17
A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje ´ le ´ Wilson, a Ghada Ajlani, b Jean-Marc Verbavatz, b Imre Vass, c Cheryl A. Kerfeld, d and Diana Kirilovsky a,1 a Unite ´ de Recherche Associe ´ e 2096, Centre National de la Recherche Scientifique, Service de Bioe ´ nerge ´ tique, 91191 Gif sur Yvette, France b Service de Biologie de Fonctionnes Membranaires, De ´ partement de Biologie Joliot-Curie, Commissariat a ` l’Energie Atomique Saclay, 91191 Gif sur Yvette, France c Biological Research Center, H-6726 Szeged, Hungary d Molecular Biology Institute, University of California, Los Angeles, California 90095-1570 Photosynthetic organisms have developed multiple protective mechanisms to survive under high-light conditions. In plants, one of these mechanisms is the thermal dissipation of excitation energy in the membrane-bound chlorophyll antenna of photosys- tem II. The question of whether or not cyanobacteria, the progenitor of the chloroplast, have an equivalent photoprotective mechanism has long been unanswered. Recently, however, evidence was presented for the possible existence of a mechanism dissipating excess absorbed energy in the phycobilisome, the extramembrane antenna of cyanobacteria. Here, we demonstrate that this photoprotective mechanism, characterized by blue light–induced fluorescence quenching, is indeed phycobilisome- related and that a soluble carotenoid binding protein, ORANGE CAROTENOID PROTEIN (OCP), encoded by the slr1963 gene in Synechocystis PCC 6803, plays an essential role in this process. Blue light is unable to quench fluorescence in the absence of phycobilisomes or OCP. The fluorescence quenching is not DpH-dependent, and it can be induced in the absence of the reaction center II or the chlorophyll antenna, CP43 and CP47. Our data suggest that OCP, which strongly interacts with the thylakoids, acts as both the photoreceptor and the mediator of the reduction of the amount of energy transferred from the phycobilisomes to the photosystems. These are novel roles for a soluble carotenoid protein. INTRODUCTION Carotenoids, the most widely occurring pigments in nature, are essential to microbial, animal, and plant life. These lipophilic macromolecules play diverse roles, functioning as colorants, precursors of visual pigments, and antioxidants. In plant and algal photosynthesis, carotenoids are known to have a dual function in chlorophyll–membrane complexes: harvesting light and photoprotection. The photoprotective mechanisms include (1) reducing the amount of energy funneled to photochemical reaction centers (by screening or thermal energy dissipation), (2) quenching triplet chlorophyll to prevent the formation of singlet oxygen, and (3) directly quenching singlet oxygen. In cyanobacteria, carotenoids are also associated with pro- teins devoid of chlorophyll. These water-soluble carotenoid bind- ing proteins were first described by Holt and Krogmann (1981) and later found in several different cyanobacterial species (re- viewed in Kerfeld, 2004a, 2004b). The soluble ORANGE CA- ROTENOID PROTEIN (OCP), a 35-kD protein that contains a single noncovalently bound carotenoid (Holt and Krogmann, 1981; Wu and Krogmann, 1997; Kerfeld, 2004a, 2004b), is the best characterized of these proteins. In Synechocystis PCC 6803, the OCP is the product of the slr1963 open reading frame (Wu and Krogmann, 1997). Highly conserved homologs of the OCP are found in the genomes of all cyanobacteria for which genomic data are available, with the exception of the prochlor- ococci (Kerfeld, 2004a, 2004b). The structure of the Arthrospira maxima OCP has been de- termined to 2.1 A ˚ resolution (Kerfeld et al., 2003). The OCP consists of two domains: an all a-helical N-terminal domain and a mixed a/b C-terminal domain. The embedded carotenoid, a 39-hydroxyequinenone, has an all-trans configuration and spans both protein domains. The protein has a large effect on the spectroscopic characteristics of the carotenoid. In organic sol- vents, the 39-hydroxyequinenone is yellow (l max ¼ 450 nm), but in the OCP, it appears orange (l max ¼ 465 and 495 nm) (Kerfeld et al., 2003; Polivka et al., 2005). In the course of OCP purifica- tion, a red carotenoid protein (RCP) is also isolated (Wu and Krogmann, 1997; Kerfeld, 2004a, 2004b). N-terminal sequencing and mass spectrometry analysis indicated that this is a 16-kD proteolytic fragment of the OCP. The proteolysis removes the entire C-terminal domain, which, without concomitant structural change, would expose nearly half of the carotenoid to solvent. A role for the OCP under stress conditions has been proposed. Indeed, microarray studies indicate that the OCP transcript levels increase by >600% upon transfer to high light (Hihara et al., 2001). Several functions for the OCP in photoprotection have been suggested based on its structure and on in vitro 1 To whom correspondence should be addressed. E-mail diana. [email protected]; fax 33-1-69088717. The authors responsible for distribution of materials integral to the findings presented in this article in accordance with the policy described in the Instructions for Authors (www.plantcell.org) are: Ghada Ajlani ([email protected]) and Diana Kirilovsky ([email protected]). Article, publication date, and citation information can be found at www.plantcell.org/cgi/doi/10.1105/tpc.105.040121. The Plant Cell, Vol. 18, 992–1007, April 2006, www.plantcell.org ª 2006 American Society of Plant Biologists

A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

A Soluble Carotenoid Protein Involved inPhycobilisome-Related Energy Dissipationin Cyanobacteria

Adjele Wilson,a Ghada Ajlani,b Jean-Marc Verbavatz,b Imre Vass,c Cheryl A. Kerfeld,d and Diana Kirilovskya,1

a Unite de Recherche Associee 2096, Centre National de la Recherche Scientifique, Service de Bioenergetique, 91191

Gif sur Yvette, Franceb Service de Biologie de Fonctionnes Membranaires, Departement de Biologie Joliot-Curie, Commissariat a

l’Energie Atomique Saclay, 91191 Gif sur Yvette, Francec Biological Research Center, H-6726 Szeged, Hungaryd Molecular Biology Institute, University of California, Los Angeles, California 90095-1570

Photosynthetic organisms have developed multiple protective mechanisms to survive under high-light conditions. In plants, one

of these mechanisms is the thermal dissipation of excitation energy in the membrane-bound chlorophyll antenna of photosys-

tem II. The question of whether or not cyanobacteria, the progenitor of the chloroplast, have an equivalent photoprotective

mechanism has long been unanswered. Recently, however, evidence was presented for the possible existence of a mechanism

dissipating excess absorbed energy in the phycobilisome, the extramembrane antenna of cyanobacteria. Here, we demonstrate

that this photoprotective mechanism, characterized by blue light–induced fluorescence quenching, is indeed phycobilisome-

related and that a soluble carotenoid binding protein, ORANGE CAROTENOID PROTEIN (OCP), encoded by the slr1963 gene in

Synechocystis PCC 6803, plays an essential role in this process. Blue light is unable to quench fluorescence in the absence of

phycobilisomes or OCP. The fluorescence quenching is notDpH-dependent, and it can be induced in the absence of the reaction

center II or the chlorophyll antenna, CP43 and CP47. Our data suggest that OCP, which strongly interacts with the thylakoids,

acts as both the photoreceptor and the mediator of the reduction of the amount of energy transferred from the phycobilisomes to

the photosystems. These are novel roles for a soluble carotenoid protein.

INTRODUCTION

Carotenoids, the most widely occurring pigments in nature, are

essential to microbial, animal, and plant life. These lipophilic

macromolecules play diverse roles, functioning as colorants,

precursors of visual pigments, and antioxidants. In plant and

algal photosynthesis, carotenoids are known to have a dual

function in chlorophyll–membrane complexes: harvesting light

and photoprotection. The photoprotective mechanisms include

(1) reducing the amount of energy funneled to photochemical

reaction centers (by screening or thermal energy dissipation), (2)

quenching triplet chlorophyll to prevent the formation of singlet

oxygen, and (3) directly quenching singlet oxygen.

In cyanobacteria, carotenoids are also associated with pro-

teins devoid of chlorophyll. These water-soluble carotenoid bind-

ing proteins were first described by Holt and Krogmann (1981)

and later found in several different cyanobacterial species (re-

viewed in Kerfeld, 2004a, 2004b). The soluble ORANGE CA-

ROTENOID PROTEIN (OCP), a 35-kD protein that contains a

single noncovalently bound carotenoid (Holt and Krogmann,

1981; Wu and Krogmann, 1997; Kerfeld, 2004a, 2004b), is the

best characterized of these proteins. In Synechocystis PCC

6803, the OCP is the product of the slr1963 open reading frame

(Wu and Krogmann, 1997). Highly conserved homologs of the

OCP are found in the genomes of all cyanobacteria for which

genomic data are available, with the exception of the prochlor-

ococci (Kerfeld, 2004a, 2004b).

The structure of the Arthrospira maxima OCP has been de-

termined to 2.1 A resolution (Kerfeld et al., 2003). The OCP

consists of two domains: an alla-helical N-terminal domain and a

mixed a/b C-terminal domain. The embedded carotenoid, a

39-hydroxyequinenone, has an all-trans configuration and spans

both protein domains. The protein has a large effect on the

spectroscopic characteristics of the carotenoid. In organic sol-

vents, the 39-hydroxyequinenone is yellow (lmax ¼ 450 nm), but

in the OCP, it appears orange (lmax ¼ 465 and 495 nm) (Kerfeld

et al., 2003; Polivka et al., 2005). In the course of OCP purifica-

tion, a red carotenoid protein (RCP) is also isolated (Wu and

Krogmann, 1997; Kerfeld, 2004a, 2004b). N-terminal sequencing

and mass spectrometry analysis indicated that this is a 16-kD

proteolytic fragment of the OCP. The proteolysis removes the

entire C-terminal domain, which, without concomitant structural

change, would expose nearly half of the carotenoid to solvent.

A role for the OCP under stress conditions has been proposed.

Indeed, microarray studies indicate that the OCP transcript

levels increase by >600% upon transfer to high light (Hihara

et al., 2001). Several functions for the OCP in photoprotection

have been suggested based on its structure and on in vitro

1 To whom correspondence should be addressed. E-mail [email protected]; fax 33-1-69088717.The authors responsible for distribution of materials integral to thefindings presented in this article in accordance with the policy describedin the Instructions for Authors (www.plantcell.org) are: Ghada Ajlani([email protected]) and Diana Kirilovsky ([email protected]).Article, publication date, and citation information can be found atwww.plantcell.org/cgi/doi/10.1105/tpc.105.040121.

