ATL - digital-csic Digital CSIC

advertisement
1
Two markers and one history: phylogeography of the edible common
2
sea urchin Paracentrotus lividus in the Lusitanian region
3
4
I. Calderón*, G. Giribet+, X. Turon#
5
6
*Department of Animal Biology, Faculty of Biology, University of Barcelona, 645 Diagonal
7
Ave, 08028 Barcelona, Spain. +Department of Organismic and Evolutionary Biology & Museum
8
of Comparative Zoology, Harvard University, 26 Oxford Street, Cambridge, MA 02138, USA.
9
#
10
Center for Advanced Studies of Blanes (CSIC), C. d’Accés a la Cala S. Francesc 14, 17300
Blanes (Girona), Spain.
11
12
13
14
Corresponding author:
15
I. Calderón
16
Department of Animal Biology, Faculty of Biology, University of Barcelona, 645 Diagonal
17
Ave, 08028 Barcelona, Spain.
18
Telephone: +34934021441
19
Fax: +34934035740
20
E-mail: calderon@ub.edu
21
1
22
Abstract
23
Benthic marine invertebrates with long-lived larvae are believed to have dispersal capabilities
24
that contribute to maintaining genetic uniformity among populations over large geographical
25
scales. However, both hydrological and biological factors may limit the actual dispersal of such
26
larvae. We studied the population genetic structure of the edible common sea urchin
27
Paracentrotus lividus (Lamarck, 1816) to explore its dispersal patterns in the Atlanto-
28
Mediterranean region and, more specifically, to ascertain the role of the Strait of Gibraltar in
29
shaping the genetic structure of this species. For this purpose, we analysed 158 individuals for
30
the mitochondrial 16S rRNA gene and 151 of these for the nuclear single-copy intron ANT
31
(Adenine Nucleotide Transporter) from 16 localities from the Atlantic and Mediterranean
32
basins, spanning over 4000 km. Mitochondrial 16S rRNA shows higher genetic diversity in the
33
Mediterranean than in the Atlantic and reveals a sharp break between the populations of both
34
basins, probably as a consequence of the barrier imposed by the Almería-Orán hydrological
35
front, situated east of the Strait of Gibraltar. Both markers suggest that a recent population
36
expansion has taken place in both basins, most probably following the Messinian salinity crisis.
37
2
38
Introduction
39
The last decades have witnessed an ever-increasing interest in discerning the role of historical
40
and current processes in shaping observed genetic structure at the intraspecific level (reviewed
41
in Avise 2000). In marine benthic invertebrates, gene flow between populations is mainly driven
42
by dispersal of larvae that, in some cases, can remain in the water column for weeks or even
43
months. Nevertheless, recent studies in several benthic invertebrates have shown little
44
correlation between larval lifespan and dispersal ability (e.g., Benzie 2000; Hellberg et al. 2002;
45
Palumbi 2004). Indeed, barriers to gene flow are not always conspicuous, especially in marine
46
habitats where geographical barriers, currents, temporal and spatial spawning patterns (Hart and
47
Scheibling 1998), physical and behavioural properties of larvae (Thomas 1994), juvenile
48
mortality (Gosselin and Qian 1996, 1997; Hunt and Scheibling 1997) and many other abiotic or
49
biotic factors can ultimately determine population structure over space and time (Moberg and
50
Burton 2000; Sponaugle et al. 2002).
51
Current and historical barriers to gene flow may leave a strong footprint on population
52
structure. A good model to study the role of such barriers on marine organisms is the Strait of
53
Gibraltar, which constitutes the limit between two marine biogeographical regions, the north-
54
eastern Atlantic Ocean and the Mediterranean Sea. Historically, the exchange of waters between
55
both basins was interrupted during the Messinian salinity crisis (Maldonado 1985; Pérès 1989),
56
which constituted one of the most dramatic events during the Cenozoic era (Duggen et al. 2003).
57
The opening of a new connection through the Strait of Gibraltar re-established water exchanges
58
between basins, leading to the recolonisation of the Mediterranean by organisms from the
59
Atlantic. Besides, fluctuations in sea level during the Quaternary also produced sporadic
60
separations between both basins (Nilsson 1982; Waelbroeck et al. 2002). Population
61
differentiation across the Atlantic-Mediterranean divide has been described in a number of
62
marine species (e.g., Borsa et al. 1997; Bargelloni et al. 2003; Baus et al. 2005). The
63
concordance in intraspecific patterns in species with diverse life history traits points to the
64
importance of the historical processes that occurred in this area. Nowadays, the so-called
65
Almería-Orán hydrological front, situated east of the Strait (Tintore et al. 1998) still hinders
66
migration between basins for numerous marine species (Patarnello et al. 2007 and references
67
therein).
