12.02.2014 Views

Rapid evolutionary divergence of Photosystem I core subunits PsaA ...

Rapid evolutionary divergence of Photosystem I core subunits PsaA ...

Rapid evolutionary divergence of Photosystem I core subunits PsaA ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Photosynthesis Research 65: 131–139, 2000.<br />

© 2000 Kluwer Academic Publishers. Printed in the Netherlands.<br />

131<br />

Regular paper<br />

<strong>Rapid</strong> <strong>evolutionary</strong> <strong>divergence</strong> <strong>of</strong> <strong>Photosystem</strong> I <strong>core</strong> <strong>subunits</strong> <strong>PsaA</strong> and<br />

PsaB in the marine prokaryote Prochlorococcus<br />

Georg W.M. van der Staay 1,2 , Seung Yeo Moon-van der Staay 1,3 , Laurence Garczarek 1 &<br />

Frédéric Partensky 1,∗<br />

Observatoire Océanologique de Rosc<strong>of</strong>f, CNRS-UPR 9042 et Université Pierre et Marie Curie, BP 74, F-29682<br />

Rosc<strong>of</strong>f Cedex, France; 2 Present address: Botanisches Institut, Lehrstuhl III, Universität zu Köln, Gyrh<strong>of</strong>straße 15,<br />

D-50931 Köln, Germany; 3 Present address: Botanisches Institut, Lehrstuhl I, Universität zu Köln, Gyrh<strong>of</strong>straße<br />

15, D-50931 Köln, Germany; ∗ Author for correspondence (e-mail: partensky@sb-rosc<strong>of</strong>f.fr; fax: +33-98292324)<br />

Received 2 December 1999; accepted in revised form 9 August 2000<br />

Key words: cyanobacteria, <strong>Photosystem</strong> I, Synechococcus, prochlorophyte, Prochlorococcus<br />

Abstract<br />

The nucleotide sequences <strong>of</strong> the genes coding for the <strong>subunits</strong> <strong>of</strong> the <strong>Photosystem</strong> I (PS I) <strong>core</strong>, <strong>PsaA</strong> and PsaB<br />

were determined for the marine prokaryotic oxyphototrophs Prochlorococcus sp. MED4 (CCMP1378), P. marinus<br />

SS120 (CCMP1375) and Synechococcus sp. WH7803. Divergence <strong>of</strong> these sequences from those <strong>of</strong> both freshwater<br />

cyanobacteria and higher plants was remarkably high, given the conserved nature <strong>of</strong> <strong>PsaA</strong> and PsaB proteins. In<br />

particular, the <strong>PsaA</strong> <strong>of</strong> marine prokaryotes showed several specific insertions and deletions with regard to known<br />

<strong>PsaA</strong> sequences. Even in between the two Prochlorococcus strains, which correspond to two genetically different<br />

ecotypes with shifted growth irradiance optima, the sequence identity was only 80.2% for <strong>PsaA</strong> and 88.9% for<br />

PsaB. Possible causes and implications <strong>of</strong> the fast evolution rates <strong>of</strong> these two PS I <strong>core</strong> <strong>subunits</strong> are discussed.<br />

Abbreviations: Chl – chlorophyll; PS – photosystem<br />

Introduction<br />

PS I is a pigment–protein complex containing 11 different<br />

polypeptides in cyanobacteria and 17 in higher<br />

plants and binding about 90 molecules <strong>of</strong> Chl a (reviewed<br />

in Golbeck 1994; Chitnis et al. 1995; Chitnis<br />

1996; Scheller et al. 1997). Most <strong>of</strong> the pigments<br />

and components <strong>of</strong> the electron-transport chain in PS<br />

I are bound to the large <strong>subunits</strong> <strong>PsaA</strong> and PsaB<br />

(Schubert et al. 1997). These are hydrophobic proteins<br />

with molecular masses <strong>of</strong> 83–84 kDa, that are<br />

chloroplast-encoded in eukaryotes. The structure <strong>of</strong><br />

PS I structure has recently been determined at 4 Å<br />

resolution (Schubert et al. 1997). Both <strong>PsaA</strong> and PsaB<br />

have 11 transmembrane helices, that are organized in<br />

a pseudo two-fold symmetry. Based on a combination<br />

<strong>of</strong> X-ray diffraction studies and biochemical analyses<br />

on purified PS I reaction centers, a topological model<br />

for <strong>PsaA</strong> and PsaB has been proposed (Sun et al.<br />

1997). Comparison <strong>of</strong> their three-dimensional structures<br />

showed remarkable similarities in the topologies<br />

<strong>of</strong> PS I and PS II (Schubert et al. 1998; Klukas et al.<br />

1999). The six N-terminal helices <strong>of</strong> <strong>PsaA</strong> and PsaB,<br />

the antenna binding domain, are organized like the six<br />

helices <strong>of</strong> the inner PS II antenna proteins CP43 and<br />

CP47. The 5 C-terminal helices <strong>of</strong> <strong>PsaA</strong> and PsaB, that<br />

bind the components <strong>of</strong> the electron transport chain,<br />

are organized comparably to the PS II <strong>core</strong> proteins<br />

D1 and D2.<br />

PS I and PS II are remarkably conserved between<br />

prokaryotes and eukaryotes, whose chloroplasts are<br />

thought to have derived from an endosymbiotic cyanobacterial<br />

ancestor (Douglas 1998). With amino<br />

acid identities between cyanobacteria and chloroplasts<br />

<strong>of</strong> about 80% and many conservative amino acid replacements,<br />

<strong>PsaA</strong> and PsaB were considered to be


132<br />

very conserved proteins (Cantrell and Bryant 1987).<br />

This point <strong>of</strong> view has recently been challenged by the<br />

determination <strong>of</strong> the psaA and psaB sequences from<br />

the din<strong>of</strong>lagellate Heterocapsa triquetra (Zhang et al.<br />

1999), that showed a much lower degree <strong>of</strong> sequence<br />

conservation. This surprising result suggests that the<br />

available sequences are not sufficiently representative<br />

<strong>of</strong> the diversity <strong>of</strong> <strong>PsaA</strong> and PsaB proteins.<br />