The Plant Cell, Vol. 18, 992–1007, April 2006, www.plantcell.org ª 2006 American Society of Plant Biologists

Page 2: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

experiments, such as a quencher of singlet oxygen, a carotenoid

transport protein, and a light screen (Kerfeld, 2004a, 2004b).

By harvesting solar energy and converting it into chemical

energy, plants, algae, and cyanobacteria provide food and oxygen

that are essential for life on earth. However, excess light can be

lethal for photosynthetic organisms, because harmful reactive

oxygen species are generated in the photochemical reaction

centers when energy absorption exceeds the rate of carbon

fixation. To survive, photosynthetic organisms have evolved sev-

eral protective processes. In plants and algae, dissipation of the

excess absorbed energy as heat in the light-harvesting chlorophyll

antenna (LHCII) of photosystem II (PSII) decreases the amount of

energy funneled to the photochemical centers (for reviews, see

Demmig-Adams, 1990; Horton et al., 1996; Niyogi, 1999; Muller

et al., 2001). A decrease of PSII-related fluorescence emission,

known as nonphotochemical quenching (NPQ; more specifically

qE), is observed when this process is triggered by acidification of

the thylakoid lumen under saturating light conditions. A decrease

in thylakoid lumen pH activates the formation of the carotenoid

zeaxanthin from violaxanthin via the xanthophyll cycle (Yamamoto,

1979; Gilmore and Yamamoto, 1993). Low pH drives the proton-

ation of PsbS, a PSII subunit that belongs to the LHC superfamily

(Li et al., 2000), inducing conformational changes in the LHCII

(Ruban et al., 1992).

Cyanobacteria do not have the integral membrane chlorophyll-

containing light-harvesting complex, LHCII. Instead, light is cap-

tured by a membrane extrinsic complex, the phycobilisome,

which is attached to the outer surface of thylakoid membranes

(Gantt and Conti, 1966). These large complexes consist of

phycobiliproteins that covalently bind bilin pigments and linker

peptides that are required for the organization of the phycobili-

some and for tuning the physical characteristics of the pigments

(for reviews, see Glazer, 1984; MacColl, 1998; Tandeau de

Marsac, 2003; Adir, 2005). Phycobilisomes are composed of a

core from which rods (usually six) radiate. The major core protein

is allophycocyanin, whereas the rods contain phycocyanin and in

some species phycoerythrin or phycoerythrocyanin (in the distal

end of the rod). The phycobilisomes are bound to the thylakoids

via the core–membrane linker protein, Lcm, which also serves as

the terminal energy acceptor of the phycobilisomes (Redlinger

and Gantt, 1982). Harvested light energy is transferred from Lcm

to the chlorophylls of the inner chlorophyll antenna, CP43 and

CP47 (containing chlorophyll a and carotenoids), and to reaction

center II. Phycobilisome can also transfer energy to photosystem

I (PSI) (Mullineaux, 1992; Rakhimberdieva et al., 2001).

Cyanobacteria have two well-characterized mechanisms that

are associated with fluorescence quenching: state transitions

and photoinhibition. In photoinhibition, high light intensities in-

duce PSII fluorescence quenching and the irreversible inactiva-

tion of PSII caused by damage and degradation of the D1 protein,

an essential constituent of PSII (for reviews, see Prasil et al.,

1992; Aro et al., 1993; Melis, 1999). Recovery of fluorescence

and oxygen-evolving activity requires the replacement of the

damaged D1 protein. State transitions regulate the redistribution

of energy between PSII and PSI; they are induced by changes in

the redox state of the plastoquinone pool (for reviews, see Allen,

1992; van Thor et al., 1998; Wollman, 2001). Exposure of pho-

tosynthetic organisms to light absorbed predominantly by PSII

causes a relative decrease of the PSII fluorescence yield. Con-

versely, illumination with light absorbed preferentially by PSI

induces a relative increase of the fluorescence yield.

It has long been assumed that cyanobacteria do not use an

antenna-related NPQ mechanism to decrease the amount of en-

ergy funneled to reaction center II (Campbell et al., 1998). Several

recent experiments refute this view. For example, an NPQ mech-

anism mediated by the Iron stress–induced A protein (IsiA) that

belongs to the core complex family of chlorophyll binding pro-

teins has been described. This protein, which is induced under

iron starvation (Laudenbach and Straus, 1988; Burnap et al.,

1993) and other stress conditions (Jeanjean et al., 2003; Yousef

et al., 2003; Havaux et al., 2005), encircles the PSI reaction center

(Bibby et al., 2001; Boekema et al., 2001). After prolonged iron

starvation, empty rings of IsiA (without PSI) are also detected

(Yeremenko et al., 2004). It was proposed that this unconnected

IsiA is involved in photoprotection (Park et al., 1999; Sandstrom

et al., 2001; Yeremenko et al., 2004; Bailey et al., 2005; Ihalainen

et al., 2005). Indeed, it was demonstrated that these IsiA aggre-

gates are in a strongly quenched state, suggesting that they are

responsible for the thermal dissipation of absorbed energy

(Ihalainen et al., 2005).

In addition, results revealing the existence of a distinct blue

light–induced NPQ mechanism, proposed to be associated with

the phycobilisome, were first described in 2000 (El Bissati et al.,

2000). Subsequently, spectral and kinetic data were presented

suggesting that blue light–activated carotenoids induce the quench-

ing of phycobilisome fluorescence emission (Rakhimberdieva et al.,

2004).

Here, we report that the carotenoid protein OCP is essential for

this phycobilisome-associated NPQ mechanism in Synechocys-

tisPCC 6803. The OCP appears to act as both the photoreceptor

and the mediator of a photoprotective energy dissipation mech-

anism, which decreases the amount of energy arriving at both

photosystems from the phycobilisome.

RESULTS

Construction of OCP Mutants

To determine the function and cellular location of the OCP, we

constructed three mutants ofSynechocystisPCC 6803: a mutant

lacking the OCP protein, in which the ocp gene (slr1963) was

inactivated (DOCP); a mutant containing an OCP–green fluores-

cent protein fusion under the control of the ocp promoter (OCP-

GFP); and a mutant in which the gene slr1964 was inactivated

(DSlr1964), used as the control strain. An antibiotic resistance

cassette was introduced into the HincII site in slr1963 (DOCP) or

into the ClaI site in slr1964 (DSlr1964 and OCP-GFP) by double

homologous recombination (Figure 1A). To confirm the insertion

sites and the complete segregation of the mutants, PCR analysis

was performed. The amplification of the genomic region con-

taining slr1963, isolated from DOCP, DSlr1964, and OCP-GFP,

with the oligonucleotides car1 and car6 gave fragments of 4.0 kb

(DOCP and DSlr1964) and 4.8 kb (OCP-GFP) (Figure 1B). No

traces of the 2-kb fragment, observed in the wild type, were

detected in the mutants, indicating complete segregation. The

Carotenoids and NPQ in Cyanobacteria 993

Page 3: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

correct position of the antibiotic cassette in DOCP and DSlr1964

was controlled by PCR amplification of slr1963 by the oligonu-

cleotides car1 and car2 (Figure 1B). In the DOCP mutant, the

3-kb PCR fragment indicated that slr1963 was interrupted by the

antibiotic cassette, whereas this gene was not interrupted in

DSlr1964 (0.9-kb PCR fragment). To check the position of the

antibiotic cassette in this mutant, the 4-kb PCR fragment ob-

tained by amplification with the oligonucleotides car1 and car6

was digested by the restriction enzyme HincII. As expected,

digestion of the DSlr1964 PCR fragment gave four fragments of

2 kb, 800 bp, 750 bp, and 440 bp, confirming that in this mutant

slr1964 is interrupted by the antibiotic cassette (Figure 1C).

Phenotype Conferred by DOCP under

High Light Intensities

We compared the behavior of DOCP and wild-type cells under

high-intensity white light. When fluorescence was measured with

a PAM fluorometer during saturating light illumination, a faster

quenching of fluorescence was observed in wild-type cells than

in DOCP cells. Figure 2 shows room temperature fluorescence

traces in Synechocystis wild-type and DOCP cells illuminated by

different light intensities. Dark-adapted wild-type and DOCP

cells presented a characteristic low level of maximal fluores-

cence (Fm). Upon illumination with low intensities of white light, a

maximal level of Fm was reached in both strains (Figure 2A). This

increase of fluorescence is related to a state 1 transition induced

by the oxidation of the plastoquinone pool upon illumination.

Subsequently, exposure of cells to saturating white light inten-

sities induced fluorescence quenching. A very fast quenching of

the maximal fluorescence (Fm9) was observed in wild-type cells,

especially during the first minutes of high light illumination. The

steady state (Fs) and minimal (Fo) fluorescence levels also de-

creased (Figure 2). In DOCP cells, the quenching of Fm9 was

slower than in the wild type, and no quenching of Fo was

detected (Figure 2).

Figure 1. Mutant Construction and Segregation in the Synechocystis PCC 6803 Genome.

(A) Gene arrangement of the slr1963 (encoding the OCP) and slr1964 genes. The positions of the oligonucleotides used for amplification and the

restriction sites are indicated. In the DOCP strain, the slr1963 gene was disrupted by insertion of the spectinomycin and streptomycin (Sp/Sm)

resistance cassette. In the Dslr1964 strain, the Sp/Sm resistance cassette interrupted slr1964. In the OCP-GFP mutant, the GFP gene was fused to the

C terminus of the OCP and the slr1964 gene was disrupted by the Sp/Sm resistance gene.

(B) Amplification of genomic Synechocystis DNA from the OCP-GFP mutant (lane 1), the wild type (lane 2), the DOCP mutant (lane 3), and the DSlr1964

mutant (lane 4) using car1 and car6 as primers. Lanes 5 and 6 show the PCR fragments obtained by amplification of slr1963 from the DSlr1964 strain

(lane 5) and from the DOCP mutant (lane 6) with car1 and car2 as primers. MW, 1-kb DNA ladder.

