toxic dinoflagellate alexandrium catenella revealed with nuclear

30
1 Unexpected Genetic Diversity among and within Populations of the 1 Toxic Dinoflagellate Alexandrium catenella Revealed with Nuclear 2 Microsatellite Markers 3 4 Estelle Masseret,* 1 Daniel Grzebyk, 1 Satoshi Nagai, 2 Benjamin Genovesi, 1,2 Bernard Lasserre, 1 5 Mohamed Laabir, 1 Yves Collos, 1 André Vaquer, 1 Patrick Berrebi 3 6 7 Université Montpellier II, CNRS, Ifremer, UMR 5119 Ecosystèmes lagunaires 5119, 8 cc093, place E. Bataillon, 34095 Montpellier, France 1 9 National Research Institute of Fisheries and Environment of the Inland Sea, Maruishi 2- 10 17-5, Hatsukaichi, Hiroshima 739-0452, Japan 2 11 Université Montpellier II, CNRS, UMR 5554 Institut des Sciences de l'Evolution, cc065, 12 place E. Bataillon, 34095 Montpellier, France 3 13 14 KEY WORDS: 15 Alexandrium catenella; dinoflagellates; microsatellites; population genetics; phylogenetic 16 diversity. 17 RUNNING TITLE: French-Japanese A. catenella population genetics 18 19 *Corresponding author. mailing address: UMR UM2-CNRS-Ifremer 5119 Ecosystèmes 20 lagunaires, Université Montpellier II, cc093, Place Eugène Bataillon, 34095 Montpellier, France. 21 Phone: 33 467144762. Fax: 33 467143719, E-mail: [email protected] 22 ACCEPTED Copyright © 2009, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved. Appl. Environ. Microbiol. doi:10.1128/AEM.01686-08 AEM Accepts, published online ahead of print on 5 February 2009 Copyright © 2009, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved. Appl. Environ. Microbiol. doi:10.1128/AEM.01686-08 AEM Accepts, published online ahead of print on 5 February 2009

Upload: independent

Post on 30-Mar-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

1

Unexpected Genetic Diversity among and within Populations of the 1

Toxic Dinoflagellate Alexandrium catenella Revealed with Nuclear 2

Microsatellite Markers 3

4

Estelle Masseret,*1 Daniel Grzebyk,

1 Satoshi Nagai,

2 Benjamin Genovesi,

1,2 Bernard Lasserre,

1 5

Mohamed Laabir,1 Yves Collos,

1 André Vaquer,

1 Patrick Berrebi

3 6

7

Université Montpellier II, CNRS, Ifremer, UMR 5119 Ecosystèmes lagunaires 5119, 8

cc093, place E. Bataillon, 34095 Montpellier, France 1 9

National Research Institute of Fisheries and Environment of the Inland Sea, Maruishi 2-10

17-5, Hatsukaichi, Hiroshima 739-0452, Japan 2 11

Université Montpellier II, CNRS, UMR 5554 Institut des Sciences de l'Evolution, cc065, 12

place E. Bataillon, 34095 Montpellier, France 3 13

14

KEY WORDS: 15

Alexandrium catenella; dinoflagellates; microsatellites; population genetics; phylogenetic 16

diversity. 17

RUNNING TITLE: French-Japanese A. catenella population genetics 18

19

*Corresponding author. mailing address: UMR UM2-CNRS-Ifremer 5119 Ecosystèmes 20

lagunaires, Université Montpellier II, cc093, Place Eugène Bataillon, 34095 Montpellier, France. 21

Phone: 33 467144762. Fax: 33 467143719, E-mail: [email protected] 22

ACCEPTED

Copyright © 2009, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved.Appl. Environ. Microbiol. doi:10.1128/AEM.01686-08 AEM Accepts, published online ahead of print on 5 February 2009

Copyright © 2009, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved.Appl. Environ. Microbiol. doi:10.1128/AEM.01686-08 AEM Accepts, published online ahead of print on 5 February 2009

2

Since 1998, blooms of Alexandrium catenella associated with paralytic shellfish 23

poisoning have been repeatedly reported in Thau lagoon (French Mediterranean coastal area). 24

From data obtained on ribosomal DNA markers, it has been suggested that these strains could be 25

closely related to the Japanese Temperate Asian ribotype of the Temperate Asian clade. In order 26

to get more insights into their origin, we carried out a genetic analysis of 61 Mediterranean and 27

23 Japanese strains on both ribosomal and microsatellite markers. Whereas the phylogeny on 28

ribosomal markers tended to confirm the previous findings, the analysis of microsatellite 29

sequences revealed an unexpected partition between the French and Japanese populations. This 30

analysis also highlighted a large intra-specific diversity that was not detected with the classical 31

rDNA markers. The Japanese strains are divided into two differentiated A. catenella lineages: the 32

Sea of Japan lineage and the east coast lineage which includes populations from the Inland Sea 33

and the Pacific Ocean. A. catenella strains isolated in Thau lagoon would belong to another 34

lineage. These findings indicate that microsatellite markers are probably better suited to 35

investigate the population genetics of this worldwide distributed species. Finally, the application 36

of the population genetics concepts available for macroorganisms could support new paradigms 37

for speciation and migration in phytoplankton assemblages. 38

39

40

41

ACCEPTED

3

Toxic phytoplankton blooms, also known as harmful algae blooms (HABs), have long 42

been considered as sporadic phenomena. These world-wide distributed events occur particularly 43

in coastal and confined waters when HAB species bloom or rise up from their “hidden status” (3). 44

Their increasing occurrence is a significant and expanding threat to health, the fisheries and 45

shellfish industries (22). Many toxic phytoplankton species involved in these toxic events belong 46

to the class Dinophyceae (50). In French coastal waters, the genus Alexandrium Halim is 47

represented by A. minutum in the North Brittany coast and A. catenella (Whedon and Kofoid) 48

Balech in the NW Mediterranean coast; both produce neurotoxins responsible for paralytic 49

shellfish poisoning (PSP) (19). In the French Mediterranean coast, A. catenella was first observed 50

in 1998 in the Thau lagoon, and was responsible for a first toxic event. Since then, the species has 51

repeatedly produced toxic blooms (years 2001, 2003, 2004, 2005, 2007) particularly in autumn 52

and/or spring periods during which the shellfish trade, a significant economic activity in this area, 53

was forbidden (26, 30). Moreover, it is regularly detected along the neighbouring Catalan coast 54

(52). 55

Previous studies based on ribosomal RNA gene sequencing have highlighted the existence 56

of a complex of morphologically similar species named the “tamarense complex”. Within this 57

complex including A. tamarense, A. catenella and A. fundyense, the ribotypes were grouped into 58

phylogeographic clusters closely related to geographic zones defined by the main oceanic coastal 59

areas (19, 45, 47). These studies accredited the obviousness of A. catenella spread by human 60

activities such as shipping or shellfish aquaculture (7). Lilly et al (29) assumed that A. catenella 61

could have recently been introduced into Thau lagoon through ballast water from Japan. Indeed, 62

PSP events of A. catenella have frequently caused damage in Japanese coastal waters over the 63

last 25 years (55). This hypothesis was recently supported after a clear link was established 64