Prochlorococcus (Chisholm et al. 1992) is a genus<br />

<strong>of</strong> planktonic photosynthetic prokaryotes very abundant<br />

in the ocean [see Partensky et al. (1999) for a<br />

review]. In contrast to other cyanobacteria, that use<br />

phycobiliproteins to collect light, the major light harvesting<br />

complexes in Prochlorococcus are membrane<br />

intrinsic proteins that bind mainly divinyl chlorophylls<br />

a and b. This genus is closely related to the marine<br />

AandBSynechococcus species, as shown by analysis<br />

<strong>of</strong> 16S rRNA, rpoC1, psbA, psbB, petB and petD<br />

gene sequences (Hess et al. 1995; Palenik and Swift<br />

1996; Urbach et al. 1998). PS I in Prochlorococcus<br />

shows several peculiar features. Recently, we have<br />

shown that the PS I <strong>subunits</strong> PsaF and PsaL in Prochlorococcus<br />

are longer and less conserved than in<br />

other organisms (van der Staay et al. 1998; van der<br />

Staay and Partensky 1999). Additionally, PS I in Prochlorococcus<br />

contains divinyl Chl b, which seems to<br />

be linked to the PS I <strong>core</strong> itself (Garczarek et al. 1998).<br />

To investigate the sequence conservation <strong>of</strong> the<br />

two large PS I <strong>subunits</strong> <strong>PsaA</strong> and PsaB, we have<br />

isolated and cloned these genes from two strains <strong>of</strong><br />

Prochlorococcus. P. marinus SS120 (CCMP1375) is<br />

a low light adapted strain which has a divinyl chlorophyll<br />

a to b ratio lower than 1 (Moore et al. 1995).<br />

This strain also contains some ‘normal’ (monovinyl)<br />

chlorophyll b (Partensky et al. 1993) and low amounts<br />

<strong>of</strong> phycoerythrin (Hess et al. 1996). Prochlorococcus<br />

sp. MED4 (CCMP1378) is adapted to grow at higher<br />

photon fluxes and has a divinyl chlorophyll a to b ratio<br />

<strong>of</strong> about 10 (Moore et al. 1995). Phylogenetic analyses<br />

using 16S rRNA consistently show MED4 belongs<br />

to the most evolved Prochlorococcus clade which includes<br />

all high light-adapted strains, whereas SS120<br />

and other low-light adapted strains belong to separate<br />

clusters genetically closer to the marine Synechococcus<br />

(Moore et al. 1998; Urbach et al. 1998; Moore and<br />

Chisholm 1999). As, prior to our study, there were<br />

no psaA and psaB gene sequences known from typical<br />

marine cyanobacteria, we have also determined their<br />

sequences from Synechococcus WH 7803. This strain<br />

is closely related to the other marine cyanobacteria<br />

and Prochlorococcus, as determined e.g. by analysis<br />

<strong>of</strong> DNA dependent RNA polymerase gene sequences<br />

(Palenik and Swift 1996). Our data confirm the close<br />

relatedness between Prochlorococcus and the marine<br />

Synechococcus. We show that in these three marine<br />

oxyphototrophs, the psaA and psaB genes evolved<br />

very rapidly.<br />

Materials and methods<br />

Isolation <strong>of</strong> clones<br />

The isolation <strong>of</strong> Lambda vectors containing the psaIpsaL<br />

gene clusters from Prochlorococcus sp. MED4<br />

and P. marinus SS120 has been previously described<br />

(van der Staay et al. 1998). The same vectors also<br />

contained the psaA-psaB genes (see ‘Results’).<br />

A genomic library <strong>of</strong> Synechococcus WH7803<br />

in Charon 35 was kindly provided by Dr D.<br />

Scanlan. A 499 bp psaB gene fragment from P.<br />

marinus SS120 was amplified by PCR with the<br />

primers CCGCTACATATTCGCCCAA and AGTGC-<br />

CGAAAACAGCATC. An initial denaturation at 94<br />

◦ C for 4 min was followed by 35 cycles <strong>of</strong> 94 ◦ C, 50<br />

◦ C and 72 ◦ C for 1 min each, with a final elongation at<br />

72 ◦ C for 7 min, in the presence <strong>of</strong> 1.5 mM Mg 2+ .The<br />

PCR product was cloned into the pCR 2.1 vector (In-<br />

Vitrogen). For probe labeling, the insert was excised<br />

with Eco RI. About 25 pg total DNA, corresponding to<br />

about 3 pg insert, was amplified by PCR in a volume<br />

<strong>of</strong> 50 µl as described above, except that the nucleotide<br />

mix was replaced by 5 µl <strong>of</strong> a fluorescein nucleotide<br />

mix (DuPont NEN). The labeled PCR product was<br />

used as a probe. Plaque hybridization proceeded as<br />

described in van der Staay et al. (1998). Hybridization<br />

occurred at 45 ◦ C, blots were washed with 5 ×<br />

SSC/0.1%SDS and 1 × SSC/0.1% SDS at 50 ◦ C. All<br />

other methods are described in van der Staay et al.<br />

(1998).<br />

Sequences used in this study<br />

The sequences determined in this study were deposited<br />

in the EMBL databank under the accession numbers<br />

AJ133190 for Synechococcus WH7803 psaA/B,<br />

AJ133191 for Prochlorococcus sp. MED4 psaA/B,<br />

AJ133192 for P. marinus SS120 psaA/B.<br />

The other sequences are available at GenBank under<br />

the following accession numbers: Anabaena variabilis:<br />

L26326, Mastigocladus laminosus (Fischerella<br />

PCC7605): AF038558, Synechococcus elongatus:


133<br />

X63768, Synechococcus PCC7002: M18165; Synechocystis<br />

PCC6803: D90906; Cyanophora paradoxa:<br />

U30821, Porphyra purpurea: U38804, Guillardia<br />

theta AF041468, Odontella sinensis Z67753, Heterocapsa<br />

triquetra: psaA AF130031, psaB AF130032;<br />

Euglena gracilis: X70810; Nephroselmis olivacea:AF<br />

137379; Chlamydomonas reinhardtii: psaA X05845,<br />

psaB X05848; Chlorella vulgaris: AB001684;<br />

Marchantia polymorpha: X04465; Pinus thunbergii<br />

(pine): D17510; Zea mays (maize): M11203; Oryza<br />

sativa (rice): X15901, Spinacia oleracea (spinach):<br />

X04131; Nicotiana tabacum (tobacco): Z00044.<br />

Protein alignments<br />

<strong>PsaA</strong> and PsaB sequences were aligned with the program<br />

ClustalX 1.64 (Thompson et al. 1994) using<br />

the PAM 250 matrix (Dayh<strong>of</strong>f et al. 1978). Alignments<br />

made with this matrix were very similar to the<br />

ones obtained using other matrices. Alignments were<br />

manually refined, keeping the number <strong>of</strong> gaps caused<br />

by insertions and deletions to a minimum. This program<br />

was also used to calculate the percentages <strong>of</strong><br />

sequence identity.<br />

Phylogenetic analysis<br />

Phylogenetic analysis used the parsimony and the<br />

neighbor-joining distance algorithms. PHYLIP Version<br />

3.5c (Felsenstein 1992) was used for all analyses.<br />

Distance matrices were constructed with the program<br />

PROTDIST using the PAM 250 matrix (Dayh<strong>of</strong>f et al.<br />

1978). Neighbor-joining distance trees were determined<br />

with the neighbor-joining (NEIGHBOR) method<br />

(Saitou and Nei 1987). Bootstrap analyses (SEQ-<br />

BOOT) with 100 replicates were applied to test the<br />

stability <strong>of</strong> the tree topology.<br />

Results<br />

Further upstream sequencing <strong>of</strong> a genomic clone containing<br />

the previously described genes coding for the<br />

PS I <strong>subunits</strong> PsaL and PsaI <strong>of</strong> Prochlorococcus sp.<br />

MED4 (van der Staay et al. 1998) revealed that this<br />

genomic fragment also contained the genes for the<br />

PS I <strong>core</strong> <strong>subunits</strong> <strong>PsaA</strong> and PsaB. The latter were<br />

separated from psaL and psaI by a gene encoding a<br />

protein that showed the closest similarity to a dolicholphosphate<br />

mannosyltransferase from Aquifex aeolicus<br />

(GenBank accession number AE000762), as determined<br />

by partial sequencing. Analogously, psaA and<br />

psaB from P. marinus SS120 were located on a genomic<br />

clone containing psaL and psaI. The psaA and<br />

psaB genes <strong>of</strong> the marine Synechococcus sp. WH<br />

7803 were isolated from a genomic library in lambda<br />

Charon 35, using a psaB probe from P. marinus<br />

SS120. As found in a study on the psbB and petB/D<br />

genes <strong>of</strong> several strains <strong>of</strong> Prochlorococcus and marine<br />

Synechococcus (Urbach et al. 1998), Prochlorococcus<br />

shows a third codon bias towards A and T<br />

(not shown), with a more pronounced bias in Prochlorococcus<br />

sp. MED4 than SS120. In contrast, Synechococcus<br />

WH7803 has a bias towards G and C at<br />

the third codon position, like reported for several other<br />

marine Synechococcus strains (Urbach et al. 1998).<br />

An alignment <strong>of</strong> the deduced amino acid sequences<br />

is shown in Figure 1. The corresponding proteins<br />

<strong>of</strong> the cyanobacterium Synechocystis PCC 6803, for<br />

which a topological model <strong>of</strong> the PS I <strong>core</strong> complex<br />

has been presented (Sun et al. 1997), are included in<br />

the alignment. The numbering <strong>of</strong> the putative membrane<br />

spanning helices and the extramembrane loops<br />

is done according to the model proposed for Synechocystis.<br />

The deduced <strong>PsaA</strong> proteins from P. marinus<br />

SS120, Prochlorococcus sp. MED4 and Synechococcus<br />

WH7803 are, respectively, 773, 767 and 767<br />

amino acids long. This is longer than <strong>PsaA</strong> sequences<br />

reported for other organisms, which vary between 732<br />

and 761 amino acids, with most species having a <strong>PsaA</strong><br />

750–755 amino acids long. None <strong>of</strong> the insertions or<br />

deletions are situated in the 11 potential membrane<br />

spanning regions, nor in regions known to be involved<br />

in the binding <strong>of</strong> components <strong>of</strong> the electron transport<br />

chain (Chitnis 1996). The <strong>PsaA</strong> proteins <strong>of</strong> all three<br />

marine prokaryotes contain a 13 amino acid insertion<br />

between the transmembrane helices III and IV (extramembrane<br />

loop D), the least conserved area <strong>of</strong> this<br />

protein (Figure 1A). A conserved insertion <strong>of</strong> three<br />

amino acids (F-P-A) is located between the helices IV<br />

and V (loop E). This loop also shows a low degree <strong>of</strong><br />

sequence identity. <strong>PsaA</strong> proteins from both Prochlorococcus<br />

strains have an additional insertion between<br />

helices VII and VIII and a deletion <strong>of</strong> 10 amino acids<br />

between the helices XI and X (loop J) not found in any<br />

other species.<br />

PsaB from P. marinus SS120 is with 747 amino<br />

acids the longest known PsaB sequence with the exception<br />

<strong>of</strong> the one <strong>of</strong> Heterocapsa triquetra, which<br />