(C) Digestion of the 2- and 4-kb PCR fragments obtained using wild-type (lane 1), DOCP (lane 2), and DSlr1964 (lane 3) DNA as templates by the

restriction enzyme HincII.

994 The Plant Cell

Page 4: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

When dark-adapted cells were exposed directly to high inten-

sities of white light, a slight increase of Fm9 was initially detected,

probably attributable to oxidation of the plastoquinone pool and

partial state 1 transition, followed by fluorescence quenching

(Figures 2B and 2C). The quenching generated in DOCP cells

was not reversible in darkness (Figure 2C). By contrast, the large

Fm9 and Fo decrease observed in wild-type cells was almost

completely reversible in the dark in the presence of chloram-

phenicol, a protein synthesis inhibitor; this finding strongly

suggested that the observed quenching was not related to D1

protein damage, which is associated with photoinhibition (Figure

2B). These results implied that, in wild-type cells, an induction of

a reversible antenna-associated NPQ occurs in high intensities

of white light that could protect PSII from photoinhibition. By

contrast, in DOCP, the decrease of maximal fluorescence ap-

peared to correlate with photoinhibition and not with NPQ.

Indeed, when the sensitivity of DOCP cells to high light

intensities was tested by following the decrease of oxygen

evolution during saturating illumination, oxygen evolution activity

decreased faster in the mutant than in the wild type, indicating

that DOCP was more sensitive to high intensities of white light

(Figure 3A). Moreover, comparison of light saturation curves of

PSII activity strongly suggested that the effective PSII antenna

size was smaller in cells that were in the quenched state. Figure

3B shows light saturation curves for oxygen-evolving activity in

wild-type cells and in 10-min photoinhibited cells. Light satura-

tion occurred at lower light intensities in wild-type control cells

than in photoinhibited cells.

Blue-Green Light–Induced Quenching and the OCP

The unexpected fluorescence characteristics of the DOCP mu-

tant prompted us to investigate the possible relationship be-

tween the OCP and the blue-green light–induced NPQ reported

previously (El Bissati et al., 2000). Blue-green light, depending on

its intensity, may be involved in state transitions or in NPQ (El

Bissati et al., 2000). Illumination of dark-adapted wild-type and

DOCP cells by dim blue-green light (which preferentially excites

PSI) increased fluorescence levels, indicating a transition to state

I induced by the oxidation of the plastoquinone pool (Figure 4).

Subsequently, exposure of dim light–adapted cells to high blue-

green light intensities induced a quenching of all fluorescence

levels in the wild type but not in DOCP (Figures 4A and 4B), even

under very high light intensities. Similar characterization of the

mutant in which the slr1964 gene, located immediately down-

stream of the ocp gene, was disrupted (DSlr1964) confirmed that

the inhibition of NPQ is attributable to the absence of the OCP

and not to the lack of transcription of slr1964 or to other polar

effects of the mutation (Figure 4C).

Transition to state II was not affected in DOCP. Illumination by

orange light (which preferentially excites PSII) of wild-type andFigure 2. Changes in Fluorescence Levels Induced by Different Inten-

sities of White Light Measured in a PAM Fluorometer.

(A) Dark-adapted wild-type and DOCP cells (3 mg chlorophyll/mL) were

successively illuminated with white light at 50 mmol�m�2�s�1 followed by

white light at 1500 mmol�m�2�s�1. Saturating pulses (1-s duration, 2000

mmol�m�2�s�1) were applied to measure Fm and Fm9 levels in darkness

and in dim light, respectively. Under high-intensity light illumination, all

centers were closed (Fs ¼ Fm9) and saturating flashes were not applied.

(B) and (C) Dark-adapted cells of the wild type (B) and DOCP (C)

were illuminated directly with high white light (1500 mmol�m�2�s�1 ) for

2 min and then incubated in darkness. During dark incubation, sat-

urating pulses were applied to measure Fm levels. Chloramphenicol was

present during all experiments.

Carotenoids and NPQ in Cyanobacteria 995

Page 5: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

DOCP cells in state I (adapted to low intensities of blue-green light)

induced a decrease of the Fm9 level, characteristic of state II

transition, in both strains (Figures 4D and 4E). Collectively, these

data demonstrate that absence of the OCP inhibits the blue-green

light–induced NPQ and does not affect state transitions.

Is the OCP-Associated NPQ Related to Phycobilisomes or

to Chlorophyll Antennae?

In cyanobacteria, fluorescence detected at room temperature is

emitted from both chlorophyll and phycobiliproteins (Campbell

et al., 1998). In a PAM fluorometer, the measuring light has a

maximum excitation at 650 nm and the fluorescence is detected

at wavelengths of >700 nm. In cyanobacteria, which lack chlo-

rophyll b, most of the measuring light is absorbed by the

phycobilisome. This is confirmed by the low levels of fluores-

cence emitted by mutants lacking phycobilisome (El Bissati and

Kirilovsky, 2001). Thus, a decrease in the fluorescence levels

observed in a PAM fluorometer could be the result of a diminution

of the phycobilisome emission or of the chlorophyll antenna emis-

sion or of a decrease in energy transfer from the phycobilisome to

PSII. It is important to distinguish between these different possible

sources of OCP-dependent NPQ. Fluorescence spectra strongly

suggested that OCP was involved in a phycobilisome-associated

NPQ (Figure 5). At room temperature, when wild-type cells exhibit-

ing blue-green light NPQ were excited at 600 nm (light principally

absorbed by the phycobilisome), the fluorescence band with a

maximum at 660 nm (phycobilisome-related) was smaller than that

of dark-adapted wild-type control cells (Figure 5A). By contrast, the

fluorescence emission spectra were identical in DOCP cells adap-

ted either to darkness or to high intensities of blue-green light

(Figure 5C). However, when wild-type andDOCP cells were excited

by 430-nm light (principally absorbed by chlorophyll), the emission

band with a maximum at 685 nm (chlorophyll-related) increased in

cells in the quenched state compared with that in dark-adapted

control cells (Figures 5B and 5D). This chlorophyll fluorescence

increase is associated with a state II–to–state I transition upon

illumination of dark-adapted cells. Figures 5E and 5F show the 77K

fluorescence emission spectra of wild-type cells illuminated for

5 min with low (control state) or high (quenched state) intensities of

blue light. In the fluorescence spectrum generated by 600-nm

excitation, emission bands related to phycocyanin (650 nm),

allophycocyanin (660 nm), the phycobilisome terminal emitter

(685 nm), PSII (685 and 695 nm), and PSI (725 nm) were observed.

The intensity of each of these bands decreased in wild-type cells in

the quenched state (Figure 5E). By contrast, the fluorescence

spectrum generated in the quenched cells by excitation at 430 nm,

in which only PSII- and PSI-related emissions were observed, was

identical to that generated in control cells (Figure 5F). Thus, the

observed fluorescence decrease observed in the PAM fluorometer

was attributable to the quenching of the phycobilisome fluores-

cence emission and a concomitant decrease in the energy transfer

from the phycobilisome to the photosystems.

OCP-Associated NPQ in Phycobilisome and

PSII-Deficient Mutants

To substantiate the hypothesis that the OCP is specifically

involved in a phycobilisome-dependent NPQ mechanism, addi-

tional mutants were characterized: (1) lacking phycocyanin (CK,

Dcpc operon) (B. Ughy and G. Ajlani, unpublished data); (2)

lacking phycocyanin and OCP (CK-DOCP) (this work); (3) lacking

allophycocyanin (DAB, DapcAB) (Ajlani et al., 1995); (4) lacking

phycobilisome (PAL, DapcAB, DapcE, PC�) (Ajlani and Vernotte,

1998); (5) lacking PSII but retaining intact phycobilisome (DCP47

[this work] and DCP43-DpsbD; psbDI/C/DII� [Yu and Vermaas,

1990]); (6) lacking both the OCP and PSII (DCP47-DOCP) (this

work); and (7) lacking IsiA (DIsiA) (this work). The construction of

the new mutants used in this study is described in Methods.

Figure 3. Photoinhibition: Effect of High Intensities of White Light on

Oxygen Evolution in Wild-Type and DOCP Cells.

(A) Decrease of oxygen evolution induced in wild-type (circles) and

DOCP (squares) cells (10 mg chlorophyll/mL) by exposure to white light

(three lamps of 1000 mmol�m�2�s�1 each). Error bars represent SE from

three independent experiments. One hundred percent of oxygen-

evolving activity was 207 6 5 mmol O2�h�1�mg chlorophyll�1 in the wild

type and 213 6 6 mmol O2�h�1�mg chlorophyll�1 in the DOCP mutant.

(B) Saturation light curves of oxygen evolution activity in control wild-

type cells (closed circles) and in 10-min photoinhibited wild-type cells

(open circles). The photoinhibited cells were incubated on ice until

measurement. Error bars represent SE from four independent experi-

ments. One hundred percent of the oxygen-evolving activity was 205 6

5 mmol O2�h�1�mg chlorophyll�1 in control cells and 160 6 5 mmol

O2�h�1�mg chlorophyll�1 in photoinhibited cells.

996 The Plant Cell

Page 6: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

In CK (lacking phycocyanin), the PSII antenna is composed of

the phycobilisome core containing allophycocyanin. In this mu-

tant, blue-green light still induced a reversible fluorescence

decrease (Figure 6A). The NPQ was smaller and the recovery

faster than in the wild type. This slight fluorescence quenching

was completely inhibited in the absence of the OCP (Figure 6B).

In PAL, devoid of phycobiliproteins, and in DAB, lacking allophy-

cocyanin, high intensities of blue-green light were unable to

induce fluorescence quenching (Figures 6C and 6D). By contrast,

in PAL (El Bissati and Kirilovsky, 2001) and in DAB (G. Ajlani and

C. Vernotte, unpublished data), changes in the redox state of the

plastoquinone pool induced state transitions. These results

suggest that the OCP may interact with the core of the phyco-

bilisomes to induce fluorescence quenching and support a role

for the OCP in a phycobilisome-associated NPQ.