ACCEPTED

4

between NW Mediterranean and Japanese A. catenella strains based on phylogenetic analyses of 65

the 5.8S rDNA gene and ITS regions (40). 66

Microsatellite markers, commonly used to analyse population genetics of various 67

terrestrial and aquatic macroorganisms (25), have recently been developed on phytoplankton 68

species: diatoms Ditylum brightwellii (41-43) and Pseudo-nitzschia pungens (12), and 69

coccolithophore Emiliania huxleyi (24). They are recognized as valuable markers for better 70

understanding the genetic structure of phytoplankton populations over both temporal and spatial 71

scales. Among the Dinophyceae, many microsatellite markers have recently been developed in 72

order to gain more insight on the development and dynamics of blooms of the genus 73

Alexandrium: A. tamarense, North American and Temperate Asian ribotypes (2, 32), A. minutum 74

(34) and A. catenella (35). In a first study of the A. tamarense population, microsatellite markers 75

evidenced some correlation between genetic and geographic distances, and several segregated 76

populations, in coastal waters around Japan (33). 77

In the present survey, we investigated the inter- and intra- specific diversity of French 78

Mediterranean and Japanese strains using a comparative genetic analysis of 12 microsatellite 79

markers developed in A. catenella (35) with population genetics tools. Their phylogenetic 80

relationships were established using the ribosomal marker spanning the region ITS1 + 5.8S 81

rDNA + ITS2 + D1/D3 28S rDNA. This survey tries to measure the genetic diversity of French 82

and Japanese A. catenella strains, in order to explore the Japanese origin of the French 83

Mediterranean strains; it is also a first opportunity to check the ploidy level of the different 84

phases of this dinoflagellate’s life cycle. 85

86

87

88

ACCEPTED

5

MATERIALS AND METHODS 89

Sampling site and isolates collection. Thau lagoon is the deepest marine lagoon located 90

on the French Mediterranean coast (43° 24’ N-3° 36’ E) covering 75 km2 of which 1/5 is 91

occupied by shellfish farming structures (Fig.1). The lagoon is connected to the Mediterranean 92

Sea by three channels, one of them opening into the Sète international harbour. Water and 93

sediment samples were collected in the Angle Creek where the A. catenella blooms are usually 94

initiated. A. catenella cultures were developed from single vegetative cells (ACTEM) or single 95

resting cysts (ACTKM) (Table 1). Resting cysts were isolated from the first 3 cm of sediment 96

cores in April 2004. Their isolation and germination were performed as described in (17). 97

Planktonic vegetative cells were isolated from the bloom occurring in November 2004. A. 98

catenella cultures were grown in Enriched Sea Natural Water (23) at 20°C with a 12:12 h 99

light:dark photocycle at a photon flux density of ~100 µmol m-2

s-1

. After centrifugation at 1800 100

g for 3 minutes, cultures were frozen at -20°C. Japanese strains were isolated in four locations in 101

the Akasaki seaport in the Sea of Japan (35°51’ 11’’ N-133°65’69’’ E), Uchino-Umi 102

(34°22’03’’N-134° 60’ 57’’ E) and Kita Nada (34° 21’ 44’’ N-134° 44’ 86’’ E) in the Inland Sea 103

and Ago bay (34° 28’ 36’’ N-136°80’ 04’’E) on the Pacific coast (Fig.1 and Table 1). A 104

collection of 61 monoclonal cultures from Thau lagoon and 23 from Japanese waters and 105

sediment were analyzed in this study (Table 1). 106

Molecular methods. Total genomic DNA was extracted from monoclonal cultures 107

according to the classical phenol:chloroform method (44). Ribosomal DNA sequencing was 108

carried out on 23 selected A. catenella strains (11 from Thau Lagoon, 12 from Japan). A 109

ribosomal DNA fragment, spanning the region ITS1 + 5.8S rDNA + ITS2 + D1/D3 28S rDNA 110

(~1500 bp), was PCR-amplified with our novel primer 18S-ITS1-Ac-F 5’-111

CTTAGAGGAAGGAGAAGTCG-3’ (forward) specifically designed to cover the full length 112

ACCEPTED

6

sequence of the ITS1 region and the reverse primer 28S-D3B-R 5’-113

TCGGAGGGAACCAGCTACTA-3’ (reverse) (38). PCR amplification was performed using the 114

proof-reading DNA polymerase PrimeSTAR HS (Takara Bio Inc., Otsu, Shiga, Japan), with the 115

following conditions: 95°C for 2 min, 35 cycles (95°C, 30 s; 55°C, 15 s; 72°C, 2 min) and a final 116

extension at 72°C for 5 minutes. The PCR products checked on agarose electrophoresis gel and 117

revealing a single band were purified using the MinElute PCR Purification Kit (Qiagen, Hilden, 118

Germany). The purified PCR products were used as templates for direct sequencing using the two 119

PCR primers and an additional internal primer. Sequencing was performed by Macrogen (Seoul, 120

South Korea) using an ABI 3730XL Genetic Analyzer (Applied Biosystems). The sequence data 121

were assembled using the software package Vector NTI (Invitrogen, Carlsbad, CA, USA). 122

Nucleotide sequence alignments, with data retrieved from Genbank, were performed using 123

ClustalX (51). The maximum likelihood phylogenetic analysis was carried out using PHYML 124

v2.4.4 (21) as follows: alignment spanning 515 A. catenella positions (ITS1 + 5.8S rDNA + 125

ITS2), 1000 bootstraps, GTR substitution model and an “invariant + Γ” among-site substitution 126

rate distribution model approximated by 8 category rates with all the parameters estimated by the 127

program. The accession numbers of the deposited sequences used in this analysis are provided in 128

the phylogenetic tree. 129

The microsatellite analysis used the twelve loci described in (35). The PCR reactions 130

were carried out in a total volume of 10 µl, containing 1 µl of 10X buffer, 2.5 mM MgCl2 131

(Promega), 0.2 mM of each dNTP (Invitrogen), 0.5 µM of each primer labelled either with CY5 132

or fluorescein (MWG-Biotech AG), 0.3 U of Taq polymerase (Sigma), 0.1 mg/ml of BSA 133

(Roche), 2 µl of DNA template (at about 5 ng/µl) and completed up to 10 µl with DNA-free water 134

(Sigma). Thermal cycling conditions consisted of an initial denaturation at 94°C for 10 min, 135

followed by 38 cycles including denaturation at 94°C for 30 s, annealing at 60°C for 30 s, 136

ACCEPTED

7

extension at 72°C for 1 min, and a final extension at 72°C for 5 min. Two µl of PCR mixture 137

from each strain were electrophoresized onto an 8% denaturing polyacrylamide gel (Bio-Rad), 138

along with a fluorescently labelled DNA ladder 100–600 bp (Promega). Labelled DNA fragments 139

were visualized with a FMBIO II fluorescent imaging system (Hitachi), and the allele sizes were 140

estimated using the FMBIOAnalysis 8.0 image analyser program (Hitachi). 141

Statistical analyses of microsatellite markers. The statistical analyses were performed 142

using the programme GENETIX (4). A series of multidimensional analyses (Factorial 143