contains 776 amino acids (Zhang et al. 1999). The<br />

PsaB proteins from Prochlorococcus sp. MED4 and<br />

Synechococcus sp. WH7803 fall, with 742 and 738<br />

amino acids, respectively, within the usual range <strong>of</strong>


134<br />

Figure 1. Alignment <strong>of</strong> the <strong>PsaA</strong> (A) and PsaB (B) from the marine prokaryotes Prochlorococcus sp. MED4, Prochlorococcus marinus SS120<br />

and Synechococcus WH7803. For comparison, the corresponding sequences <strong>of</strong> the freshwater cyanobacterium Synechocystis PCC6803 is<br />

included. The putative membrane spanning helices I–XI and the extramembrane loops A–L, determined and numbered based on the analogy<br />

to the model presented by Sun et al. (1997) are indicated. The cysteines that ligate the F x electron acceptor are indicated with a block. For<br />

positions where amino acids were identical, only the amino acid for Prochlorococcus sp. MED4 is given. Gaps are indicated by -.<br />

731–743 amino acids reported for most other organisms.<br />

Most diverse is loop E between the helices IV<br />

and V and loop H between helices VII and VIII, the<br />

areas that are the least conserved in all PsaB pro-


135<br />

Figure 1. Continued.<br />

teins. The C-terminal domains <strong>of</strong> <strong>PsaA</strong> and PsaB, from<br />

helices VIII to XI, that are thought to bind the components<br />

<strong>of</strong> the electron transport chain including the<br />

F x -binding domain, are the most conserved regions.<br />

In all phylogenetic analyses using the concatenated<br />

<strong>PsaA</strong> and PsaB sequences, the Prochlorococcus<br />

strains grouped together with the marine Synechococcus<br />

strain (Figure 2), separately from the other cyanobacteria.<br />

Analyses <strong>of</strong> either <strong>PsaA</strong> or PsaB lead<br />

to comparable results. Inclusion <strong>of</strong> the sequences<br />

from the din<strong>of</strong>lagellate Heterocapsa triquetra resulted<br />

in a destabilization <strong>of</strong> the trees, due to the high<br />

<strong>divergence</strong> <strong>of</strong> these sequences. Depending on small<br />

differences in the alignments and the inclusion or<br />

omission <strong>of</strong> gaps, Heterocapsa grouped together with<br />

Guillardia and Odontella or with the Prochlorococcus/Synechococcus<br />

cluster (not shown).


136<br />

Figure 2. Phylogenetic analysis <strong>of</strong> concatenated <strong>PsaA</strong> and PsaB proteins using the neighbor-joining distance method. For this analysis, all<br />

gaps created by the alignment were excluded. The first number at nodes represent the bootstrap percentage from 100 replicates for the<br />

neighbor-joining distance method, the second number indicates the bootstrap support for this node found by the maximum parsimony method.<br />

Values below 50% are indicated by – or not shown. Synechocystis PCC6803 was arbitrarily chosen as an outgroup. The scale bar indicates<br />

substitutions per amino acid position.<br />

The neighbor-joining <strong>PsaA</strong>/B tree exhibits unexpectedly<br />

long branches within the Synechocococcus/Prochlorococcus<br />

cluster (Figure 2), despite the<br />

fact that these species are phylogenetically close. This<br />

suggests considerable differences in the <strong>evolutionary</strong><br />

rates <strong>of</strong> the psaA/B operon both in between marine<br />

cyanobacteria and between these and the other phototrophs.<br />

The high <strong>divergence</strong> in PS I protein sequences<br />

between these three strains is confirmed by the comparison<br />

<strong>of</strong> sequence identities (Table 1). The <strong>PsaA</strong> sequences<br />

are more divergent than the PsaB sequences.<br />

Prochlorococcus sp. SS120 shows the highest deviation<br />

<strong>of</strong> both the <strong>PsaA</strong> and PsaB sequences. For<br />

both <strong>PsaA</strong> and PsaB, Heterocapsa showed by far the<br />

highest deviation (not shown). Since the larger number<br />

<strong>of</strong> insertions and deletions in these proteins from Heterocapsa<br />

complicated the alignment, leading to more<br />

gaps, they were excluded from the Table. Compared<br />

to the species indicated in Table 1, <strong>PsaA</strong> from H. triquetra<br />

showed sequence identities between 51.4% to<br />

P. marinus SS120 and 56.6% to Cyanophora paradoxa.<br />

For Heterocapsa PsaB, the sequence identity<br />

ranged between 51.3% to Prochlorococcus sp. MED4<br />

and 55.9% to Cyanophora paradoxa.<br />

Discussion<br />

The amino acid sequence analysis <strong>of</strong> the PS I <strong>subunits</strong><br />

<strong>PsaA</strong> and PsaB from one Synechococcus strain<br />

(WH 7803), representative <strong>of</strong> the Marine-Cluster A<br />

(Waterbury and Rippka 1989), and two Prochlorococcus<br />

strains, representative <strong>of</strong> the low-light (SS120)<br />

and high-light adapted (MED4) ecotypes (Moore et al.<br />

1995, 1998; Urbach et al. 1998), are in agreement with<br />

previous molecular studies showing that these marine<br />

oxyphototrophs are closely related and probably<br />

evolved from a common ancestor (Hess et al. 1995;<br />

Urbach et al. 1998; Honda et al. 1999; Turner et al.<br />

1999).<br />

The high <strong>evolutionary</strong> rate <strong>of</strong> the <strong>PsaA</strong> and PsaB<br />

sequences from all the three marine oxyphototrophs


137<br />

Table 1. Sequence identity <strong>of</strong> <strong>PsaA</strong> and PsaB <strong>of</strong> cyanobacteria and chloroplasts. Sequences were aligned with the program<br />

