizek, J., Bernardini, P., Sacchi, A. & Komenda, J. 1999.

Photoadaptation of two members of the Chlorophyta (Scenedesmus and Chlorella) in laboratory

and outdoor cultures: changes in chlorophyll fluorescence quenching and the xanthophyll cycle.

Planta 209: 126-135.

Moon-van der staay, S. Y., De Watcher, R. & Vaulot, D. 2001. Oceanic 18S rDNA sequences

from picoplankton reveal unsuspected eukaryotic diversity. Nature 409: 607-610.

Niyogi, K. K., Li, X.P., Rosenberg, V., Jung, H.S. 2004. Is PsbS the site of non-photochemical

quenching in photosynthesis? J. Exp. Bot. 56: 375-382.

22 Not, F., Massana, R., Latasa, M., Marie, D., Colson, C., Eikrem, W., Pedrós-Alió, C., Vaulot, D.

& Simon, N. 2005. Late summer community composition and abundance of photosynthetic

picoeukaryotes in Norwegian and Barents Sea. Limnol. Oceanogr. 50: 1677-1686.

Pascal, A. A., Liu, Z., Broess, K., van Oort, B., van Amerongen, H., Wang, C., Horton, P.,

Robert, B., Chang, W. & Ruban, A. 2005. Molecular basis of photoprotection and control of

photosynthetic light-harvesting. Nature 436: 134-138.

Raven, J.A. 1998. The twelfth Tansley lecture. Small is beautiful: the picophytoplankton. Funct.

Ecol. 12: 503.

Ruban, A. & Horton, P. 1999. The xanthophyll cycle modulates the kinetics of non- photochemical energy dissipation in isolated light-harvesting complexes, intact and leaves. Plant Physiol. 119: 531-542.

Schubert, H., Kroon, B.M.A. & Matthijs, H.C.P. 1994. In vivo Manipulation of the xanthophyll

cycle and the role of zeaxanthin in the protection against photodamage in the green alga

Chlorella pyrenoidosa. J. Biol. Chem. 269: 7267-7272.

Sobrino, C., Neale, P.J. & Lubian, L.M. 2005. Interaction of UV radiation and inorganic carbon supply in the inhibition of photosynthesis: Spectral and temporal responses of two marine picoplankton. Photochem. Photobiol. 81: 384-393.

Stomp, M., Huisman, J., de Jongh, F., Veraart, A.J., Gerla, D., Riijkeboer, M., Ibelings, B.W.,

Wollenzien, U.I.A. & Stal, L.J. 2004. Adaptive divergence in pigment composition promotes phytoplankton biodiversity. Nature 432: 104-107.

Strzepek, R.F. & Harrison, P.J. 2004. Photosynthetic architecture differs in coastal and oceanic diatoms. Nature 431: 689-692.

23 Tardy, F. & Havaux, M. 1996. Photosynthesis, chlorophyll fluorescence, light-harvesting system

and photoinhibition resistance of zeaxanthin-accumulating mutant of Arabidopsis thaliana.

Photochem. Photobiol. 34: 87-94.

Tassan, S. & Ferrari, G.M. 1995. An alternative approach to absorption measurements of aquatic particles retained on filters. Limnol. Oceanogr. 40: 1358-1368.

Timmermans, K.R., van der Wagt, B., Veldhuis, M.J.W., Maatman, A. & de Baar, H.J.W. 2005.

Physiological responses of three species of marine pico-phytoplankton to ammonium, phosphate, iron and light limitation. J. Sea Res. 53: 109-120.

Vaulot, D., Le Gall, F., Marie, D., Guillou, L. & Partensky, F. 2004. The Roscoff Culture

Collection (RCC): a collection dedicated to marine picoplankton. Nova Hedwigia 79: 49-70.

Veldhuis, M.J.W., Timmermans, K.R., Croot, P. & van der Wagt, B. 2005. Picophytoplantkon: a comparative study of their biochemical composition and photosynthetic properties. J. Sea Res.

53: 7-24. Villareal, T.A. 2004. Single-cell pulse amplitude modulation fluorescence measurements of the giant diatom Ethmodiscus (Bacillariophyceae). J. Phycol. 40: 1052-1061.

Wilhelm, C., Kramer, P.P. & Lenartz-Weiler, I. 1989. The energy distribution between the photosystems and light-induced changes in the stoichiometry of system I and system II reaction centers in the chlorophyll b-containing alga Mantoniella squamata ().

Photosynth. Res. 20: 221-233.

24

LEGENDS

Fig.1: Progressive light-increase experiment. Temporal variation of (a): light intensity (µmol photons·m-2·s-1), (b): Vx/chl a (diamond), Ax/chl a (circle) and Zx/chl a (triangle) in mol·mol chl

-1 a and (c): DPS (triangle) and effNPQ (circle). Values are means (± SD), n = 3 from independent experiments.

Fig. 2: Progressive light-increase experiment. Mean curves of ETR (a) and NPQ (b) vs PFD obtained from RLC measurements.

Fig. 3: Progressive light-increase experiment. Temporal variation of cellular RALS (a), FALS

(b) and red fluorescence (c) in the changing light culture (diamond) and in the control (circle).