In the DCP47 mutant, which lacks PSII reaction centers and

contains only traces of unconnected CP43 (Eaton-Rye and

Vermaas, 1991), fluorescence at room temperature is almost

entirely emitted by the phycobilisomes. As expected, because

of the lack of active PSII, no variable fluorescence was detected

in this mutant. In dark-adapted DCP47 cells, blue-green light

induced a large amount of fluorescence quenching that recov-

ered in the dark in the presence of a protein synthesis inhibitor

(Figures 7A and 7B). The quenching of the 660-nm emission

fluorescence band was even larger than that observed in wild-

type cells (Figure 7A).

The fluorescence quenching increased with blue-green light

intensity in wild-type and DCP47 cells (Figures 7C and 7D). Blue-

green light up to 160mmol�m�2�s�1 induced the transition to state

I in dark-adapted, low-fluorescence wild-type cells (data not

shown) and did not induce any quenching in the DCP47 mutant

(Figure 7D). Above this intensity, fluorescence quenching was

induced, and it increased with light intensity (Figures 7C and 7D).

Green light also induced fluorescence quenching with high

efficiency in dark-adapted and low-light-adapted cells (Figures

7C, 7E, and 7F). By contrast, orange-red light did not induce any

NPQ (Figures 7E and 7F). Similar results were obtained with the

DCP43-DPsbD mutant, lacking both interior chlorophyll anten-

nas (CP43 and CP47) and PSII (Yu and Vermaas, 1990) (data not

shown). In the DCP47-DOCP double mutant, lacking both the

OCP and the PSII core, even very high intensities of blue-green

light were unable to induce fluorescence quenching (Figure 7D).

We also examined the relationship between the OCP-associated

NPQ and the NPQ mediated by the IsiA chlorophyll binding

Figure 4. Effect of the OCP on the Development of Blue-Green Light–

Induced NPQ and on State Transitions.

(A) to (C) Measurements of fluorescence yield by a PAM fluorometer in

dark-adapted wild-type (A), DOCP (B), and DSlr1964 (C) cells illuminated

successively with low-intensity blue-green light (400 to 550 nm,

80 mmol�m�2�s�1 ) and high-intensity blue-green light (740 mmol�m�2�s�1).

(D) and (E) Measurements of fluorescence yield by a PAM fluorometer in

dark-adapted wild-type (D) and DOCP (E) cells illuminated successively

with low-intensity blue-green light (400 to 550 nm, 80 mmol�m�2�s�1) and

orange light (600 to 650 nm, 20 mmol�m�2�s�1). Blue-green illumination

induced the state 1 transition (high fluorescence state), and then orange

illumination induced the state 2 transition (low fluorescence state).

Saturating pulses separated by 30 s were applied to assess Fm9.

Carotenoids and NPQ in Cyanobacteria 997

Page 7: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

Figure 5. Involvement of Phycobilisomes in Thermal Energy Dissipation.

(A) to (D) Room temperature fluorescence spectra of dark-adapted wild-type and DOCP cells (solid lines) and after 5 min of high-intensity blue-green

light illumination (dashed lines).

(E) and (F) 77K fluorescence spectra of wild-type cells after 5 min of illumination with low-intensity (gray lines) and high-intensity (dashed lines) blue-

green light. These spectra were normalized to fluorescence emitted by known concentrations of phycoerythrin (PE; excitation, 600 nm) or fluorescein

(excitation, 430 nm) added to the samples just before recording the spectra. APC, allophycocyanin; PC, phycocyanin.

Each spectrum shown is the mean of 12 spectra from three independent experiments (mean of four spectra per experiment). Excitation was performed

at 600 nm ([A], [C], and [E]) and at 430 nm ([B], [D], and [F]).

998 The Plant Cell

Page 8: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

protein. Figure 7H shows that in a mutant lacking the IsiA protein,

blue-green light induced NPQ similar to that induced in the wild

type, demonstrating that this protein is not involved in the OCP-

mediated process.

Thus, the blue-green light–induced NPQ associated with OCP

can occur in the absence of the chlorophyll core antenna (CP43,

CP47, and IsiA) as well as in the absence of reaction center II.

Moreover, the kinetics of Fm9 quenching induced by strong blue-

green light were similar in the absence or presence of DCMU,

a PSII electron transport inhibitor, or nigericin, an uncoupler

(Figure 7G). These chemicals had no observable effect on the

recovery kinetics (Figure 7G). This finding indicates that the

cyanobacterial NPQ induced by blue-green light was not related

to the redox state of the plastoquinone pool or to DpH. Instead, it

seems to be related to the energy collected by the OCP, which

absorbs in the blue and green regions of the visible spectrum but

not in the orange-red region (Kerfeld, 2004b; Polivka et al., 2005).

The OCP Is Associated with Thylakoids and Phycobilisomes

The quenching phenomenon we observed suggests that the

OCP is associated with phycobilisomes and/or thylakoids. To

identify its cellular location, a Synechocystis PCC 6803 strain

containing an OCP-GFP fusion protein was prepared (Figure 1).

Panel 1 in Figure 8A shows that in the OCP-GFP mutant, green

fluorescence appeared to be distributed throughout the cell. It

is not at the periphery, indicating that OCP-GFP does not pre-

ferentially accumulate in the cell wall or in the cytoplasmic

membrane. No significant green fluorescence was observed in

wild-type cells (Figure 8A, row 2).

Cells containing the OCP-GFP fusion protein were also studied

by immunogold labeling with analysis by electron microscopy.

Visualization of the anti-GFP antibodies showed that most of the

OCP-GFP fusion proteins were localized in the interthylakoid

cytoplasmic region, the phycobilisome side of the membranes

(Figures 8B and 8C, panel 1). The density of the gold particles in

the cytoplasmic interthylakoid region (which represents 65% of

the cellular surface) was significantly higher (average ¼ 46.33 6

3.93 particles/mm2) than in the rest of the cytoplasm (average ¼17.06 6 1.78 particles/mm2; P < 10�6). Thus, 82% of the gold

particles were in the thylakoid region. In the absence of the

primary antibody, no labeling was observed (Figure 8C, panel 2).

Evidence for the interaction between the OCP and the thyla-

koids was corroborated by the presence of the OCP-GFP fusion

protein in phycobilisome-associated (MP) and phycobilisome-

free (M) membrane preparations (Figure 9). The membrane

fractions were obtained by centrifugation of cells broken in a

phosphate/citrate buffer (to obtain MP) or in a MES buffer (to

obtain M). The GFP fluorescence emission spectra of different

Figure 6. Strong Blue-Green Light Effect in Different Phycobilisome

Mutants.

Measurements of fluorescence yield by a PAM fluorometer in dark-

adapted CK ([A]; phycocyanin-deficient mutant), CK-DOCP (B), DAB ([C];

allophycocyanin-deficient mutant), and PAL ([D]; phycobilisome-deficient

mutant) cells illuminated successively with low-intensity blue-green light

(400 to 550 nm, 80 mmol�m�2�s�1; for PAL, 300 mmol�m�2�s�1) and high-

intensity blue-green light (740 mmol�m�2�s�1; for PAL, 1700 mmol�m�2�s�1).

Carotenoids and NPQ in Cyanobacteria 999

Page 9: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

Figure 7. Blue-Green Light–Induced OCP-Related Fluorescence Quenching in DCP47 and DIsiA Mutant Cells.

(A) Room temperature fluorescence spectra of control (solid line) and quenched (dashed line) DCP47 cells. Excitation was performed at 600 nm.

(B) Fluorescence level changes in dark-adapted cells of the DCP47 mutant during successive illumination by high blue-green light followed by dark

incubation for fluorescence recovery.

1000 The Plant Cell

Page 10: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

fractions are shown in Figures 9A and 9B. When the soluble

proteins were separated from the membrane fractions by cen-

trifugation, most of the OCP-GFP coprecipitated with the mem-

brane fractions (Figure 9A). The GFP fluorescence emission in

the M fraction was slightly smaller than that in the MP fraction,

whereas in both supernatants, the GFP emission was very small,

being smaller in the phosphate/citrate than in the MES superna-

tant. Isolation of the MP and M fractions in a sucrose gradient

gave similar results (data not shown). When the MP fraction was

resuspended in a low-phosphate buffer to dissociate the phy-

cobilisomes from the membrane, most of the OCP-GFP remained

attached to the membrane (Figure 9B). By contrast, allophyco-

cyanin and phycocyanin emissions were prominent in the MP

fraction but small in the M fraction (data not shown). These

results were supported by protein separation by SDS-PAGE and

detection of the OCP-GFP fusion protein by an anti-GFP anti-

body. This antibody reacted with a 65-kD polypeptide corre-

sponding to the expected molecular mass of the whole fusion

protein. The antibody binding was greater in the MP and M

fractions (Figure 9D) than in the supernatant fractions (Figure 9C).

Thus, our preliminary localization results suggest that there is a

relatively strong interaction between the OCP and the thylakoids,

with almost all of the OCP-GFP in the MP fraction.

The OCP was very sensitive to a membrane protease activated

during membrane solubilization (20 min, 4 or 408C) in the various

sample buffers used for protein separation by gel electrophore-

sis. Typically, we observed a band of 50 kD instead of the ex-

pected 65 kD, which corresponds to the whole fusion protein

(Figure 9D). Only solubilization of the MP and M fractions at 958C

for 3 min prevented the proteolysis (no trace of the 50-kD band

was detected). By contrast, the OCP-GFP detected in the soluble

protein fraction was not proteolyzed under any conditions. As

noted above, a 16-kD RCP corresponding to the N terminus of the

OCP was isolated from cells concomitantly with OCP. The 50-kD

band may correspond to the GFP fused to the C-terminal domain

of proteolyzed OCP. The biological significance of this derivative

form, if any, is unknown, but in several cyanobacterial genomes

(Gloeobacter violaceus, Nostoc punctiforme, and Nostoc PCC

7120), in addition to the full-length OCP gene there are other genes

corresponding to RCP-like proteins (the N-terminal domain of the

OCP) (Kerfeld, 2004b). Moreover, in Thermosynechoccocus elon-

gatus, there are single-copy open reading frames for the N- and

C-terminal domains of the OCP. These data suggest that the N-

terminal RCP may have a distinct function and/or that different

combinations of N- and C-terminal domains result in different

OCPs that may be tailored for different (photoprotective) roles.