ClustalX and manually refined. The percentage sequence identity was calculated from a distance matrix constructed with the<br />

program ClustalX. Positions with gaps were excluded from the alignment. Upper triangle: <strong>PsaA</strong>, lower triangle: PsaB<br />

<strong>PsaA</strong><br />

1 2 3 4 5 6 7 8 9 10 11 12 13 14<br />

1 Prochlorococcus sp. MED4 – 80.2 75.7 69.9 70.3 70.9 69.1 69.7 70.9 69.4 67.2 70.2 69.9 68.8<br />

2 Prochlorococcus marinus SS120 88.9 – 70.8 68.2 67.1 68.9 65.8 66.4 67.9 66.8 65.4 67.5 67.4 65.8<br />

3 Synechococcus WH7803 81.5 80.6 – 77.6 77.7 78.4 75.8 77.7 77.3 75.1 71.3 76.1 77.7 74.7<br />

4 Anabaena variabilis 76.1 74.5 81.9 – 94.4 87.6 82.9 85.0 82.6 80.9 77.2 82.6 83.2 80.6<br />

5 Mastigocladus 74.4 74.5 80.3 90.6 – 87.4 81.4 84.4 82.6 80.7 76.7 81.9 82.6 80.9<br />

6 Synechococcus elongatus 76.8 76.0 81.6 89.6 87.7 – 84.3 87.9 84.3 85.0 79.3 85.1 86.1 85.6<br />

7 Synechococcus PCC7002 79.0 78.5 87.6 87.0 85.5 87.1 – 88.4 81.5 82.1 75.9 81.8 81.5 79.0<br />

8 Synechocystis PCC6803 76.8 76.8 84.1 86.4 85.6 87.1 92.6 – 83.1 84.2 75.7 83.7 83.6 80.9<br />

9 Cyanophora paradoxa 74.6 74.8 82.0 82.3 82.4 84.8 86.4 84.8 – 85.6 79.9 85.4 85.0 82.4<br />

10 Porphyra purpurea 74.4 74.1 80.4 80.5 81.1 81.7 84.4 83.5 83.5 – 78.9 86.2 85.1 82.3<br />

11 Euglena gracilis 70.9 71.3 76.4 77.4 76.7 79.2 80.5 78.3 80.4 79.0 – 82.5 82.2 79.3<br />

12 Chlamydomonas reinhardtii 73.0 73.0 79.7 79.0 79.0 81.6 82.7 80.9 82.2 81.5 80.4 – 89.8 86.8<br />

13 Marchantia polymorpha 73.8 73.6 81.2 79.3 79.3 81.8 84.1 82.3 83.8 84.5 83.2 85.0 – 93.2<br />

14 Spinacia oleracea 72.7 72.2 79.8 77.9 77.5 80.5 81.8 79.8 81.7 83.2 82.3 83.8 92.4 –<br />

PsaB<br />

is unexpected. Most striking is the low percent identity<br />

(80.2%) between the <strong>PsaA</strong> proteins <strong>of</strong> the two<br />

Prochlorococcus strains. This percent identity is as<br />

low as the identity between freshwater cyanobacteria<br />

and chloroplasts, that are thought to have separated<br />

about one billion years ago. By comparison, the respective<br />

16S rRNA sequences <strong>of</strong> these ecotypes are<br />

98.4% identical (Urbach et al. 1998). One may wonder<br />

which <strong>evolutionary</strong> constraint led to this rapid <strong>divergence</strong><br />

<strong>of</strong> <strong>PsaA</strong> (and to a lesser extent PsaB). It has<br />

been observed for other proteins that their <strong>evolutionary</strong><br />

rate differs among taxa. (Lockhart et al. 1996;<br />

Lopez et al. 1999). The underlying mechanism is not<br />

well understood yet. For Prochlorococcus, themost<br />

dramatic difference (with regard to photosynthesis)<br />

between the respective niches <strong>of</strong> these ecotypes in the<br />

field is certainly the amount <strong>of</strong> available light (Campbell<br />

and Vaulot 1993; Moore et al. 1998; Moore and<br />

Chisholm 1999). Besides having shifted growth irradiance<br />

optima and divinyl Chl a to divinyl Chl b ratios<br />

(Partensky et al. 1993, Moore et al. 1995), the Prochlorococcus<br />

ecotypes MED4 and SS120 also have<br />

dissimilar thylakoid protein pr<strong>of</strong>iles (Partensky et al.<br />

1997; Garczarek et al. 1998). The antenna system<br />

is probably the component <strong>of</strong> the photosynthetic apparatus<br />

the most differentiated between these strains<br />

(LaRoche et al. 1996). We recently discovered that<br />

SS120 possesses multiple pcb genes encoding seven<br />

different antenna proteins whereas MED4 possesses<br />

a single pcb gene (Garczarek et al. 2000). Another<br />

major difference between SS120 and MED4 is the<br />

presence <strong>of</strong> a large gene cluster implicated in the<br />

biosynthesis <strong>of</strong> phycoerythrin and its associated phycobilins<br />

in the latter but not the former strain (Hess et<br />

al. 1999). Thus, it seems that light has been a major<br />

driving force in the evolution <strong>of</strong> several key photosynthetic<br />

proteins in Prochlorococcus and this might be<br />

the case for the PS I <strong>core</strong> as well. In the open ocean,<br />

the wavelengths <strong>of</strong> photons found at depths below<br />

100 m (i.e. in the environment where the low light adapted<br />

Prochlorococcus ecotype thrives), are narrowly<br />

centered around 470 nm (Morel 1978). These photons<br />

are most efficiently captured by the divinyl Chl b and<br />

much less by divinyl Chl a. So to cope with these low<br />

blue light levels, the low-light adapted ecotypes must<br />

have evolved to optimize the capture <strong>of</strong> photons by PS<br />

I, e.g. either by binding Chl b molecules (Garczarek et<br />

al. 1998) or by recruiting as an antenna one or several<br />

<strong>of</strong> the multiple divinyl Chl a/b-binding Pcb proteins<br />

present in the cell. Comparative analysis <strong>of</strong> <strong>PsaA</strong>/B<br />