Correspondence Analysis or FCA, (5) provided the overall genetic structuring of the samples. 144

The objective of this analysis was to represent each individual according to its complete 145

genotype. For this, the genotypic data were first coded according to the presence of each allele 146

with value 0 (allele absent), 1 (heterozygous for the allele), 2 (homozygous for the allele) (49). 147

The computation then aimed at finding composite axes which were a combination of the 148

variables and optimised the differences between the individuals analysed. The relationships 149

among individuals can be visualized on two or three axes. In order to compare genotypes of 150

different ploidies in the same analysis, we had to weight the haploid genotypes as 2, meaning that 151

a haploid genotype was given the same weight as the two alleles of a diploid heterozygote 152

genotype. The inertia values (i.e. the proportion of the total information contained by an axis, 153

given as a percent) along each axis, were shown to be equivalent to linear combinations of the 154

monolocus Fixation Index (Fst) values (20). Standard parameters of population genetics were 155

calculated for each group of strains: allele frequencies, P(0.95) and P(0.99) (proportion of 156

polymorphic loci) and A (mean number of alleles per locus) parameters. The observed 157

heterozygosity (Hobs) and unbiased expected heterozygosity (Hnb) (unbiased according to the 158

sample sizes) (36) were estimated for the diploid samples (cyst cultures). Regarding the F 159

statistics of Wright, the Fis value (f parameter of Weir and Cockerham (54)) was used to test the 160

ACCEPTED

8

departures from Hardy-Weinberg equilibrium of the cyst collection (expected to be diploid), and 161

the Fst value (f of Weir and Cockerham (54)) was used to measure the differentiation between 162

samples using both cells and cysts. For these tests, the significance of departure from zero (null 163

hypothesis) was estimated by (5000) permutation tests, using GENETIX options. The ploidy 164

level of each cycle phase has seldom been tested and microsatellite data can be used to estimate 165

this level according to the heterozygote genotypes observed. However, apparent heterozygotes 166

can also result from false monoclonal cultures. Indeed, in spite of thorough verifications during 167

the cell isolation procedure, two swimming cells could sometimes be isolated, thus giving rise to 168

a false monoclonal culture. Here, we compared the proportion of heterozygote genotypes in 169

cultures of expected haploid cells and in cultures of expected diploid resting cysts. 170

Nucleotide sequence accession numbers. Sequence data have been deposited with 171

GenBank accession numbers FM211455 to FM211477. 172

ACCEPTED

9

RESULTS 173

Ribotypes analysis. Though direct sequencing of rDNA amplicons revealed some 174

polymorphisms, as reported by Lilly et al (28), the ribosomal DNA sequences spanning the (ITS1 175

+ 5.8S rDNA + ITS2 + 28S rDNA D1/D3) region (~1500 bp) obtained from the 23 selected A. 176

catenella strains from Thau Lagoon and Japan were identical. Over the (ITS1 + 5.8S rDNA + 177

ITS2) region (515 bp), the consensus sequences were identical, or nearly, to most other A. cf. 178

catenella sequences deposited since 2000. This is featured in the phylogenetic tree of the 179

Alexandrium genus showing our new data among Genbank deposited sequences (Fig. 2). 180

Microsatellite analyses. The average proportion of missing genotypes per locus was 181

12.2% and varied from 4.5% (loci Acat02 and 49) to 24.7% (locus Acat10). Missing genotypes 182

represented 11.6% and 13.8% in the French and Japanese strains, respectively. This lack of 183

amplification is not likely to originate from cross-priming (the primers were designed from a 184

Japanese strain DNA) but rather from sample preservation problems and/or null alleles. 185

Microsatellite genotypes obtained in cultures originating from planktonic cells and germinated 186

resting cyst cultures exhibited few heterozygote patterns. Though Chi2 test was not significant, 187

the rate of heterozygous genotypes in the French Mediterranean isolates, was higher in cyst-188

derived cultures (5/114 genotypes = 4.4%) than in cultures obtained from planktonic cells (8/477 189

= 1.7%). The number of investigated Japanese strains was not large enough to make the same 190

comparison. In the following calculations, heterozygous cysts were considered as true 191

heterozygotes, whereas heterozygous planktonic/vegetative cells were assumed to be artefactual 192

(e.g. possibly due to polyclonal cultures mistakenly produced during isolation or to an occasional 193

duplication and mutation of a microsatellite target region). In the French strains, heterozygotes were 194

detected in loci Acat14, 49 and in loci Acat14, 16, 20, 34 and 50, in cyst- and vegetative cell-195

derived cultures, respectively. In the Japanese samples, only one heterozygous genotype was 196

ACCEPTED

10

observed at locus Acat14 in a vegetative cell-derived culture. For cyst samples only, expected to 197

contain a diploid genetic reservoir, the Hardy-Weinberg Equilibrium (HWE, genotype 198

frequencies expected in panmixia) was tested by estimating the significance of Fis departure from 199

zero. The Fis values were between 0.87 and 1.00, and the permutation tests (5000 permutations) 200

showed that the heterozygote deficit was highly significant (p<0.001). The genetic diversity of 201

the A. catenella samples was investigated through several methods. The classical descriptive 202

parameters of population genetics showed that the genetic polymorphism was higher in Japan 203

than in Thau (Table 2). The Akasaki strains were considerably different from other Japanese and 204

Mediterranean strains, showing ten private alleles in seven loci (Acat02, Acat06, Acat14, Acat22, 205

Acat37, Acat 44 and Acat50). Parameter A (mean number of alleles per locus) which is 206

dependent on the sample size, reached 2.7 (42 strains) and 2.1 (10), respectively, for vegetative 207

cells and cysts sampled in 2004 in Thau, whereas the main Japanese sample (Akasaki excluded 208

because of its too small sample size) reached 3.7 with only 16 strains. The value obtained from 209

the strains isolated in Akasaki seaport could not be considered because of its too small sample 210

size (6 strains). This group of strains was thereafter separated from the other Japanese ones 211

because of its high differentiation (see FCA paragraph, below). In order to avoid strong 212

distortions, only nearly complete genotypes were taken into account in the FCA analyses, i.e. 213

with 3 (i.e. 25%) missing genotypes (either due to failed PCR or null alleles) at the most among 214

the 12 loci. As a result, the numbers of strains used in the analyses were as follows: 20 Japanese 215

strains, and 42 Thau 2004 vegetative cells strains (but due to the transformation of heterozygotes 216

by the creation of virtual haploid strains, this number rose up to 53), 10 Thau 2004 cyst strains 217

(heterozygote genotypes were retained since they were considered as true), plus ATTL01, ACT1 218

ACT2 and ACT3 (Table 1). The FCA of the whole dataset showed that the Japanese strains from 219

Akasaki were genetically very different from all the other strains (diagram not shown); the entire 220

ACCEPTED

11

analyse was thus devoted to the Akasaki/non-Akasaki differences. After excluding Akasaki, the 221