DISCUSSION

OCP Provides Photoprotection through NPQ

The results presented in this work demonstrate that the OCP is

specifically involved in NPQ that appears to be associated with a

photoprotective energy-dissipation mechanism. In the absence

of the OCP, the NPQ induced by strong white or blue-green light

in Synechocystis PCC 6803 cells is completely inhibited, and the

cells are more sensitive to high light intensities. The observation

that the effective antenna size was smaller in the cells in the

quenched state and that the oxygen-evolving activity decreased

faster in the DOCP mutant under high-light conditions strongly

supports the hypothesis that the OCP-related mechanism dis-

sipates the excess absorbed energy, thereby decreasing the

amount of energy arriving at the photochemical centers.

OCP-Related NPQ Is Associated with Phycobilisomes

Intense blue-green light was able to induce NPQ in Synecho-

cystis PCC 6803 mutants lacking the inner chlorophyll antenna,

CP43 and CP47, but it was unable to induce NPQ in mutants

lacking phycobilisomes or the phycobilisome core. In addition,

fluorescence spectra showed that when the cells were in the

quenched state, the fluorescence emitted by the phycobilisomes

decreased and that there was less energy transfer from the

phycobilisomes to PSII and PSI. These results demonstrated that

NPQ induced by blue-green light and inhibited by the absence

of the OCP is related to the phycobilisomes. The evidence for

Figure 7. (continued).

(C) Wild-type cells adapted to low blue-green light intensities (high fluorescence state) were illuminated with strong blue-green light (400 to 550 nm) at

300 (circles), 470 (squares), and 730 (diamonds) mmol�m�2�s�1 or with green light (500 to 550 nm, 510 mmol�m�2�s�1; triangles) to induce the quenched

state.

(D) Fluorescence level changes in dark-adapted DCP47 mutant cells during illumination at 150 (dashed line), 350 (dotted line), and 1000 (solid line)

mmol�m�2�s�1 blue-green light and in dark-adapted DCP47-DOCP mutant cells during illumination at 1000 mmol�m�2�s�1 blue-green light (solid line).

The DCP47 and DCP47-DOCP mutants lack variable fluorescence because of the absence of PSII reaction centers.

(E) Dark-adapted wild-type cells were illuminated with orange-red (600 to 650 nm; squares), blue-green (400 to 550 nm; circles), or green (triangles) light

at 470 mmol�m�2�s�1. Saturating pulses separated by 30 s were applied to assess Fm9. The orange-red light at this intensity closed 95% of PSII centers,

whereas blue-green and green light closed only 20 and 10% of centers, respectively.

(F) Fluorescence level changes in dark-adapted DCP47 cells illuminated by orange-red (600 to 700 nm, 3000 mmol�m�2�s�1; dashed line), green (500 to

550 nm, 470 mmol�m�2�s�1; dotted line), and blue-green (400 to 550 nm, 470 mmol�m�2�s�1; solid line) light.

(G) Measurements of fluorescence yield by a PAM fluorometer for wild-type cells adapted to low blue-green light intensities (high fluorescence state) in

the presence of DCMU (solid line), nigericin (squares), or without additions (circles) illuminated for 200 s with strong blue-green light (740 mmol�m�2�s�1 )

to induce the quenched state and then illuminated with low blue-green light (80 mmol�m�2�s�1) to allow fluorescence recovery.

(H) Measurements of fluorescence yield by a PAM fluorometer for dark-adapted DIsiA cells illuminated successively with low-intensity blue-green light

(400 to 550 nm, 80 mmol�m�2�s�1) and high-intensity blue-green light (740 mmol�m�2�s�1).

Carotenoids and NPQ in Cyanobacteria 1001

Page 11: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

an interaction between the OCP and the phycobilisomes and

thylakoids necessary for this NPQ was supported by the coiso-

lation of the OCP-GFP fusion protein with the phycobilisome-

associated membrane fraction. Our results are in accord with

those of a proteomic study of Synechocystis PCC 6803 thyla-

koids (Srivastava et al., 2005) in which the OCP was observed in

a purified thylakoid preparation. Furthermore, our studies by im-

munogold labeling and electron microscopy showed that most

of the OCP-GFP fusion protein is present in the interthylakoid

region of the cell.

This NPQ also occurs in a Synechocystis PCC 6803 mutant

lacking phycocyanin (CK) in which only phycobilisome cores are

present. This finding suggests that the OCP interacts with a

component of the phycobilisome core. The Lcm, the linker-

membrane protein that acts as the terminal energy acceptor in

the phycobilisome, is a good candidate for this role. The faster

recovery from the quenched state in CK compared with the wild

type implies a weaker interaction between the OCP and the

phycobilisome and/or the involvement of the movement of

the phycobilisome relative to the membrane, as suggested by

Joshua et al. (2005).

Under Fe-starvation conditions, a large reversible quenching

of Fo and Fm levels, suggested to be mediated by the IsiA

protein, is also induced by blue light (Cadoret et al., 2004; Bailey

et al., 2005; Joshua et al., 2005). Because we demonstrated that

the blue-green light–induced NPQ is not mediated by the IsiA

protein, we propose that the blue light–induced NPQ observed

under Fe-starvation conditions is also related to phycobilisome

and OCP; the presence of IsiA only increases the extent of the

quenching. If part of the phycobilisomes transfer absorbed en-

ergy to IsiA complexes, that contributes to the Fo level (Joshua

et al., 2005); in cells containing quenched phycobilisomes, less

energy will arrive at the IsiA complexes and a larger fluorescence

quenching will be observed. This blue light mechanism seems

to be independent of the quenched state of the large IsiA

oligomers that appear in long-term iron-depleted cells, as

Figure 8. In Situ Localization of the OCP-GFP Fusion Protein.

(A) Distribution of green fluorescence. Phase contrast (left), GFP fluorescence (middle), and chlorophyll and phycobiliprotein fluorescence (right)

micrographs of OCP-GFP (row 1) and wild-type (row 2) cells are shown. Bar ¼ 3 mm.

(B) Histogram of gold particle density in whole cells (bar 1), interthylakoid region (bar 2), and non-thylakoid-related cytoplasm (bar 3). Thirty-three cells

were analyzed and 1120 gold particles were counted. On average, the thylakoid region represents 65% of the cellular surface. Error bars indicate SE.

(C) Immunogold labeling of a thin section of OCP-GFP–transformed cells. Panel 1, OCP-GFP cells were labeled with a polyclonal antibody against the

GFP coupled to 10-nm gold particles; panel 2, no labeling was observed without the primary antibody. Bar ¼ 0.5 mm.

1002 The Plant Cell

Page 12: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

already suggested by Ihalainen et al. (2005). Energy dissipation in

IsiA aggregates was independent of the quality of the excitation

light.

NPQ Appears to Be Induced via Activation of the OCP by

Light Absorption

In higher plants, the antenna-associated NPQ is induced by low

pH in the thylakoid lumen. By contrast, the phycobilisome-

associated NPQ described here is not dependent on the pres-

ence of a transthylakoid DpH, as shown by the inability of

uncouplers to affect the kinetics of quenching and recovery.

Several lines of evidence indicate that the induction of the

quenching is also independent of excitation pressure on PSII or

changes in the redox state of the plastoquinone pool. On the one

hand, the intensities of blue light that induce the NPQ are not

saturating for oxygen evolution, and NPQ can occur when a small

number of PSII centers are closed (El Bissati et al., 2000). On the

other hand, NPQ could be induced under conditions in which the

plastoquinone pool becomes (1) more reduced (by transfer from

darkness or dim blue light to high intensities of white light) or (2)

more oxidized (when going from dark to blue light, in the pres-

ence of DCMU) or (3) when its redox state is unaffected (in the

PSII-lacking mutant). In addition, because blue light is absorbed

by PSI, we cannot discount the possible involvement of PSI

activity, but this is unlikely because green light is very efficient in

the induction of NPQ.

Recently, two independent studies showed that in two different

Synechocystis PCC 6803 mutants lacking PSII, blue-green light

induces a quenching of phycobilisome emission (Rakhimberdieva

et al., 2004; Scott et al., 2005). The action spectrum for the

Figure 9. GFP Fluorescence Measurement, Gel Electrophoresis, and

Protein Gel Blot Analysis of Soluble and Membrane Fractions.

(A) Fluorescence emission spectra. GFP emission in the membrane frac-

tions MP (for membrane-bound phycobilisomes) and M (for membrane-free

phycobilisomes) at 4 mg chlorophyll/mL and the soluble fractions sup MES

(for MES supernatant) and sup P/C (for phosphate-citrate supernatant) at

a concentration corresponding to that of the membrane fractions (see

Methods). Excitation was at 480 nm. Values shown are means of four

independent experiments.

(B) Fluorescence emission spectra. GFP in the MP fraction (3.5 mg

chlorophyll/mL) and in the M and phycobilisome (PBS) fractions obtained

after suspension of the MP fraction in MES buffer (1-h incubation) and

subsequent centrifugation. Values shown are means of three different

experiments.

(C) Coomassie blue–stained gel electrophoresis and immunoblot detec-

tion of the OCP-GFP fusion protein of the M fraction (lane 1), the MP

fraction (lane 2), and the soluble fractions of cells broken in P/C buffer

(lane 4) and MES (lane 5); lane 3, molecular mass markers. In the

immunoblot (lane 3), the heavy chain of the mouse IgG is visualized.

(D) Coomassie blue–stained gel electrophoresis and immunoblot detec-

tion of the OCP-GFP fusion protein of the MP (lanes 2 and 4) and M (lanes

1 and 5) fractions solubilized for 3 min at 958C (lanes 1 and 2) or for 20 min

at 48C (lanes 4 and 5); lane 3, molecular mass markers. In the immunoblot

(lane 3), the heavy chain of the mouse IgG is visualized.

Each slot contained 2 mg of chlorophyll of the MP or M fraction or the

corresponding volumes of the supernatants (see Methods). The rela-

tionship between the volumes of membrane and soluble fractions was

the same for the fluorescence spectra and gels.