and other proteins associated with the photosystems<br />

from other representatives <strong>of</strong> the low-light and high<br />

light ecotypes should help us to get more insight in the<br />

underlying adaptation processes.<br />

The analysis <strong>of</strong> PS I-enriched fractions from<br />

MED4 and SS120 strains previously showed that they


138<br />

have similar PS I protein pr<strong>of</strong>iles, but both possess<br />

two proteins with apparent molecular mass <strong>of</strong> 21 and<br />

25 kDa, which have no equivalent in cyanobacteria,<br />

including the marine Synechococcus WH 8103 (Garczarek<br />

et al. 1998). These proteins were identified as<br />

the PS I <strong>subunits</strong> PsaF and PsaL, respectively, and<br />

their anomalous length was found to be due to specific<br />

gene insertions (van der Staay et al. 1998; van<br />

der Staay and Partensky 1999). Here we demonstrate<br />

that the two large <strong>core</strong> proteins <strong>PsaA</strong> and PsaB in<br />

Prochlorococcus show some specific insertions, too,<br />

but also deletions. Not surprisingly, the C-terminal<br />

part <strong>of</strong> both these proteins is the most conserved. This<br />

part is thought to bind the components <strong>of</strong> the electron<br />

transport chain. Therefore, a differentiation <strong>of</strong> this region<br />

is severely restricted. In the remaining part <strong>of</strong> the<br />

regions, variations are more tolerated, allowing insertions<br />

and deletions. The main insertions in <strong>PsaA</strong> <strong>of</strong> the<br />

three marine prokaryotes occur in loop D, located in<br />

the lumen [compared to topographic model <strong>of</strong> the PS<br />

I <strong>core</strong> proteins proposed by Sun et al. (1997)] and the<br />

cytoplasmic loop E. Both these loops contain less conserved<br />

regions in all species. Whereas no function has<br />

been assigned to loop D, loop E might interact with the<br />

subunit PsaE. An insertion in <strong>PsaA</strong> <strong>of</strong> Prochlorococcus,<br />

but not <strong>of</strong> other species, is located in the luminal<br />

loop H. This loop is thought to interact with PsaF. It is<br />

tempting to speculate that both this insertion in <strong>PsaA</strong><br />

and the ones in PsaF from Prochlorococcus might be<br />

involved in the interaction between these proteins. A<br />

significant deletion <strong>of</strong> 10 amino acids is present in<br />

the loop J <strong>of</strong> <strong>PsaA</strong> from Prochlorococcus. Thecorresponding<br />

loop in PsaB was shown to be involved in<br />

the interaction with soluble electron transporters (Sun<br />

et al. 1999). Assuming a pseudo tw<strong>of</strong>old symmetry <strong>of</strong><br />

<strong>PsaA</strong> and PsaB in the PS I complex, loop J might have<br />

a similar function in <strong>PsaA</strong>. In PsaB, major insertions<br />

occur in the cytoplasmic loop E and the luminal loop<br />

H. Insertions in both <strong>of</strong> these loops, that are the least<br />

conserved in all species, are found in PsaB from other<br />

species, too.<br />

Our results on marine cyanobacteria, in addition to<br />

the ones obtained with the din<strong>of</strong>lagellate Heterocapsa<br />

triquetra (Zhang et al. 1999), show that PS I <strong>core</strong> proteins<br />

can be more variable than previously assumed.<br />

We demonstrated that <strong>PsaA</strong> <strong>of</strong> marine cyanobacteria<br />

has characteristic features that distinguish them from<br />

the corresponding proteins <strong>of</strong> all other groups. In addition,<br />

based on the characteristic insertion and deletion<br />

in the <strong>PsaA</strong> sequence, and the much lower GC content<br />

in the psaA and psaB genes, representatives <strong>of</strong> the<br />

genus Prochlorococcus can probably be distinguished<br />

from marine Synechococcus. Therefore, these genes<br />

might constitute useful genetic markers for studies on<br />

the biodiversity <strong>of</strong> natural picoplanktonic communities.<br />

To get a more complete picture, sequences from<br />

other organisms should be obtained. Of special interest<br />

would be the cyanobacterium Gloeobacter violaceus.<br />

It is considered as a representative <strong>of</strong> the most primitive<br />

group <strong>of</strong> cyanobacteria (Honda et al. 1999; Turner<br />

et al. 1999) and, like Prochlorococcus (Garczarek et<br />

al. 1998), its PS I lacks the characteristic fluorescence<br />

at 77 K (Koenig and Schmidt 1995). PS I from Prochlorococcus<br />

appears to be unique, because it binds a<br />

divinyl form <strong>of</strong> Chl a, and probably divinyl Chl b as<br />

well (Garczarek et al. 1998). Since Chl b has also been<br />

reported in the PS I <strong>of</strong> Prochloron and Prochlorothrix<br />

(Hiller and Larkum 1985; van der Staay et al. 1992),<br />

obtaining sequences from the <strong>PsaA</strong> and PsaB <strong>of</strong> these<br />

two organisms might cast some light on the potential<br />

effect at the protein sequence level <strong>of</strong> the kind <strong>of</strong><br />

bound pigment. Also <strong>of</strong> particular interest in this context<br />

is Acaryochloris marina, an oxygenic prokaryote,<br />

the PS I <strong>of</strong> which contains almost exclusively Chl d<br />

(Hu et al. 1998).<br />

References<br />

Campbell L and Vaulot D (1993) Photosynthetic picoplankton community<br />

structure in the subtropical North Pacific Ocean near<br />

Hawaii (station ALOHA) Deep-Sea Res 40: 2043–2060<br />

Cantrell A and Bryant DA (1987) Molecular cloning and nucleotide<br />

sequence <strong>of</strong> the psaA and psaB genes <strong>of</strong> the cyanobacterium<br />