FCA showed a clear separation between the French Mediterranean and Japanese strains in the 222

first FCA axis. The dispersion was much larger in the Japanese isolates, indicating a higher 223

polymorphism (Fig. 3). However, a cyst strain from Thau seemed to be closer to the Japanese 224

strains, and one strain from Ago Bay fell among the Thau population (a second one was 225

discriminated along the third axis - not shown). When analyzed alone, the 61 Thau strains 226

showed no heterogeneity in the population structure (diagram not shown). 227

Fst estimations were calculated and the statistical evaluation of the significance of the 228

departure from zero tested by 5000 permutations (Table 3). These values indicated that (i) there 229

was no differentiation between vegetative cells and cysts from Thau in 2004 (θ=0.04, ns), (ii) the 230

differentiation was significant to highly significant between France and the Pacific coast of Japan 231

(0.15<θ<0.29), (iii) the differentiation was higher between France and the Sea of Japan 232

(0.55<θ<0.58), (iv) the differentiation between the two Japanese coasts was significant only 233

between Akasaki and Ago Bay (θ=0.48). The fact that this significant value was lower that non 234

significant values of θ between Akasaki and the other Pacific stations (0.52< θ <0.54) indicated 235

that this distortion was probably due to very small sample sizes. Finally, it was clear that there 236

was no differentiation among Pacific stations in Japan (-0.02< θ<0.05, ns). 237

238

DISCUSSION 239

Phylogenetic relationships among A. catenella ribotypes. Until now, the differentiation 240

among Alexandrium lineages had mainly been investigated using ribosomal sequences (19, 39, 241

40, 45-48). These analyses showed that the Alexandrium lineages are notably related to their 242

geographic origin defined as the North American, Western European, Mediterranean, Temperate 243

Asian, Tropical Asian and Tasmanian ribotypes. Since 1998, when a toxic event of A. catenella 244

ACCEPTED

12

was observed for the first time in Thau lagoon, many studies have been carried out in order to 245

understand its ecophysiology and growth strategies (8-10, 26). Among the causation theories 246

advanced by (22) concerning HAB expansions, the appearance of A. catenella in Thau lagoon 247

through dispersal of resting cysts in ballast water (29) and in transplanted shellfish stocks from 248

aquaculture cultivation (26) has been considered as a putative source of this new and recurrent 249

outbreak. Lilly et al (29) first described the strains isolated in 1998 as belonging to the Japanese 250

ribotype of the Temperate Asian clade, and recently renamed them as Group IV in the A. 251

tamarense-complex phylogeny (28). According to the phylogenetic analysis, isolates from 252

November 2004 are similar to those isolated in 1998, 2002 and 2003. The phylogenetic tree (Fig. 253

3), gathering most of the published ITS1-5.8S rDNA-ITS2 sequences from the A. 254

tamarense/catenella/fundyense species complex including the 23 new French and Japanese A. 255

catenella strains, pointed out that the ribosomal operon could not elucidate the intraspecific 256

variability within and between A. catenella populations from a common geographic origin. 257

Furthermore, the A. catenella cluster contains some A. tamarense morphotypes from China. The 258

phylogenic grouping of isolates from this complex is consistent with the ribotype clades but not 259

with the morpho-species that form the complex (39). They are in accordance with the recently 260

developed trends (28, 39, 53) which question the phylogenetic, proteomic, biological and 261

morphological species definition within the A. tamarense complex. Because ribosomal genes are 262

ubiquitous, orthologous (15) and display little polymorphism due to a limited mutation rate (27), 263

they should still be useful and valuable markers for the species level. However, they may not 264

represent the most judicious and accurate molecular markers to analyse the intraspecific diversity 265

essential to understand migration and speciation phenomena within A. catenella assemblages. 266

Genetic diversity within the Temperate Asian A. catenella clade evidenced with 267

microsatellite markers. In contrast, it appears that microsatellite markers may be appropriate 268

ACCEPTED

13

genetic markers to investigate the intraspecific diversity of the A. catenella ribotype. For the first 269

time, Japanese and Mediterranean A. catenella isolates belonging to the Japanese Temperate 270

Asian ribotype of the Temperate Asian clade, were compared with these highly polymorphic 271

genetic markers designed from Japanese A. catenella strains. Thus, the analysis distinctly showed 272

that the studied Japanese samples were divided into two differentiated “catenella” lineages: the 273

Sea of Japan lineage (including the differentiated sample from Akasaki seaport), and the east 274

coast lineage which included populations from the Inland sea (Uchino-Umi, Kita Nada) and the 275

Pacific Ocean (Ago Bay) (Fig.1). This is the first time such heterogeneity within Japanese A. 276

catenella populations is described. It also appeared that A. catenella strains isolated in Thau 277

lagoon could belong to another lineage. Several discordances (private alleles) and Fst values were 278

always significant between France and Japan (Table 3). 279

Genetic diversity within the A. catenella lineage from Thau lagoon. Microsatellite 280

markers revealed a relatively low genetic diversity in whole Mediterranean isolates from Thau 281

lagoon (ATTL01, ACT1-2-3, ACTEM, ACTKM) as compared to Japanese populations. 282

According to these markers, these Thau strains could belong to the same lineage. This could be 283

explained by several causes: i) the introduction of a small exogenous population which would 284

have found its ecological niche; ii) the specific environmental characteristics of Thau lagoon (i.e. 285

high temperature (up to 26°C) and high salinity (up to 39-40 PSU in the summer) could have 286

selected adapted ecotypes. Despite a weak recorded diversity with microsatellites, the A. 287

catenella individuals within the population would not homogeneously participate in the formation 288

of the following generation because no panmixia was apparent according to Fis values. When the 289

following generation is not the product of a random pooling in the initial gametic pool (16) pre-290

zygotic selection should be considered. Such a low polymorphism and absence of panmixia in 291

cysts could be explained by a differential reproductive success. (i) If very few cells breed resting 292

ACCEPTED

14

cysts during the bloom, a founding effect would be generated at each reproduction. (ii) The 293

original settlement of the Thau population would stem from very few individuals of unknown 294

origin. (iii) Resting cysts sampled in 2004, which can settle for several years in the sediment, 295

could not represent a population at equilibrium. Thus, more accurate genetic markers are required 296

for the description of the diversity within A. catenella population and consequently the 297

understanding of its adaptative strategy and the success of its recurrent dominancy. 298

Ploidy level and artefacts. Many dinoflagellates, including the genus Alexandrium, have 299

a complex life cycle with a predominant vegetative haploid (n) phase with asexual replication and 300

a short sexual diploid (2n) phase (6, 18, 31). Haploid cells divide mitotically in the water column, 301

and some cells differentiate into a gametic stage. Compatible gametes (n) fuse in the water 302

column to form a short-living diploid zygote called planozygote (2n) which afterward transforms 303

into a resting cyst (2n). After a dormancy period, the cyst germinates and regenerates a 304

planomeiocyte (2n) which undergoes the meiotic division to give new haploid cells (n) that 305

develop as vegetative cells (n). The distinction between haploid and diploid cells through 306

population genetics can be made by examining the genotypes of polymorphic nuclear markers. A 307

polymorphic nuclear marker will show heterozygous genotypes in a diploid individual but not in 308

a haploid one. In our study, the genotypes were obtained from a monoclonal culture of a 309