Carotenoids and NPQ in Cyanobacteria 1003

Page 13: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

phycobilisome quenching, in both cases, was reminiscent of a

carotenoid absorption spectrum. Moreover, Rakhimberdieva

et al. (2004) proposed that a carotenoid activated by blue light

induces the reversible NPQ. Here, we have shown that the ab-

sence of the OCP completely inhibits the blue light–induced

phycobilisome-related NPQ in PSII-lacking mutants and in the

wild type, demonstrating that this carotenoid is associated with

the OCP. The action spectra (Rakhimberdieva et al., 2004)

resembled the absorption spectra of the OCP in Arthrospira

maxima (Polivka et al., 2005). We cannot completely rule out the

involvement of a cryptochrome or a BLUF protein as the blue light

sensor, but the fact that green light (500 to 550 nm) induced NPQ

very efficiently renders this improbable. The oxidized flavin ade-

nine dinucleotide, the cofactor of these blue photoreceptors, has

a similar spectrum in the blue region, but it absorbs very weakly in

the green region (for the BLUF [Slr1694] spectrum of Synecho-

cystis PCC 6803, see Masuda et al., 2004).

Working Models for OCP Activity

Many questions arise from our findings about the role of the OCP

in the phycobilisome-related NPQ. First, what are the changes

that are induced in the carotenoid and in the protein by high light

intensities that activate the OCP to induce the quenching? For

example, blue-green light absorption could lead to carotenoid

isomerization, inducing conformational changes in the chromo-

phore and in the protein. Alternatively (or in addition), proton or

electron transfer could be induced as in the flavin-containing

blue receptors. Changes in the tertiary or quaternary structure

(OCP is a dimer) or proteolysis of the protein may also be involved

in the mechanism.

Our results do not permit conclusions to be drawn on the

mechanism by which energy is dissipated. There are several

possibilities. The OCP could be the light sensor, a mediator of

quenching, the quencher itself, or play a combination of these

roles. The activated OCP, through interaction with the core of the

phycobilisome, could cause an alteration of the phycobilisome

structure, leading to the quenched state. Alternatively, the ca-

rotenoid of the OCP could interact directly with a phycobilin

chromophore (most probably the terminal acceptor) and dissi-

pate the absorbed energy.

This study demonstrated that the OCP is specifically involved

in a phycobilisome-associated NPQ and not in other mecha-

nisms affecting the levels of fluorescence (e.g., state transitions

or D1 damage). This mechanism, induced by both blue-green

light and saturating intensities of white light, likely protects PSII

from high light exposure. We report here a novel function for a

soluble carotenoid protein as a mediator of a photoprotective

response. Our results reveal the role of the OCP and confirm the

hypothesis that in cyanobacteria, as in higher plants, there exists

a mechanism involving energy dissipation in the antenna (phy-

cobilisomes). The molecular mechanism of this novel process

awaits elucidation. We propose a working model in which the

OCP acts as a photoreceptor that responds to blue-green light

and subsequently induces energy dissipation (and fluorescence

quenching) through interaction with the phycobilisome core.

Further testing of this hypothesis is likely to open new frontiers in

carotenoid function.

METHODS

Plasmid Construction

DOCP Plasmid

The ocp gene (slr1963) was amplified by PCR using genomic DNA of

Synechocystis PCC 6803 as template. Two synthesized oligonucleotides

were used as primers: car1, 59-CTACGCGGATCCATTGACTCTGCCCG-

CGGAATTT-39, and car2, 59-CGGCCGCTCGAGCAAAGTTGAGTAATT-

CTTTGGG-39. The resulting PCR product was digested by EagI and AvaI

restriction enzymes and then cloned in the polylinker EagI-XhoI restriction

sites of pBluescript SKþ (Stratagene) plasmid. The slr1963 gene was

interrupted by inserting a 2.2-kb DNA fragment (v cassette) containing

the aadA gene from Tn7, conferring resistance to Sp/Sm, in the unique

restriction site HincII.

OCP-GFP Plasmid

The slr1963gene ofSynechocystisPCC 6803 was amplified by PCR using

the oligonucleotides car1 and car4 (59-CTGAAGGGAGTTAGGATCCC-

GAGCAAAGTTGAG-39) containing a BamHI restriction site. The stop

codon was suppressed. The PCR product was cleaved with SphI and

BamHI restriction enzymes and cloned into the polylinker SphI and

BamHI restriction sites of a pEGFP ampicillin-resistant vector to create a

recombinant slr1963-GFP gene. In addition, the immediately downstream

slr1964 open reading frame, encoding a hypothetical protein, was am-

plified by PCR using two other oligonucleotides: the NotI-creating primer

(car5,59-TTACTAACTTTGGCGGCCGCAATAACTCCCTTCAGAG-39)and

the SpeI-creating primer (car6, 59-CACCGGACTAGTCAAAAACTAT-

CTGCTGGCGATCG-39). This second PCR fragment was cloned into

the pEGFP/OCP ampicillin-resistant vector opened with the same en-

zymes. The slr1964 gene was inactivated by insertion of the Sp/Sm

cassette in the unique ClaI restriction site.

Dslr1964 Plasmid

The OCP-GFP plasmid was cleaved with both NotI and SpeI restriction

enzymes. The digestion product of 3.1 kb contained the slr1964 gene

inactivated by insertion of the Sp/Sm cassette in the unique ClaI restric-

tion site. This DNA fragment was cloned into a pBluescript SKþ ampi-

cillin-resistant vector opened with the same enzymes to obtain the

Dslr1964 plasmid.

DisiA Plasmid

The isiA gene (sll0247) of Synechocystis PCC 6803 was amplified by PCR

and cloned in the polylinker SrfI restriction site of a pPCR-Script SKþampicillin-resistant vector. The isiA gene was inactivated by inserting

Sp/Sm cassette in the unique PmlI restriction site.

DCP47 Plasmid

The DNA region of Synechocystis PCC 6803 containing the psbB gene,

encoding the CP47 protein, was amplified by PCR using the oligonucle-

otides BM (59-CATGGTGATAATCAAGGGATG-39) and BK (59-CGCTT-

TCGTCGTGGCCGGTAC-39). The PCR fragment containing the psbB

gene was cloned onto the EcoRV site of the pBC SKþ plasmid. In the

resulting plasmid, a 550-bp BstEII fragment was substituted by the eryth-

romycin resistance cassette.

Transformation, Selection, and Genetic Analysis of Mutants

The DOCP, OCP-GFP, DisiA, DCP47, and Dslr1964 plasmid constructs

were used to transform wild-type Synechocystis sp PCC 6803. To obtain

1004 The Plant Cell

Page 14: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

the double mutants DCP47-DOCP and CK-DOCP, DCP47 and CK cells

were transformed with the DOCP plasmid. Transformants were selected

under dim light at 328C on plates containing different antibiotics: 25mg/mL

spectinomycin and 12 mg/mL streptomycin, 20 mg/mL erythromycin, or

40 mg/mL kanamycin. The nonphotosynthetic DpsbB mutant was grown

in the presence of 20 mM glucose because it has an obligate heterotroph

phenotype.

Genomic DNA was isolated from Synechocystis PCC 6803 essentially

as described by Cai and Wolk (1990). To confirm the homoplasmicity and

complete segregation of the different mutants, PCR analysis and specific

digestions by restriction enzymes were performed.

Culture Conditions

Wild-type and mutant cells were grown photoautotrophically in a mod-

ified BG11 medium as described by Herdman et al. (1973) containing

twice the concentration of sodium nitrate. Cells were shaken in a rotary

shaker (120 rpm) at 308C and illuminated by fluorescent white lamps

giving a total intensity of ;30 to 40 mmol�m�2�s�1 under a CO2-enriched

atmosphere. The selected mutants were grown in the presence of the

appropriate antibiotics. The cells were maintained in the logarithmic phase

of growth and were collected having optical densities of 0.6 to 0.8 at 800 nm.

Photoinhibition

For the photoinhibition experiment (Figure 3A), cells (10 mg chlorophyll/

mL) were incubated at 308C in a glass tube (3 cm diameter) under stirring

and illuminated with three Atralux spots of 150 W (1000 mmol�m�2�s�1

each lamp). The effective intensity of light absorbed by each cell was

much less as a result of self-absorption by the cells in the glass tube. For

comparison, when wild-type cells (10mg chlorophyll/mL) were illuminated

in the PAM cuvette (1500 mmol�m�2�s�1), oxygen-evolving activity was

abolished within 30 min. Therefore, we estimate the actual amount of

incident light to be <1500 mmol�m�2�s�1. For the experiment shown in

Figure 3B, wild-type cells were illuminated for 10 min (same conditions as

in Figure 3A) and then incubated on ice to inhibit the recovery of NPQ

(data not shown) until oxygen evolution measurements could be com-

pleted. Oxygen evolution of intact cells (10 mg chlorophyll/mL) was

measured polarographically using a Clark-type oxygen electrode with the

addition of 1 mM 2,6-dichlorobenzoquinone as an artificial PSII electron

acceptor.

Fluorescence Measurements and Detection

In the course of our studies, we always observed the blue-green light–

induced NPQ in wild-type cells; however, the extent of the reversible NPQ

and the kinetics of recovery were sensitive to cell concentration, phyco-

bilisome concentration, PSII:PSI and Fv:Fo ratios, etc. Thus, wild-type

and mutant cells were grown in similar conditions (see Culture Conditions)

and collected at an equal cell concentration and phycocyanin:chlorophyll

ratio. The yield of chlorophyll fluorescence was monitored in a modulated

fluorometer (PAM; Walz, Effelrich, Germany) adapted to a Hansatech

oxygen electrode as described previously (El Bissati et al., 2000). All NPQ

induction and recovery experiments were performed in a stirred cuvette

of 1 cm diameter (328C) at a chlorophyll concentration of 3 mg/mL in the

presence of chloramphenicol (30 mg/mL) to inhibit protein synthesis.

Recovery was in darkness or under 50 to 80 mmol�m�2�s�1 blue-green

light. The nomenclature used was as follows: Fo, minimal fluorescence

level, the fluorescence emitted by open reaction centers in dark-adapted

cells; Fm, maximal fluorescence level in dark-adapted samples; Fm9,

maximum fluorescence under illumination, corresponding to the fluores-

cence emitted by the maximum concentration of closed reaction centers;

Fs, steady state fluorescence level; Fv, variable fluorescence ¼ Fm � Fo.