Values are means (± SD) expressed in units relative to the beads, n = 3 from independent experiments.

Fig. 4: High to low PFD experiment. Temporal variation of: (a): Vx/chl a (diamond), Ax/chl a

-1 (circle) and Zx/chl a (triangle) in mol·mol chl a and (b): Fq'/Fm' (operating efficiency of PSII, triangle) and Fv/Fm (maximum photochemical efficiency, circle). Values are means (± SD), n =

3 from independent experiments.

Fig. 5: DTT-added experiment. Temporal variation of (a): effNPQ (DTT-treated culture: (circle; control: triangle), and Vx/chl a (diamond), Ax/chl a (circle) and Zx/chl a (triangle) in mol·mol

25 chl a-1 in the DTT-treated culture (b) and in the control (c). Values are means (± SD) with n = 3 from independent experiments, except the control (n = 1).

Fig. 6: NF-added experiment. Temporal variation of: Vx/chl a (diamond), Ax/chl a (circle) and

Zx/chl a (triangle) in the NF-treated culture (a) and in the control (b); ß-carotene/chl a in the NF- treated culture (c) and in the control (d); effNPQ (e, NF-treated culture: circle; control: triangle)

and Fv/Fm (f, NF-treated culture: circle; control: triangle). All the pigment ratios are in mol·mol

-1 chl a and the values are means (± SD) with n = 3 from independent experiments, except the

control (n =1).

26 ) 1 - 400 s

.

2 a - m

.

s 300 n o t o h p l 200 o m µ ( y t i

s 100 n e t n i t h

g 0 i

L 0 100 200 300 400 Time (min)

0.05 ) 1 - a l

h 0.04

c b l o m

.

0.03 l o m ( o

i 0.02 t a r t n e 0.01 m g i P 0.00 0 100 200 300 400 Time (min)

1.0 c 0.5

0.8 0.4 e

0.6 0.3 f f

S N P P D Q 0.4 0.2

0.2 0.1

0.0 0.0 0 100 200 300 400 Time (min) DIMIER et al. Fig. 1 16 ) 1 - a h .

1 12 - a l h c g

. 8 - e l o m ( 4 R T E

0 0 400 800 1200 1600 PFD (µmol photons m-2 s-1)

2.0

1.6 b

1.2 Q P N 0.8

0.4

0.0 0 400 800 1200 1600 PFD (µmol photon. s m.-2 s-1)

DIMIER et al. Fig. 2 0.036

) 0.032

s a t i n u

e 0.028 v i t a l e r (

0.024 S L A

R 0.020

0.016 0 100 200 300 400 Time (min)

0.20 b ) s

t 0.16 i n u e v i t a 0.12 l e r (

S L

A 0.08 F

0.04 0 100 200 300 400 Time (min)

2.0 ) s t i

n c u

1.9 e v i t a l e

r 1.8 (

e c n e

c 1.7 s e r o u l

f 1.6

a l h c 1.5 0 100 200 300 400 Time (min) DIMIER et al. Fig. 3 0.06 ) 1 - a

l 0.05 a h c

l o 0.04 m

.

l o 0.03 m (

o i t

a 0.02 r

t n e

m 0.01 g i P 0.00 0 40 80 120 160 200 Time (min)

0.68

0.66 ' m F

/ 0.64 ' q F

d 0.62 n

a b

m F

/ 0.60 v F

0.58

0.56 0 40 80 120 160 200 Time (min)

DIMIER et al. Fig. 4 0.12 a

0.08 Q P N

f f e 0.04

0.00 0 40 80 120 Time (min)

0.04 ) 1 -

a b l h c

0.03 l o m

.

l o 0.02 m (

o i t a r

t

n 0.01 e m g i P 0.00 0 40 80 120 Time (min)

0.04 ) 1

- c a l h c 0.03 l o m

.

l o 0.02 m (

o i t a r

t

n 0.01 e m g i P 0.00 0 40 80 120 Time (min) DIMIER et al. Fig. 5 0.04 0.04 ) ) 1 1 - -

a a a l l b h h c c

0.03

0.03 l l o o m m

. .

l l o o 0.02 0.02 m m ( (

o o i i t t a a r r

t t

n 0.01 n 0.01 e e m m g g i i P P 0.00 0.00 0 20 40 60 80 100 0 20 40 60 80 100 Time (min) Time (min)

0.040 0.046 )

1 c ) - 1 d a -

l 0.038 a

h l 0.044 c h

l c

o l o

m 0.036

. m 0.042

l .

o l o m ( m 0.034 ( e 0.040 n e e n t e o t r o a 0.032 r

c a 0.038 - c ß - ß 0.030 0.036 0 20 40 60 80 100 0 20 40 60 80 100 Time (min) Time (min)

0.50 0.75 e f 0.40 0.70

0.65 0.30 Q P m N F

/ 0.60 f v f

e 0.20 F 0.55

0.10 0.50

0.00 0.45 0 40 80 120 0 40 80 120 Time (min) Time (min) DIMIER et al. Fig. 6