“swimming cell” or a resting cyst. Isolating a single cell or cyst is a critical step where false 310

heterozygotes can be obtained if two cells (or more) are inadvertently isolated together, or a 311

planomeiocyte (2n) or planozygote (2n) is isolated rather than a haploid vegetative cell. 312

In the present study, the rate of heterozygote genotypes was higher in cyst cultures (4.4%) 313

than in planktonic cell cultures (1.7%) where the number of vegetative cells genotypes obtained 314

was more than four times higher than that of cyst genotypes. Despite the fact that this difference 315

in heterozygote genotypes is was not significant, we could attribute the 1.7% vegetative cell 316

ACCEPTED

15

heterozygote genotypes to mishandlings in the long process of monoclonal culture and/or a 317

percentage of planomeiocytes and planozygotes isolated in situ among the “swimming cells”. We 318

also considered that within the estimated 4.4% of cyst heterozygote genotypes, false 319

heterozygotes also occurred in the same proportion as in cell samples. This could show that the 320

analysed global diploid phase presents a low heterozygosity which has been overestimated (Ho < 321

0.04). This preliminary result is encouraging, but it is obvious that new highly polymorphic 322

microsatellites are necessary. This is why new markers were designed by the same team (37). 323

When analysing the expected Hardy-Weinberg equilibrium in the three cyst samples large 324

enough to allow the test, we observed an extreme Fis disequilibrium between 0.9 and 1 325

(p<0.001), meaning that only a few expected heterozygous genotypes were detected. Panmixia 326

(free sexuality) was not observed, highlighting the particular sexual crossing (i.e. heterothallism) 327

which occurred in the Alexandrium catenella populations. The fusion of compatible gametes 328

from the same clone (selfing, homothallism) results in the lack of effective recombination, whereas 329

the fusion of gametes from different

clones (outcrossing, heterothallism) may result in 330

recombination. Therefore, A. catenella from Thau lagoon may exhibit a heterothallic behaviour 331

as described by Yoshimatsu (56) for Japanese strains. The life cycle of A. catenella must be 332

studied in more details because sexual stages still remain unclear, contrasting with A. excavatum 333

(11), A. tamutum, A. minutum (14) and A. peruvianum (13), for which they are well described. 334

This knowledge is essential for the good understanding of the origin and occurrence of the toxic 335

events due to A. catenella. 336

Is the recent Mediterranean invasion of Japanese origin? We are far from having a 337

complete description of Japanese A. catenella polymorphism. The higher polymorphism and 338

structure of the Japanese isolates confirm that east-Asia is the natural origin of distribution of A. 339

catenella. In contrast, the weak genetic diversity among the Mediterranean strains from Thau 340

ACCEPTED

16

Lagoon could be related to a recent introduction of genotypes whose exact origin remains to be 341

determined. These preliminary results suggest that the hypothesis of the Japanese origin of A. 342

catenella blooming in Thau lagoon (French Mediterranean) should be precisely explored. 343

According to the observed and potential differentiation of the species in Asian coasts, more 344

strains from both regions must be analysed before drawing a conclusion. This study urges one to 345

compare other A. catenella isolates from the Mediterranean area with the same markers in order 346

to (i) determine if these populations belong to the same lineage and (ii) understand the genetic 347

distribution of the species in the Mediterranean basin. The genetic analysis will have to use more 348

microsatellite markers (37) and more samples, not limited to Japan. This genetic study must be 349

extended to other Asiatic sites and geographic areas in the world in order to clarify the origin of 350

A. catenella established in Thau lagoon. We consider the present survey as the first step of a 351

multi-year project prospecting A. catenella all around the world in order to understand the long 352

distance exchanges related to human activity. 353

These findings should bring insights into other environmental factors affecting the niche 354

adaptations of these microorganisms in situ. Large variations in single-cell characteristics and 355

physiological capabilities (adaptation to temperature, salinity, nutrient assimilation, allelopathic 356

activity) can be expected within and between lineages, which could have ecological 357

consequences, such as HAB events. Such a genetic approach could underpin a new paradigm for 358

speciation and migration in phytoplankton assemblages. 359

360

ACKNOWLEDGEMENTS 361

We thank Pierre-Alexandre Gagnaire for his statistical help; Annie Pastoureaud (Ifremer 362

LER/LR-Sète) for her helpful collaboration, Patrick Gentien (Ifremer DYNECO-Brest) for his 363

ACCEPTED

17

continuous support and the Ifremer staff (LER/LR-Sète) for its technical assistance. We would 364

also like to thank the anonymous reviewers for their helpful comments and suggestions. 365

This research was supported by grants from the French National Programme “Ecosphère 366

Continentale et Côtière” (EC2CO-PNEC) and the Agence Nationale de la Recherche (ANR -05-367

BLAN-0219 XPressFloral), and by the program ALCAT of Ifremer. 368

REFERENCES 369

1. Adachi, M., Y. Sako, and Y. Ishida. 1996. Analysis of Alexandrium (Dinophyceae) 370

species using sequences of the 5.8S ribosomal DNA and internaltranscribed spacer 371

regions. Journal of Phycology 32:424-432. 372

2. Alpermann, T. J., U. E. John, L. K. Medlin, K. J. Edwards, P. K. Hayes, and K. M. 373

Evans. 2006. Six new microsatellite markers for the toxic marine dinoflagellate 374

Alexandrium tamarense. Molecular Ecology Notes 6:1057-1059. 375

3. Anderson, D. M. 1989. Toxic algal blooms and red tides: a global perspective, p. 11-16. 376

In T. Okaichi, D. M. Anderson, and T. Nemoto (ed.), Red Tides: Biology, Environmental 377

Science and Toxicology. Elsevier. 378

4. Belkhir, K., P. Borsa, J. Goudet, L. Chikhi, and F. Bonhomme 1998, posting date. 379

GENETIX, logiciel sous Windows TM pour la génétique des populations. Laboratoire 380

Génome et Populations. [Online.] 381

5. Benzecri, J. P. 1982. L'analyse des données. II. L'analyse des correspondances. Dunod, 382

Paris. 383

6. Blackburn, S. I., and N. Parker. 2005. Microalgal life cycles: encystment and 384

excystment, p. 399-417. In R. A. Anderson (ed.), Algal Culturing Techniques. Elsevier 385

Academic Press, Amsterdam. 386

7. Bolch, C. J. S., and M. F. de Salas. 2007. A review of the molecular evidence for ballast 387

water introduction of the toxic dinoflagellates Gymnodinium catenatum and the 388

Alexandrium "tamarensis complex" to Australasia. Harmful Algae 6:465-485. 389

8. Collos, Y., C. Gagne, M. Laabir, A. Vaquer, P. Cecchi, and P. Souchu. 2004. 390

Nitrogenous nutrition of Alexandrium catenella (Dinophyceae) in cultures and in Thau 391

lagoon, southern France. Journal of Phycology 40:96-103. 392

9. Collos, Y., M. Lespilette, A. Vaquer, M. Laabir, and A. Pastoureaud. 2006. Uptake 393

and accumulation of ammonium by Alexandrium catenella during nutrient pulses. African 394