The Fo level was determined by illuminating dark-adapted cells with a low

intensity of red-modulated light (pulses of 1 ms, 1.6 kHz, 0.024 mmol�m�2�s�1). Saturating pulses (2000 mmol�m�2�s�1, 1 s) were applied to

measure Fm and Fm9 levels. Application of such pulses that transiently

close all PSII centers serves to distinguish between photochemical

quenching and NPQ.

The fluorescence emission spectra were recorded on a Hitachi F-3010

fluorescence spectrophotometer. Excitation was done at 600 or 430 nm.

The cells were at a concentration of 5mg chlorophyll/mL. Phycoerythrin or

fluorescein (1.75mM) was added to facilitate normalization of the spectra.

Phycoerythrin (isolated from Rhodella violacea) was added to obtain a

phycoerythrin emission slightly higher than those of Synechocystis phy-

cocyanin and allophycocyanin.

To detect green fluorescence from the GFP-OCP, cyanobacteria were

deposited on a glass slide and mounted for observation. Micrographs

were taken with an Olympus VAN-Ox AH2 fluorescence microscope fitted

with a color cooled charge-coupled device camera (Coolsnap; Roper

Scientific).

MP and M Membrane Preparations

Cells were resuspended in a buffer of 0.5 M K-phosphate and 0.3 M Na-

citrate (pH 6.8) (P/C buffer) to obtain MP and in a 20 mM MES, pH 6.8,

buffer to obtain M at a chlorophyll concentration of 1 mg/mL and broken

in a mini-bead-beater in the presence of glass beads. The M and MP

fractions were collected by centrifugation and frozen at –808C until use for

gel electrophoresis. The homogenates containing the broken cells were

loaded on a continuous sucrose gradient (0 to 50%) prepared in the P/C

buffer and centrifuged. The colored band containing the MP fraction or

the M fraction was collected, pelleted by centrifugation, and stored at

�808C. We are aware that the M and MP fractions obtained only by

centrifugation could be contaminated by cytoplasmic membranes. How-

ever, similar results on detection of the OCP-GFP fusion protein were

obtained using the M and MP fractions collected directly by centrifugation

and those collected by the sucrose gradient that are more purified (data

not shown).

The OCP-GFP in the M, MP, and soluble fractions compared in protein

gel blots and fluorescence spectra (Figure 9) was isolated and quantified

as follows: 0.5 mL of broken cells (1 mg chlorophyll/mL) (in MES or P/C

buffer) was centrifuged. The pellet containing the membrane fraction was

resuspended in 100mL of MES buffer. The supernatant was concentrated

to 100 mL. We assumed that 1 mL of the concentrated supernatant

corresponded to ;1 mL of the resuspended membrane fraction. This

assumption was confirmed by the observation that the chlorophyll:

phycocyanin absorption ratio in a solution containing 5 mL of supernatant

and 5 mL of membranes was equivalent to that of whole cells. If

necessary, adjustments were made by dilution to equalize the ratios

between the two samples. For the fluorescence spectra shown in Figure

9B, 20 mL of MP suspended in P/C buffer was diluted with 2 mL of MES

buffer and incubated on ice for 1 h. The membranes were subsequently

precipitated by centrifugation and resuspended in 100 mL of MES. The

supernatant was concentrated to 100 mL. Again, the chlorophyll:phyco-

cyanin ratio was used to normalize the samples.

Gel Electrophoresis and Protein Gel Blots

Proteins of the MP and M fractions and of the soluble fractions of the

OCP-GFP mutant were analyzed by SDS-PAGE on a 12% polyacryla-

mide/2 M urea gel in a Tris/MES system (Kashino et al., 2001). The OCP-

GFP fusion protein was detected by a monoclonal antibody against GFP

(Clontech).

Electron Microscopy

Samples were fixed in 4% paraformaldehyde and embedded in Unicryl.

For immunogold labeling, the sections were incubated with the primary

Carotenoids and NPQ in Cyanobacteria 1005

Page 15: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

antibody (rabbit polyclonal antibody against GFP [Abcam]; 3 mg/mL final

dilution) in buffer T (20 mM Tris-HCl, 154 mM NaCl, 0.1% NaN3, 0.1%

BSA, 0.05% Tween 20, and 0.1% fish gelatin); then, after washing, they

were incubated with a 1:20 dilution of 10-nm gold-conjugated anti-rabbit

antibodies (Amersham) and stained in 2% uranyl acetate followed by lead

citrate. The sections were observed on a Philips EM 400 electron

microscope (FEI).

Accession Numbers

Sequence data from this article can be found in the GenBank/EMBL data

libraries under the following accession numbers: slr1963, NP_441508;

slr1964, NP_441509; isiA (sll0247), NP_441268; psbB (slr0906),

NP_442388; psbC (sll0851), NP_441119; psbD (sll0849), NP_441120;

psbD2 (slr0927), NP_442780; apcA (slr2067), NP_441194; apcB (slr1986),

NP_441195; apcA (slr2067), NP_441194; apcB (slr1986), NP_441195;

apcE (slr0335), NP_441972; cpcA (sll1578), NP_440551; cpcB (sll1577),

NP_440552; cpcC2 (sll1579), NP_440550; cpcC1 (sll1580), NP_440549;

cpcD (ssl3093), NP_440548.

ACKNOWLEDGMENTS

We thank W. Vermaas for the gift of the DpsbD-DpsbC mutant, A.

William Rutherford for stimulating discussions and critical reading of the

manuscript, Bernard Lagoutte for helpful advice and discussions, Estelle

Delphin for the low-temperature fluorescence spectra, and Krisztian Cser

for the light saturation curves of oxygen evolution. This research was

partially supported by European Union network INTRO2.

Received December 7, 2005; revised February 6, 2006; accepted Feb-

ruary 17, 2006; published March 10, 2006.

REFERENCES

Adir, N. (2005). Elucidation of the molecular structures of components

of the phycobilisome: Reconstructing a giant. Photosynth. Res. 85,

15–32.

Ajlani, G., and Vernotte, C. (1998). Construction and characterization of

a phycobiliprotein-less mutant of Synechocystis sp. PCC 6803. Plant

Mol. Biol. 37, 577–580.

Ajlani, G., Vernotte, C., DiMagno, L., and Haselkorn, R. (1995).

Phycobilisome core mutants of Synechocystis PCC 6803. Biochim.

Biophys. Acta 1231, 189–196.

Allen, J.F. (1992). Protein phosphorylation in regulation of photosyn-

thesis. Biochim. Biophys. Acta 1098, 275–335.

Aro, E.M., Virgin, I., and Andersson, B. (1993). Photoinhibition of

photosystem II. Inactivation, protein damage and turnover. Biochim.

Biophys. Acta 1143, 113–134.

Bailey, S., Mann, N., Robinson, C., and Scanlan, D.J. (2005).

The occurrence of rapidly reversible non-photochemical quenching

of chlorophyll a fluorescence in cyanobacteria. FEBS Lett. 579,

275–280.

Bibby, T.S., Nield, J., and Barber, J. (2001). Iron deficiency induces the

formation of an antenna ring around trimeric photosystem I in cyano-

bacteria. Nature 412, 743–745.

Boekema, E.J., Hifney, A., Yakushevska, A.E., Piotrowski, M.,

Keegstra, W., Berry, S., Michel, K.P., Pistorius, E.K., and Kruip,

J. (2001). A giant chlorophyll-protein complex induced by iron defi-

ciency in cyanobacteria. Nature 412, 745–748.

Burnap, R., Troyan, T., and Sherman, L. (1993). The highly abundant

chlorophyll-protein of iron-deficient Synechococcus sp PCC 7942

(CP439) is encoded by the isiA gene. Plant Physiol. 103, 893–902.

Cadoret, J.-C., Demouliere, R., Lavaud, J., van Gorkom, H., Houmard,

J., and Etienne, A.-L. (2004). Dissipation of excess energy triggered by

blue light in cyanobacteria with CP439 (isiA). Biochim. Biophys. Acta

1659, 100–104.

Cai, Y., and Wolk, C.P. (1990). Use of a conditionally lethal gene in

Anabaena sp strain PCC7120 to select for double recombination and

entrap insertion sequences. J. Bacteriol. 172, 3138–3145.

Campbell, D., Hurry, V., Clarke, A., Gustafsson, P., and Oquist, G.

(1998). Chlorophyll fluorescence analysis of cyanobacterial photosyn-

thesis and acclimation. Microbiol. Mol. Biol. Rev. 62, 667–683.

Demmig-Adams, B. (1990). Carotenoids and photoprotection in plants:

A role for the xanthophyll zeaxanthin. Biochim. Biophys. Acta 1020,

1–24.

Eaton-Rye, J., and Vermaas, W. (1991). Oligonucleotide-directed

mutagenesis of psbB, the gene encoding CP47, employing a deletion

mutant strain in the cyanobacterium Synechocystis sp. PCC 6803.

Plant Mol. Biol. 17, 1165–1177.

El Bissati, K., Delphin, E., Murata, N., Etienne, A.-L., and Kirilovsky,

D. (2000). Photosystem II fluorescence quenching in the cyanobac-

terium Synechocystis PCC 6803: Involvement of two different mech-

anisms. Biochim. Biophys. Acta 1457, 229–242.

El Bissati, K., and Kirilovsky, D. (2001). Regulation of psbA and psaE

expression by light quality in Synechocystis species PCC 6803. A

redox control mechanism. Plant Physiol. 125, 1988–2000.

Gantt, E., and Conti, S.F. (1966). Granules associated with the chlo-

roplast lamellae of Porphyridium cruentum. J. Cell Biol. 29, 423–434.

Gilmore, A., and Yamamoto, H. (1993). Linear models relating xan-

thophylls and lumen acidity to non-photochemical fluorescence

quenching. Evidence that antheraxanthin explains zeaxanthin-inde-

pendent quenching. Photosynth. Res. 35, 67–68.

Glazer, A. (1984). Phycobilisome. A macromolecular complex optimised

for light energy transfer. Biochim. Biophys. Acta 768, 29–51.

Havaux, M., Guedeney, G., Hagemann, M., Yeremenko, N., Matthijs,

H., and Jeanjean, R. (2005). The chlorophyll-binding protein IsiA is

inducible by high light and protects the cyanobacterium Synechocys-

tis PCC6803 from photooxidative stress. FEBS Lett. 579, 2289–2293.