68
Sea urchins (Echinodermata, Echinoidea) are a diverse group of marine deuterostomes
69
(Smith et al. 2006) that play an important role in structuring benthic communities (e.g., Palacín
70
et al. 1998; Sala et al. 1998; Sivertsen 2006). The common edible sea urchin Paracentrotus
71
lividus (Lamarck, 1816) is a commercially exploited species found in the north-eastern Atlantic
72
and throughout the Mediterranean (Boudouresque and Verlaque 2001). Larval lifespan has been
3
73
estimated between 20 and 40 days (Fenaux et al. 1985; Pedrotti 1993) and thus, individuals of
74
this species are potentially able to disperse over long distances. Nonetheless, high spatial,
75
bathymetric and temporal variability in settlement suggests that biotic or abiotic factors may
76
affect dispersal (Hereu et al. 2004; Tomas et al. 2004).
77
Notwithstanding the vast amount of literature on the ecology and biology of
78
Paracentrotus lividus (e.g., Savy 1987; Turon et al. 1995; López et al. 1998; Tomas et al. 2006),
79
little is still known about its genetic structure. Iuri et al. (2007), in a study based on two
80
mitochondrial and one nuclear markers, stated that P. lividus presents no genetic differentiation
81
within the Gulf of Naples. At a larger scale, Duran et al. (2004) observed panmixia within
82
Atlantic and Mediterranean basins using cytochrome c oxidase subunit I (hereafter COI) DNA
83
sequences, but detected also a slight but significant pattern of genetic differentiation between the
84
two basins.
85
The aim of our study was to obtain a more detailed picture of the population genetic
86
structure of Paracentrotus lividus throughout the Atlanto-Mediterranean region and, especially,
87
to ascertain the role of the Strait of Gibraltar in shaping the genetic structure of this species. In
88
order to achieve such a goal, we sampled the common sea urchin in16 locations from the
89
Atlantic and Mediterranean basins and analysed two molecular markers of different
90
characteristics: a mitochondrial ribosomal gene and a nuclear intron.
91
92
Material and methods
93
Sixteen locations were sampled for Paracentrotus lividus by scuba along the Atlanto-
94
Mediterranean arch (Figure 1). Distances between sampling sites ranged from 20 to around 4400
95
km. The gonads were dissected from live specimens, fixed in 96% ethanol and stored at –80ºC
96
until processing.
97
Genomic DNA was extracted using the REALPURE extraction kit (Durviz, Spain) and
98
two molecular markers were analysed. A fragment of the mitochondrial 16S rRNA gene was
99
amplified for 158 individuals with universal primers 16Sa and 16Sb (Kessing et al. 1989). The
100
single-copy intron of the nuclear Adenine Nucleotide Transporter (ANT) gene was amplified for
101
151 individuals by EPIC-PCR, using degenerate universal primers designed by Jarman et al.
102
(2002). In both cases, amplifications were performed in a final volume of 25 μL using 2.5 mM
103
of MgCl2, 1 mM of dNTPs, 0.5 μM of each primer and 1 U of Taq polymerase. For the
104
mitochondrial 16S rRNA, 1 μL of DMSO was added per sample. PCR amplicons were vacuum-
105
cleaned and labelled using BigDye® Terminator v.3.1 (Applied Biosystems, Branchburg, New
106
Jersey, USA). Sequences were obtained on an ABI 3730 and 3100 Genetic Analyzer (Applied
107
Biosystems) for 16S and ANT, respectively.
4
108
In order to reconstruct the allelic phase from the ANT genotypic data we used the
109
program PHASE v2.1 (Stephens et al. 2001; Stephens and Scheet 2005). To confirm the
110
existence of only two alleles per individual and to check the results provided by PHASE, PCR
111
products of six individuals from different populations were cloned with the pGEM-Easy Vector
112
cloning kit (Promega, Wisconsin, USA) following manufacturer’s instructions. Four to six
113
colonies per individual were sequenced.
114
115
Population genetics analyses
116
Haplotype and nucleotide diversity values were calculated with DnaSP v.4.10.3 (Rozas et al.