Synechococcus sp PCC 7002. Plant Mol Biol 9: 453–468<br />

Chisholm SW, Frankel SL, Goericke R, Olson RJ, Palenik B,<br />

Waterbury JB, West-Johnsrud L and Zettler ER (1992) Prochlorococcus<br />

marinus nov. gen. nov. sp.: An oxyphototrophic<br />

marine prokaryote containing divinyl chlorophyll a and b. Arch<br />

Microbiol 157: 297–300<br />

Chitnis PR (1996) <strong>Photosystem</strong> I – update on photosynthetic electron<br />

transport. Plant Physiol 111:<br />

Chitnis PR, Xu Q, Chitnis VP and Nechushtai R (1995) Function<br />

and organization <strong>of</strong> <strong>Photosystem</strong> I polypeptides. Photosynth Res<br />

44: 23–40 661–669<br />

Dayh<strong>of</strong>f MO, Schwartz RM and Orcutt BC (1978) A model <strong>of</strong> <strong>evolutionary</strong><br />

change in proteins. In Dayh<strong>of</strong>f MO (ed) Atlas <strong>of</strong> Protein<br />

Sequence and Structure, pp 345–352. Natural Biomedical<br />

Research Foundation, Washington, DC<br />

Douglas SE (1998) Plastid evolution: Origins, diversity, trends. Curr<br />

Opin Genet Develop 8: 655–661<br />

Felsenstein J (1992) PHYLIP (phylogeny interference package).<br />

University <strong>of</strong> Washington, Seattle<br />

Garczarek L, van der Staay GWM, Thomas JC and Partensky F<br />

(1998) Isolation and characterization <strong>of</strong> <strong>Photosystem</strong> I from<br />

two strains <strong>of</strong> the marine oxychlorobacterium Prochlorococcus.<br />

Photosynth Res 56: 131–141


139<br />

Garczarek L, Hess W, Holtzendorff J, van der Staay, GWM and<br />

Partensky, F (2000) Multiplication <strong>of</strong> antenna genes as a major<br />

adaptation process to growth at low light in a marine prokaryote.<br />

Proc Natl Acad Sci USA 97: 4098–4101<br />

Hess WR, Weihe A, Loiseaux-de Goer S, Partensky F and Vaulot D<br />

(1995) Characterization <strong>of</strong> the single psbA gene <strong>of</strong> Prochlorococcus<br />

marinus CCMP1375 (Prochlorophyta). Plant Mol Biol 27:<br />

1189–1196<br />

Hess WR, Partensky F, van der Staay GWM, Garcia-Fernandez JM,<br />

Börner T and Vaulot D (1996) Coexistence <strong>of</strong> phycoerythrin and<br />

a chlorophyll a/b antenna in a marine prokaryote. Proc Natl Acad<br />

Sci USA 93: 11126–11130<br />

Hess WR, Steglich C, Lichtlé C and Partensky F (1999) Phycoerythrins<br />

<strong>of</strong> the oxyphotobacterium Prochlorococcus marinus<br />

are associated to the thylakoid membrane and are encoded by<br />

a single large gene cluster. Plant Mol Biol 40: 507–521<br />

Hiller RD and Larkum AWD (1985) The chlorophyll–protein complexes<br />

<strong>of</strong> Prochloron sp. (Prochlorophyta). Biochim Biophys<br />

Acta 806: 107–115<br />

Honda D, Yokota A and Sugiyama J (1999) Detection <strong>of</strong> seven<br />

major <strong>evolutionary</strong> lineages in cyanobacteria based on the 16S<br />

rRNA gene sequence analysis with new sequences <strong>of</strong> five marine<br />

Synechococcus strain. J Mol Evol 48: 723–739<br />

Hu Q, Miyashita H, Iwasaki I, Kurano N, Miyachi S, Iwaki M and<br />

Itoh S (1998) A <strong>Photosystem</strong> I reaction center driven by chlorophyll<br />

d in oxygenic photosynthesis. Proc Natl Acad Sci USA 95:<br />

13319–13323<br />

Klukas O, Schubert W-D, Jordan P, Krauß N, Fromme P, Witt HAT<br />

and Saenger W (1999) <strong>Photosystem</strong> I, an improved model <strong>of</strong> the<br />

stromal <strong>subunits</strong> PsaC, PsaD and PsaE. J Biol Chem 274: 7351–<br />

7360<br />

Koenig F and Schmidt M (1995) Gloeobacter violaceus –investigation<br />

<strong>of</strong> an unusual photosynthetic apparatus – absence <strong>of</strong> the<br />

long wavelength emission <strong>of</strong> <strong>Photosystem</strong> I in 77 K fluorescence<br />

spectra. Physiol Plant 94: 621–628<br />

Lockhart PJ, Larkum AWD, Steel MA, Waddell PJ and Penny D<br />

(1996) Evolution <strong>of</strong> chlorophyll and bacteriochlorophyll: The<br />

problem <strong>of</strong> invariant sites in sequence analysis. Proc Natl Acad<br />

Sci USA 93: 1930–1934<br />

Lopez P, Forterre P and Philippe H (1999) The root <strong>of</strong> the tree <strong>of</strong><br />

life in the light <strong>of</strong> the covarion model. J Mol Evol 49: 496–508<br />

Moore LR and Chisholm SW (1999) Photophysiology <strong>of</strong> the marine<br />

cyanobacterium Prochlorococcus: Ecotypic differences among<br />

cultured isolates. Limnol Ocean 44: 628–638<br />

Moore LR, Goericke R and Chisholm SW (1995) Comparative<br />

physiology <strong>of</strong> Synechococcus and Prochlorococcus: Influence<br />