Herdman, M., Delaney, S.F., and Carr, N.G. (1973). A new medium for

the isolation and growth of auxotrophic mutants of the blue-green

alga Anacystis nidulans. J. Gen. Microbiol. 79, 233–237.

Hihara, Y., Kamei, A., Kanehisa, M., Kaplan, A., and Ikeuchi, M.

(2001). DNA microarray analysis of cyanobacterial gene expression

during acclimation to high light. Plant Cell 13, 793–806.

Holt, T.K., and Krogmann, D.W. (1981). A carotenoid protein from

cyanobacteria. Biochim. Biophys. Acta 637, 408–414.

Horton, P., Ruban, A.V., and Walters, R.G. (1996). Regulation of light

harvesting in green plants. Annu. Rev. Plant Physiol. Plant Mol. Biol.

47, 655–684.

Ihalainen, J., D’Haene, S., Yeremenko, N., van Roon, H., Arteni, A.,

Boekema, E., van Grondelle, R., Matthijs, H., and Dekker, J.

(2005). Aggregates of the chlorophyll-binding protein IsiA (CP439)

dissipate energy in cyanobacteria. Biochemistry 44, 10846–10853.

Jeanjean, R., Zuther, E., Yeremenko, N., Havaux, M., Matthijs, H.,

and Hagemann, M. (2003). A photosystem I psaFJ-null mutant of the

cyanobacterium Synechocystis 6803 expresses the isiAB operon

under iron replete conditions. FEBS Lett. 549, 52–56.

Joshua, S., Bailey, S., Mann, N., and Mullineaux, C. (2005). Involve-

ment of phycobilisome diffusion in energy quenching in cyanobacte-

ria. Plant Physiol. 138, 1577–1585.

Kashino, Y., Koike, K., and Satoh, K. (2001). An improved SDS-PAGE

system for the analysis of membrane protein complexes. Electropho-

resis 22, 1004–1007.

1006 The Plant Cell

Page 16: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

Kerfeld, C.A. (2004a). Structure and function of the water-soluble

carotenoid-binding proteins in cyanobacteria. Photosynth. Res. 81,

215–225.

Kerfeld, C.A. (2004b). Water-soluble carotenoid proteins of cyanobac-

teria. Arch. Biochem. Biophys. 430, 2–9.

Kerfeld, C.A., Sawaya, M., Brahmandam, V., Cascio, D., Ho, K.,

Trevithick-Sutton, C., Krogmann, D.W., and Yeates, T.O. (2003).

The crystal structure of a cyanobacterial water-soluble protein.

Structure 11, 1–20.

Laudenbach, D., and Straus, N. (1988). Characterization of a cyano-

bacterial iron stress-induced gene similar to psbC. J. Bacteriol. 170,

5018–5026.

Li, X.-P., Bjorkman, O., Shih, C., Grossman, A.R., Rosenquist, M.,

Jansson, S., and Niyogi, K.K. (2000). A pigment-binding protein

essential for regulation of photosynthetic light harvesting. Nature 403,

391–395.

MacColl, R. (1998). Cyanobacterial phycobilisomes. J. Struct. Biol. 124,

311–334.

Masuda, S., Hasegawa, K., Ishii, A., and Ono, T. (2004). Light-induced

structural changes in a putative blue-light receptor with a novel FAD

binding fold sensor of blue-light using FAD (BLUF); Slr1694 of

Synechocystis sp PCC 6803. Biochemistry 43, 5304–5313.

Melis, A. (1999). Photosystem II damage and repair cycle in chloro-

plasts: What modulates the rate of photodamage in vivo. Trends Plant

Sci. 4, 130–135.

Muller, P., Li, X.-P., and Niyogi, K. (2001). Non-photochemical quench-

ing. A response to excess light energy. Plant Physiol. 125, 1558–1566.

Mullineaux, C.K. (1992). Excitation energy transfer from phycobili-

somes to photosystem I in a cyanobacterium. Biochim. Biophys. Acta

1100, 285–292.

Niyogi, K. (1999). Photoprotection revisited: Genetic and molecular

approaches. Annu. Rev. Plant Physiol. Plant Mol. Biol. 50, 333–359.

Park, Y., Sandstrom, P., Gustafsson, P., and Oquist, G. (1999). Ex-

pression of the isiA gene is essential for the survival of the cyano-

bacterium Synechococcus sp PCC 7942 by protecting photosystem II

from excess light under iron limitation. Mol. Microbiol. 32, 123–129.

Polivka, T., Kerfeld, C.A., Pascher, T., and Sundstrom, V. (2005).

Spectroscopic properties of the carotenoid 39-hydroxyechinenone in

the orange carotenoid protein from the cyanobacterium Arthrospira

maxima. Biochemistry 44, 3994–4003.

Prasil, O., Adir, N., and Ohad, I. (1992). Dynamics of photosystem II:

Mechanism of photoinhibition and recovery processes. In The Pho-

tosystems: Structure, Function and Molecular Biology, J. Barber, ed

(Amsterdam: Elsevier Science Publishers), pp. 295–348.

Rakhimberdieva, M., Boichenko, V.A., Karapetyan, N., and Stadnichuk,

I. (2001). Interaction of phycobilisomes with photosystem II dimers and

photosystem I monomers and trimers in the cyanobacterium Spirulina

platensis. Biochemistry 40, 15780–15788.

Rakhimberdieva, M., Stadnichuk, I., Elanskaya, I., and Karapetyan,

N. (2004). Carotenoid-induced quenching of the phycobilisome fluo-

rescence in photosystem II-deficient mutant of Synechocystis sp.

FEBS Lett. 574, 85–88.

Redlinger, T., and Gantt, E. (1982). A Mr 95000 polypeptide in

Porphyridium cruentum phycobilisomes and thylakoids: Possible

function in linkage of phycobilisomes to thylakoids and in energy

transfer. Proc. Natl. Acad. Sci. USA 79, 5542–5546.

Ruban, A.V., Ress, D., Pascal, A.A., and Horton, P. (1992). Mecha-

nism of DpH-dependent dissipation of absorbed excitation energy by

photosynthetic membranes. II. The relationship between LHCII ag-

gregation and QE in isolated thylakoids. Biochim. Biophys. Acta 1102,

39–44.

Sandstrom, S., Park, Y., Oquist, G., and Gustafsson, P. (2001).

CP439, the isiA gene product, functions as an excitation energy

dissipator in the cyanobacterium Synechococcus sp PCC 7942.

Photochem. Photobiol. 74, 431–437.

Scott, M., Vasil’ev, S., McCollum, C., Crozier, C., Espie, G., and

Bruce, D. (2005). Blue light induced fluorescence quenching in a PSII-

less mutant of Synechocystis PCC 6803. In Photosynthesis: Funda-

mental Aspects to Global Perspectives, A. van der Est and D. Bruce,

eds (Amsterdam: Elsevier Science Publishers), pp. 577–579.

Srivastava, R., Pisareva, T., and Norling, B. (2005). Proteomic studies

of the thylakoid membrane of Synechocystis sp PCC 6803. Proteo-

mics 5, 4905–4916.

Tandeau de Marsac, N. (2003). Phycobiliproteins and phycobilisomes:

The early observations. Photosynth. Res. 76, 197–205.

van Thor, J.J., Mullineaux, C.W., Matthijs, H.C., and Hellingwerf, K.J.

(1998). Light harvesting and state transitions in cyanobacteria. Bot.

Acta 111, 430–443.

Wollman, F.-A. (2001). State transitions reveal the dynamics and

flexibility of the photosynthetic apparatus. EMBO J. 20, 3623–3630.

Wu, Y.P., and Krogmann, D.W. (1997). The orange carotenoid protein

of Synechocystis PCC 6803. Biochim. Biophys. Acta 1322, 1–7.

Yamamoto, H. (1979). Biochemistry of the violaxanthin cycle in higher

plants. Pure Appl. Chem. 51, 639–648.

Yeremenko, N., Kouril, R., Ihalainem, J., D’Haene, S., van Oosterwijk,

N., Andrizhiyevkaya, E., Keegstra, W., Dekker, H., Hagemann, M.,

Boekema, E., Matthijs, H., and Dekker, J. (2004). Supramolecular

organization and dual function of the IsiA chlorophyll-binding protein in

cyanobacteria. Biochemistry 43, 10308–10313.

Yousef, N., Pistorius, E.K., and Michel, K.-P. (2003). Comparative

analysis of idiA and isiA transcription under iron starvation and

oxidative stress in Synechococcus elongatus PCC 7942 wild-type

and selected mutants. Arch. Microbiol. 180, 471–483.

Yu, J., and Vermaas, W. (1990). Transcript levels and synthesis of

photosystem II components in cyanobacterial mutants with inactivated

photosystem II genes. Plant Cell 2, 315–322.

Carotenoids and NPQ in Cyanobacteria 1007

Page 17: A Soluble Carotenoid Protein Involved in …A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in Cyanobacteria Adje´le´ Wilson, aGhada Ajlani, bJean-Marc

DOI 10.1105/tpc.105.040121; originally published online March 10, 2006; 2006;18;992-1007Plant Cell

Adjélé Wilson, Ghada Ajlani, Jean-Marc Verbavatz, Imre Vass, Cheryl A. Kerfeld and Diana KirilovskyCyanobacteria

A Soluble Carotenoid Protein Involved in Phycobilisome-Related Energy Dissipation in

 This information is current as of September 21, 2020

 

References /content/18/4/992.full.html#ref-list-1

This article cites 55 articles, 11 of which can be accessed free at:

Permissions https://www.copyright.com/ccc/openurl.do?sid=pd_hw1532298X&issn=1532298X&WT.mc_id=pd_hw1532298X

eTOCs http://www.plantcell.org/cgi/alerts/ctmain

Sign up for eTOCs at:

CiteTrack Alerts http://www.plantcell.org/cgi/alerts/ctmain

Sign up for CiteTrack Alerts at:

Subscription Information http://www.aspb.org/publications/subscriptions.cfm

is available at:Plant Physiology and The Plant CellSubscription Information for

ADVANCING THE SCIENCE OF PLANT BIOLOGY © American Society of Plant Biologists