117
2003). Genetix v 4.05.2 (Belkhir et al. 2004) was used to calculate haplotype frequencies and
118
inbreeding coefficients from the data obtained for ANT with PHASE. Pairwise genetic distances
119
(FST) for both markers were calculated with Arlequin ver. 3.1 (Excoffier et al. 2005) and their
120
significance was assessed by performing 10,000 permutations. A Multidimensional Scaling
121
(MDS) was performed with Systat 11 (SPSS) to graphically visualise these results.
122
SAMOVA 1.0 (Dupanloup et al. 2002) was used to define groups of populations that are
123
geographically homogeneous and with the highest differentiation among each other. Analyses
124
were performed for K=2 groups with 10,000 simulated annealing procedures. AMOVA, as
125
implemented in Arlequin, was performed to further examine hierarchical population structure.
126
Finally, the correlation of genetic and geographical distances was tested with the Mantel test
127
procedure available in Arlequin.
128
129
Haplotype network
130
For 16S rRNA, the complete data set was used to build a median-joining network using
131
Network v4.2.0.1 (Bandelt et al. 1999). For ANT, however, due to the high number of
132
haplotypes found, we limited the network to populations surrounding the Strait of Gibraltar. The
133
loops observed in the networks were solved using criteria derived from coalescent theory
134
(Templeton et al. 1987; Templeton and Sing 1993).
135
136
Demographic analyses
137
Neutrality tests and mismatch distribution analyses can provide hints to infer population
138
demographic events. Tajima’s D (Tajima 1989a), Fu’s FS (Fu 1997) and R2 (Ramos-Onsins and
139
Rozas 2002) were calculated with DnaSP. Mismatch distributions (Rogers and Harpending
140
1992; Harpending 1994), as well as goodness-of-fit tests for demographic and spatial
141
expansions, were calculated with Arlequin.
142
5
143
Results
144
Diversity and population structure
145
For the mitochondrial 16S rRNA, a fragment of 582 bp was sequenced from 158 individuals. A
146
multi-T region was observed in the amplified fragment, containing from 6 to 9 Ts. Except for
147
these, all other changes observed were substitutions. Thirty-one polymorphic sites and 38
148
haplotypes were observed. Of these, 16 haplotypes were present in Atlantic samples (9 of these
149
were exclusive to this basin) whereas Mediterranean samples comprised 29 (22 haplotypes
150
exclusive to this basin; Table 1 in Appendix I). Haplotype diversity is thus higher in the
151
Mediterranean than in the Atlantic basin, and so is nucleotide diversity (Table 1).
152
For the ANT intron, 323 bp were sequenced for 151 out of the 158 individuals sequenced
153
for 16S rRNA. Fifty-seven variable sites were observed. Cloning of six randomly chosen
154
individuals proved that ANT was a single-copy marker and confirmed in every case the
155
haplotype assignment provided by PHASE. The allelic reconstruction estimated 142 haplotypes
156
(Tables 2 and 3 in Appendix I), 86 in the Atlantic (60 haplotypes exclusive to this basin) and 82
157
in the Mediterranean (56 of which were exclusive to this basin). Haplotype and nucleotide
158
diversity did not differ between basins (Table 1). Twenty-nine out of 151 individuals appeared
159
to be homozygotes, the number of homozygote individuals being evenly distributed in both
160
basins. Eight populations showed a significant departure from Hardy-Weinberg equilibrium
161
(Table 1).
162
FST for 16S rRNA showed higher levels of population differentiation between basins
163
(mean FST = 0.109) than within the Atlantic and Mediterranean basins (means of 0.021 and –
164
0.004 respectively; Table 4 in Appendix I). As a consequence, a sharp separation between
165
Atlantic and Mediterranean populations was observed in the MDS representation (Figure 2a).
166
On the contrary, differences between basins as revealed by FST measures were not as clear for
167
the ANT data, mostly due to the high variability observed in the sampled populations (Table 4 in
168
Appendix I). Likewise, MDS did not show a clear separation between basins (Figure 2b).