Journal of Marine Science 28:313-318. 395

10. Collos, Y., A. Vaquer, M. Laabir, E. Abadie, T. Laugier, A. Pastoureaud, and P. 396

Souchu. 2007. Contribution of several nitrogen sources to growth of Alexandrium 397

catenella during blooms in Thau lagoon, southern France. Harmful Algae 6:781-789. 398

11. Destombe, C., and A. Cembella. 1990. Mating-type determination, gametic recognition 399

and reproductive success in Alexandrium excavatum (Gonyaulacales, Dinophyta), a toxic 400

red-tide dinoflagellate. Phycologia 29:316-325. 401

12. Evans, K. M., S. F. Kuhn, and P. K. Hayes. 2005. High levels of genetic diversity and 402

low levels of genetic differentiation in North Sea Pseudo-nitzschia pungens 403

(Bacillariophyceae) populations. Journal of Phycology 41:506-514. 404

ACCEPTED

18

13. Figueroa, R. I., I. Bravo, and E. Garces. 2008. The significance of sexual versus 405

asexual cyst formation in the life cycle of the noxious dinoflagellate Alexandrium 406

peruvianum. Harmful Algae 7:653-663. 407

14. Figueroa, R. I., E. Garces, and I. Bravo. 2007. Comparative study of the life cycles of 408

Alexandrium tamutum and Alexandrium minutum (Gonyaulacales, Dinophyceae) in 409

culture. Journal of Phycology 43:1039-1053. 410

15. Fitch, W. M. 1970. Toward defining the tree of maximum parsimony, p. 160-178. In G. 411

F. Estabrook (ed.), Eighth International conference on numerical taxonomy, Freeman: San 412

Francisco. 413

16. Genovesi-Giunti, B. 2006. Initiation, maintien et récurrences des efflorescences toxiques 414

d'Alexandrium catenella (Dinophyceae) dans une lagune méditerranéenne (Thau, France): 415

rôle du kyste dormant. Ph.D. thesis. Montpellier, France, Montpellier, France. 416

17. Genovesi-Giunti, B., M. Laabir, and A. Vaquer. 2006. The benthic resting cyst: A key 417

actor in harmful dinoflagellate blooms - A review. Vie Et Milieu-Life and Environment 418

56:327-337. 419

18. Giacobbe, M. G., and X. M. Yang. 1999. The life history of Alexandrium taylori 420

(Dinophyceae). Journal of Phycology 35:331-338. 421

19. Guillou, L., E. Nezan, V. Cueff, E. E. L. Denn, M. A. Cambon-Bonavita, P. Gentien, 422

and G. Barbier. 2002. Genetic diversity and molecular detection of three toxic 423

dinoflagellate genera (Alexandrium, Dinophysis, and Karenia) from French coasts. Protist 424

153:223-238. 425

20. Guinand, B. 1996. Use of a multivariate model using allele frequency distributions to 426

analyse patterns of genetic differentiation among populations. Biological Journal of the 427

Linnean Society 58:173-195. 428

21. Guindon, S., and O. Gascuel. 2003. A simple, fast, and accurate algorithm to estimate 429

large phylogenies by maximum likelihood. Systematic Biology 52:696-704. 430

22. Hallegraeff, G. M. 1993. A review of harmful algal blooms and their apparent global 431

increase. Phycologia 32:79-99. 432

23. Harrison, P. J., R. E. Waters, and F. J. R. Taylor. 1980. A broad spectrum artificial 433

seawater medium for coastal and open ocean phytoplankton. Journal of Phycology 16:28-434

35. 435

24. Iglesias-Rodriguez, M. D., O. M. Schofield, J. Batley, L. K. Medlin, and P. K. Hayes. 436

2006. Intraspecific genetic diversity in the marine coccolithophore Emiliania huxleyi 437

(Prymnesiophyceae): The use of microsatellite analysis in marine phytoplankton 438

population studies. Journal of Phycology 42:526-536. 439

25. Jarne, P., and P. J. L. Lagoda. 1996. Microsatellites, from molecules to populations and 440

back. Trends in Ecology & Evolution 11:424-429. 441

26. Laabir, M., Z. Amzil, P. Lassus, E. Masseret, Y. Tapilatu, R. De Vargas, and D. 442

Grzebyk. 2007. Viability, growth and toxicity of Alexandrium catenella and 443

Alexandrium minutum (Dinophyceae) following ingestion and gut passage in the oyster 444

Crassostrea gigas. Aquatic Living Resources 20:51-57. 445

27. Lecointre, G., H. Philippe, H. L. Van Le, and H. Le Guyader. 1993. Species sampling 446

has a major impact on phylogenetic inference. Molecular Phylogenetics and Evolution 447

2:205-24. 448

28. Lilly, E. L., K. M. Halanych, and D. M. Anderson. 2007. Species boundaries and 449

global biogeography of the Alexandrium tamarense complex (Dinophyceae). Journal of 450

Phycology 43:1329-1338. 451

ACCEPTED

19

29. Lilly, E. L., D. M. Kulis, P. Gentien, and D. M. Anderson. 2002. Paralytic shellfish 452

poisoning toxins in France linked to a human-introduced strain of Alexandrium catenella 453

from the western Pacific: evidence from DNA and toxin analysis. Journal of Plankton 454

Research 24:443-452. 455

30. Masselin, P., A. Z., A. E., E. Nézan, C. Le Bec, A. Carreras, C. Chiantella, and P. 456

Truquet. 2001. Presented at the Harmful Algal blooms 2000, Proceedings of the IXth 457

international Conference Harmful Algal Blooms, Hobart, Tasmania. 458

31. Montresor, M. 1995. The life history of Alexandrium pseudogonyaulax (Gonyaulacales, 459

Dinophyceae). Phycologia 34:444-448. 460

32. Nagai, S., C. Lian, M. Hamaguchi, Y. Matsuyama, S. Itakura, and T. Hogetsu. 2004. 461

Development of microsatellite markers in the toxic dinoflagellate Alexandrium tamarense 462

(Dinophyceae). Molecular Ecology Notes 4:83-85. 463

33. Nagai, S., C. Lian, S. Yamaguchi, M. Hamaguchi, Y. Matsuyama, S. Itakura, H. 464

Shimada, S. Kaga, H. Yamauchi, Y. Sonda, T. Nishikawa, C. H. Kim, and T. 465 Hogetsu. 2007. Microsatellite markers reveal population genetic structure of the toxic 466

dinoflagellate Alexandrium tamarense (Dinophyceae) in Japanese coastal waters. Journal 467

of Phycology 43:43-54. 468

34. Nagai, S., L. McCauley, N. Yasuda, D. L. Erdner, D. M. Kulis, Y. Matsuyama, S. 469

Itakura, and D. M. Anderson. 2006. Development of microsatellite markers in the toxic 470

dinoflagellate Alexandrium minutum (Dinophyceae). Molecular Ecology Notes 6:756-471

758. 472

35. Nagai, S., M. Sekino, Y. Matsuyama, and S. Itakura. 2006. Development of 473

microsatellite markers in the toxic dinoflagellate Alexandrium catenella (Dinophyceae). 474