<strong>of</strong> light and temperature on growth, pigments, fluorescence and<br />

absorptive properties. Mar Ecol Prog Ser 116: 259–275<br />

Moore LR, Rocap G and Chisholm SW (1998) Physiology and<br />

molecular phylogeny <strong>of</strong> coexisting Prochlorococcus ecotypes.<br />

Nature 393: 464–467<br />

Morel A (1978) Available, usable and stored radiant energy in<br />

relation to marine photosynthesis. Deep-Sea Res 25: 673–688<br />

Palenik B and Swift H (1996) Cyanobacterial evolution and prochlorophyte<br />

diversity as seen in DNA-dependent RNA polymerase<br />

gene sequences. J Phycol 32: 638–646<br />

Partensky F, Hoepffner N, Li WKW, Ulloa O and Vaulot D<br />

(1993) Photoacclimation <strong>of</strong> Prochlorococcus sp. (Prochlorophyta)<br />

strains isolated from the North Atlantic and the Mediterranean<br />

Sea. Plant Physiol 101: 285–296<br />

Partensky F, LaRoche J, Wyman K and Falkowski P (1997) The<br />

divinyl chlorophyll a/b-protein complexes <strong>of</strong> two strains <strong>of</strong> the<br />

oxyphototrophic marine prokaryote Prochlorococcus – characterization<br />

and response to changes in growth irradiance. Photosynth<br />

Res 51: 209–222<br />

Partensky F, Hess WR, and Vaulot D (1999) Prochlorococcus, a<br />

marine photosynthetic prokaryote <strong>of</strong> global significance. Microbiol<br />

Mol Biol Rev 63: 106–127<br />

Saitou N and Nei M (1987) The neighbor-joining method: A new<br />

method for reconstructing phylogenetic trees. Mol Biol Evol 4:<br />

406–425<br />

Scheller HV, Naver H and Møller BL (1997) Molecular aspects <strong>of</strong><br />

<strong>Photosystem</strong> I. Physiol Plant 100: 842–851<br />

Schubert W-D, Klukas O, Krauß N, Saenger W, Fromme P and Witt<br />

HT (1997) <strong>Photosystem</strong> I <strong>of</strong> Synechococcus elongatus at 4 Å<br />

resolution: Comprehensive structure analysis. J Mol Biol 272:<br />

741–769<br />

Schubert W-D, Klukas O, Saenger W, Witt HAT, Fromme P and<br />

Krauss N (1998) A common ancestor for oxygenic and anoxygenic<br />

photosynthetic systems – a comparison based on the<br />

structural model <strong>of</strong> <strong>Photosystem</strong> I. J Mol Biol 280: 297–314<br />

Sun J, Xu Q, Chitnis VP, Jin P and Chitnis PR (1997) Topography<br />

<strong>of</strong> the <strong>Photosystem</strong> I <strong>core</strong> proteins <strong>of</strong> the cyanobacterium<br />

Synechocystis sp. PCC 6803. J Biol Chem 272: 21793–21802<br />

SunJ,XuW,HervásM,NavarroJA,DeLaRosamandChitnis<br />

PR (1999) Oxidizing side <strong>of</strong> the cyanobacterial <strong>Photosystem</strong> –<br />

Evidence for interaction between the electron donor proteins and<br />

a luminal surface helix <strong>of</strong> the PsaB subunit. J Biol Chem 274:<br />

19048–19054<br />

Thompson JD, Higgins DG and Gibson DJ (1994) Clustal W: Improving<br />

the sensitivity <strong>of</strong> multiple sequence alignment through<br />

sequence weighting, position specific gap penalties and matrix<br />

choice. Nucleic Acids Res 22: 4673–4680<br />

Turner S, Pryer KM, Miao VPW and Palmer JD (1999) Investigating<br />

deep phylogenetic relationships among cyanobacteria and<br />

plastids by small subunit rRNA sequence analysis. J Eukaryot<br />

Microbiol 46: 327–338<br />

Urbach E, Scanlan DJ, Distel DL, Waterbury JB and Chisholm SW<br />

(1998) <strong>Rapid</strong> diversification <strong>of</strong> marine picophytoplankton with<br />

dissimilar light harvesting structures inferred from sequences <strong>of</strong><br />

Prochlorococcus and Synechococcus (cyanobacteria). J Mol Evol<br />

46: 188–201<br />

van der Staay GWM, Brouwer A, Baard RL, van Mourik F and<br />

Matthijs HCP (1992) Separation <strong>of</strong> <strong>Photosystem</strong>s I and II from<br />

the oxychlorobacterium (prochlorophyte) Prochlorothrix hollandica<br />

and association <strong>of</strong> chlorophyll b binding antennae with<br />

<strong>Photosystem</strong> II. Biochim Biophys Acta 1102: 220–228<br />

van der Staay GWM, Moon-van der Staay SY, Garczarek L and<br />

Partensky F (1998) Characterization <strong>of</strong> the <strong>Photosystem</strong> I <strong>subunits</strong><br />

PsaI and PsaL from two strains <strong>of</strong> the marine oxyphototrophic<br />

prokaryote Prochlorococcus. Photosynth Res 57: 183–<br />

192<br />

van der Staay GWM and Partensky F (1999) The 21 kDa protein<br />

associated with <strong>Photosystem</strong> I in Prochlorococcus marinus is<br />

the PsaF protein (Accession No. AJ131438). (PGR99-067). Plant<br />

Physiol 120: 339<br />

Waterbury JB and Rippka R (1989) Order Chroococcales Wettstein<br />

1924, Emend. Rippka et al. 1979. In: Kreig NR and Holt JB (eds)<br />

Bergey’s Manual <strong>of</strong> Systematic Bacteriology, pp 1728-01746.<br />

Williams and Wilkins, Baltimore<br />

Zhang Z, Green BR and Cavalier-Smith T (1999) Single gene circles<br />

in din<strong>of</strong>lagellate chloroplast genomes. Nature 400: 155–159

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!