Molecular Ecology Notes 6:120-122. 475

36. Nei, M. 1978. Estimation of average heterozygosity and genetic distance from a small 476

number of individuals. Genetics 89:583-590. 477

37. Nishitani, G., S. Nagai, E. Masseret, C. Lian, S. Yamaguchi, N. Yasuda, S. Itakura, 478

D. Grzebyk, P. Berrebi, and M. Sekino. 2007. Development of compound microsatellite 479

markers in the toxic dinoflagellate Alexandrium catenella (Dinophyceae). Plankton and 480

Benthos Research. 481

38. Nunn, G., B. Theisen, B. Christensen, and P. Arctander. 1996. Simplicity-correlated 482

size growth of the nuclear 28S ribosomal RNA D3 expansion segment in the crustacean 483

order isopoda. Journal of Molecular Evolution 42:211-223. 484

39. Penna, A., S. Fraga, M. Maso, M. G. Giacobbe, I. Bravo, E. Garces, M. Vila, E. 485

Bertozzini, F. Andreoni, A. Luglie, and C. Vernesi. 2008. Phylogenetic relationships 486

among the Mediterranean Alexandrium (Dinophyceae) species based on sequences of 487

5.8S gene and Internal Transcript Spacers of the rRNA operon. European Journal of 488

Phycology 43:163-178. 489

40. Penna, A., E. Garces, M. Vila, M. G. Giacobbe, S. Fraga, A. Luglie, I. Bravo, E. 490

Bertozzini, and C. Vernesi. 2005. Alexandrium catenella (Dinophyceae), a toxic 491

ribotype expanding in the NW Mediterranean Sea. Marine Biology 148:13-23. 492

41. Rynearson, T. A., and E. V. Armbrust. 2000. DNA fingerprinting reveals extensive 493

genetic diversity in a field population of the centric diatom Ditylum brightwellii. 494

Limnology and Oceanography 45:1329-1340. 495

42. Rynearson, T. A., and E. V. Armbrust. 2004. Genetic differentiation among 496

populations of the planktonic marine diatom Ditylum brightwellii (Bacillariophyceae). 497

Journal of Phycology 40:34-43. 498

ACCEPTED

20

43. Rynearson, T. A., and E. V. Armbrust. 2005. Maintenance of clonal diversity during a 499

spring bloom of the centric diatom Ditylum brightwellii. Molecular Ecology 14:1631-500

1640. 501

44. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory 502

manual. Cold Spring Harbor Laboratory Press, New York. 503

45. Scholin, C. A., and D. M. Anderson. 1994. Identification of group-specific and strain-504

specific genetic-markers for globally distributed Alexandrium (dinophyceae) .1. RFLP 505

analysis of SSU ribosomal-RNA genes. Journal of Phycology 30:744-754. 506

46. Scholin, C. A., and D. M. Anderson. 1996. LSU rDNA-based RFLP assays for 507

discriminating species and strains of Alexandrium (Dinophyceae). Journal of Phycology 508

32:1022-1035. 509

47. Scholin, C. A., M. Herzog, M. Sogin, and D. M. Anderson. 1994. Identification of 510

group-specific and strain-specific genetic-markers for globally distributed Alexandrium 511

(dinophyceae) .2. Sequence-analysis of a fragment of the LSU ribosomal-RNA gene. 512

Journal of Phycology 30:999-1011. 513

48. Sebastian, C. R., S. M. Etheridge, P. A. Cook, C. O'Ryan, and G. C. Pitcher. 2005. 514

Phylogenetic analysis of toxic Alexandrium (Dinophyceae) isolates from South Africa: 515

implications for the global phylogeography of the Alexandrium tamarense species 516

complex. Phycologia 44:49-60. 517

49. She, J. X., M. Autem, G. Kotoulas, N. Pasteur, and F. Bonhomme. 1987. Multivariate 518

analysis of genetic exchanges between Solea aegyptiaca and Solea senegalensis (Teleosts, 519

Soleidae). Biological Journal of the Linnean Society 32:357-371. 520

50. Sournia, A. 1995. Red tide and toxic marine phytoplankton of the world ocean: an 521

inquiry into biodiversity, p. 103-112. In P. Lassus, G. Arzul, E. Erard-Le Denn, P. 522

Gentien, and C. Marcaillou-Le Baut (ed.), Harmful Marine Algal Blooms. Lavoisier, 523

Intercept Ltd, Paris. 524

51. Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanmougin, and D. G. Higgins. 1997. 525

The CLUSTAL_X windows interface: flexible strategies for multiple sequence alignment 526

aided by quality analysis tools. Nucleic Acids Research 25:4876-4882. 527

52. Vila, M., E. Garces, M. Maso, and J. Camp. 2001. Is the distribution of the toxic 528

dinoflagellate Alexandrium catenella expanding along the NW Mediterranean coast. 529

Marine Ecology-Progress Series 222:73-83. 530

53. Wang, D. Z., L. Lin, H. F. Gu, L. L. Chan, and H. S. Hong. 2008. Comparative studies 531

on morphology, ITS sequence and protein profile of Alexandrium tamarense and A. 532

catenella isolated from the China Sea. Harmful Algae 7:106-113. 533

54. Weir, B. S., and C. C. Cockerham. 1984. Estimating F-statistics for the analysis of 534

population-structure. Evolution 38:1358-1370. 535

55. Yoshida, T., Y. Sako, and A. Uchida. 2001. Geographic differences in paralytic shellfish 536

poisoning toxin profiles among Japanese populations of Alexandrium tamarense and 537

Alexandrium catenella (Dinophyceae) Phycological Research 29:13-21. 538

56. Yoshimatsu, S. 1981. Sexual reproduction of Protogonyaulax catenella in culture I. 539

Heterothallism. Bulletin of Plankton Society of Japan 28 131-139. 540

541

542

ACCEPTED

21

TABLE CAPTIONS 543

544

TABLE 1. Designation and location from which the French and Japanese A.catenella strains 545

analysed in the study were isolated. 546

547

TABLE 2. Alleles frequencies and classical parameters calculation. P(0.95) and P(0.99) are the 548

proportion of polymorphic loci following the criterions 0.95 and 0.99 respectively; A is the mean 549

number of alleles by locus. ACTEM: French cultures obtained from planktonic cells, ACTKM: 550

French cultures obtained from cysts, J06: Japanese cultures obtained from planktonic cells or 551

cysts, Akasaki: cultures obtained from cysts in Akasaki harbour. 552

553

TABLE 3. Fst values calculated according to the Weir and Cockerham (1984) θ estimator. Using 554

GENETIX software, 5000 permutations allowed to calculate the p values which were not 555

significant (ns), significant at 0.05 level (*) or at 0.01 level (**). 556 ACCEPTED

22

557

558

TABLE 1. Designation and location of the French and Japanese A.catenella strains

analysed in this study

Strain

designation N

Sampling

Date Locality Prefecture

Isolation

source

France

ATTL01 1 1998 Thau - seawater

ACT1-2 2 2002 Thau - sediment

ACT3 1 2003 Thau - seawater

ACTEM1 to 47 47 2004 Thau - seawater

ACTKM1 to 10 10 2004 Thau - sediment

Japan

AC0202-01 to 18 6 2001 Uchino-Umi Tokushima sediment

AC0206-06 to 20 7 2002 Ago Bay Mie seawater

AC0310-01 to 14 4 2003 Kita Nada Kagawa sediment

AC0409-02 to 13 6 2004 Akasaki Seaport Tottori sediment

559

560

561

562

563

564

565

566

ACCEPTED

23

567

568

TABLE 2. Alleles frequencies and classical parameters

calculation.

Locus ACTEM ACTKM J06 Akasaki

Acat02 (N) 42 10 16 1

288 1 1 1 0

300 0 0 0 1

Acat06 (N) 41 9 14 4

226 0 0 0.0714 0

230 0.4390 0.3333 0 0

232 0.5122 0.5556 0.5714 0

234 0 0.1111 0.1429 0

236 0 0 0.1429 0

238 0.0488 0 0.0714 0

240 0 0 0 0.7500

248 0 0 0 0.2500

Acat10 (N) 39 7 13 0

264 0.0513 0 0 -

266 0.8462 1 0.3846 -

268 0.0513 0 0.0769 -

270 0.0513 0 0.3077 -

272 0 0 0.2308 -

Acat14 (N) 40 10 15 4

186 0.7625 0.3500 0.6333 0

188 0 0 0.3333 0.7500

190 0 0 0.0333 0

192 0.2375 0.6500 0 0

198 0 0 0 0.2500

569

ACCEPTED

24

570

Locus ACTEM ACTKM J06 Akasaki

Acat16 (N) 41 9 15 4

250 0.0244 0 0.0667 0

252 0 0 0.0667 0

254 0.2439 0 0.2000 0

256 0.0366 0.4444 0.4000 0

258 0.6220 0.4444 0.0667 1

260 0.0732 0.1111 0.2000 0

Acat20 (N) 38 9 13 0

277 0 0 0.0769 ------

279 0.9605 0.8889 0.7692 ------

281 0.0395 0.1111 0.1538 ------

Acat22 (N) 40 10 16 4

212 0 0 0.0625 0.7500

214 0.9000 1 0.5000 0

216 0.1000 0 0.4375 0

238 0 0 0 0.2500

Acat34 (N) 42 10 16 4

172 0.4881 0.7000 0 1

174 0 0 0.0625 0

176 0 0 0.0625 0

178 0 0 0.2500 0

180 0.4881 0.3000 0.5625 0

182 0.0238 0 0.0625 0

571

572

573 ACCEPTED

25

574

575

576

Locus ACTEM ACTKM J06 Akasaki

Acat37 (N) 42 10 16 4

219 0 0 0 0.7500

223 1 1 1 0

229 0 0 0 0.2500

Acat44 (N) 37 9 12 4

208 0 0 0.0833 0

212 0.2973 0.2222 0.3333 0

214 0.4865 0.7778 0.3333 0

216 0.1351 0 0.1667 0

218 0 0 0.0833 0

220 0.0811 0 0 0.7500

228 0 0 0 0.2500

Acat49 (N) 41 10 16 4

232 0.0244 0 0.3125 0.7500

234 0.4390 0.4500 0.5625 0.2500

236 0 0.0500 0.1250 0

238 0.5366 0.5000 0 0

Acat50 (N) 42 10 16 4

164 0 0 0.0625 0.2500

166 0 0.0500 0.1250 0

168 0.2500 0.1000 0.0625 0

170 0.5357 0.7000 0.6250 0

172 0.2143 0.1500 0.0625 0

174 0 0 0.0625 0

176 0 0 0 0.7500

(N) 42 10 16 4

P(0.95) 0.7500 0.6667 0.8333 0.7000

P(0.99) 0.8333 0.6667 0.8333 0.7000

A 2.7500 2.0833 3.7500 1.7000

P(0.95) and P(0.99) are the proportion of polymorphic loci 577

following the criterions 0.95 and 0.99, respectively; A is the 578

mean number of alleles per locus. ACTEM: French cultures 579

obtained from planktonic cells, ACTKM: French cultures 580

obtained from cysts, J06: Japanese cultures obtained from 581

planktonic cells or cysts, Akasaki: cultures obtained from 582

cysts in Akasaki harbour. 583

584

585

ACCEPTED

26

586

587

588

TABLE 3. Fst values calculated according to the Weir and Cockerham

(1984) θ estimator. Using GENETIX software, 5000 permutations

allowed to calculate the p values which were not significant (ns),

significant at 0.05 level (*) or at 0.01 level (**).

ACTKM Uchino-Umi Ago Bay Kita Nada Akasaki

ACTEM 0.04326ns 0.26852** 0.14961** 0.26801** 0.55077**

ACTKM 0.27977** 0.17752** 0.29159* 0.58133*

Uchino-Umi -0.01687ns 0.05485ns 0.53721ns

Ago Bay 0.02373ns 0.48041ns

Kita Nada 0.52256ns

589

590

591

592

593

594 ACCEPTED

27

FIGURE CAPTIONS 595

596

597

598

599

600

601

602

603

604

605

606

607

608

609

610

611

612

613

614

FIGURE 1. Sampling locations in France and in Japan. (Top panel reprinted from Souchu et al.

[Mar. Ecol. Prog. Ser. 218:141-152, 2001] with permission of the publisher. Bottom panel

courtesy of Satoshi Nagai.)

FIGURE 2. Maximum likelihood phylogenetic tree of genus Alexandrium, with L. polyedrum as

an outgroup, using ribosomal DNA sequences (ITS1 + 5.8S rDNA + ITS2), spanning 515 A.

catenella nucleotide positions. The scale bar (bottom left corner) represent 0.1 nucleotide

substitution per site. Identical sequences have been obtained from 23 selected A. catenella strains

(12 from Japan, 11 from Thau Lagoon), which are also identical to most other sequences in this

clade. The tree displayed in the small frame (bottom right corner) represents the relative genetic

distance between the sequences belonging to the A. catenella clade; the scale bar (0.002

nucleotide substitution per site) is equal to one nucleotide difference for the analyzed sequences.

The sequence data for A. catenella strains TNX22, TNY11, Y-2, OFX102, OFY101, ko-3 and

MI7 (accession number AB006990) are from Adachi et al (1996) (1).

FIGURE 3. Diagram of the first plan of the CA (axis 1 and axis 2) including all samples

excepted those from Akasaki, considered as outliers due to their high genetic difference. The two

axes inertia reached 16.83% for a total of 59 axes.

ACCEPTED

28

613

614

615

616

617

618

619

620

FIG.1. 621

622

623

Ago Bay

UchinoUmi

Akasaki

seaport

450 km Kita Nada

ACCEPTED

29

624

FIG. 2. 625

ACCEPTED

30

626

627

628

-2

-1,5

-1

-0,5

0

0,5

1

1,5

2

-2 -1,5 -1 -0,5 0 0,5 1

Thau vegetative cells 2004

Thau cysts 2004

Thau cysts 2002

Thau vegetative cells 2003

Thau vegetative cells 1998

Uchino-Umi cysts 2002

Ago Bay vegetative cells 2002

Kita Nada cysts 2003

629

FIG. 3. 630

631

632

ACCEPTED