169
For 16S rRNA, SAMOVA for K=2 clustered Atlantic populations, plus Ceuta and
170
Tarifa, in one group, and the remaining Mediterranean populations in another group. Ceuta and
171
Tarifa, although located in the Mediterranean Sea, are placed west of the Almería-Orán
172
hydrological front. Results from AMOVA showed a significant variance component associated
173
with the differentiation between basins (11.20%, considering Ceuta and Tarifa as Atlantic
174
populations). The variance between populations within basins was low and not significant
175
(0.08%), but increased to 6.39% and became significant when no groups were specified (Table
176
2a). The Mantel test showed a significant correlation coefficient between genetic and
177
geographical distances (r=0.395, p=0.009) for the whole sample. However, neither the
6
178
correlation coefficients estimated within basins (r=0.05 and r=0.225 for the Atlantic and the
179
Mediterranean, respectively) nor the coefficient for between-basins pairs alone (r=0.184) were
180
significant, indicating that the overall significant results may, in fact, be an artefact. Indeed, this
181
outcome may stem from the fact that the global analysis lumped together comparisons within
182
basins, which corresponded to populations poorly differentiated and geographically close, with
183
between-basin comparisons, which relied on populations that tended to be more divergent and
184
widely separated. This suggests that there is no isolation by distance between our basins, but
185
only a sharp genetic break at the Almería-Orán front.
186
As for ANT, SAMOVA showed a clear differentiation between Nao and all other populations
187
when K=2 groups. When K>2, populations were separated one by one from the main group and
188
the sharp distinction between the Atlantic and Mediterranean basins observed for 16S rRNA was
189
not observed. For the sake of comparison, we computed AMOVA (and subsequent analyses)
190
with the same Atlantic and Mediterranean groups as for 16S rRNA. A much higher variability
191
within populations was observed, accounting for more than 97% of the overall variation (Table
192
2b). Variation among populations within basins was in general small but significant (between 2
193
and 3%) whereas variation between groups (basins) was not significant. As expected, the Mantel
194
test provided a smaller, non-significant correlation coefficient (r=0.137, p>0.05).
195
Haplotype network
196
The network obtained for the 16S rRNA data suggests that haplotype 1 is the ancestral
197
haplotype due to its high frequency, its wide geographical distribution and its central position in
198
the network. All haplotypes are separated by a few mutational steps (Figure 3). For ANT, the
199
haplotype network was only built for the 4 populations around the Strait of Gibraltar (Cádiz,
200
Ceuta, Tarifa and Gata), comprising 78 individuals representing 51 haplotypes. The single
201
network obtained presented a high amount of loops that were almost always unambiguously
202
resolved (Figure S1).
203
204
Demographic analyses
205
Neutrality tests for the 16S rRNA detected a population expansion for the whole sample set,
206
with significant values for all three tests. However, only the Fs and R2 statistics detected
207
significant expansions within each basin (Table 3). In the case of ANT, neutrality tests provided
208
the same results observed for 16S rRNA, although R2 did not detect a significant expansion for
209
the Atlantic basin (Table 3).
210
The mismatch distributions for both 16S rRNA and ANT presented a unimodal
211
distribution, characteristic of a sudden expansion model (Rogers and Harpending 1992) for each
212
basin separately as well as for the whole area (Figure 4 and Table 4). The test for spatial
7
213
expansion for the 16S rRNA in Mediterranean populations failed to converge with the algorithm
214
implemented in Arlequin ver. 3.1.
215
According to Rogers and Harpending (1992), the wave’s crest is determined at τ=2ut,
216
where τ is the mode of mismatch distribution and t represents the approximate time of
217
expansion. In this equation, u is the mutation rate of the entire region under study (mutation rate
218
per nucleotide times the number of nucleotides of the fragment analysed). A mutation rate of
219
0.5% per nucleotide per million years (Myr) was used for the 16S rRNA, following that in other
220
sea urchins (Chenuil and Féral 2003). Assuming independent demographic events within each
221
basin, the expansion in the Atlantic would have taken place around 270,000 years ago, whereas
222
in the Mediterranean it would have occurred 360,000 years ago. According to previous studies
223
(e.g., Lozano et al. 1995; Turon et al. 1995) Paracentrotus lividus has a generation time of 3
224
years. Therefore, expansions in each basin would have occurred 90,000 and 120,000 generations
225
ago, respectively. On the contrary, if we assume that a single expansion occurred in the whole
226
area, this would have taken place 351,000 years or 117,600 generations ago.
227
Likewise, the intron data detected population expansions both for the Atlantic and
228
Mediterranean basins, as well as for the whole distribution. Lack of data for the mutation rate of
229
nuclear introns prevented us from estimating expansions times.
230
231
Discussion
232
The study of population genetics of Paracentrotus lividus in the Atlanto-Mediterranean arch
233
using a mitochondrial ribosomal gene and a nuclear intron reveals two salient points: a lack of
234
differentiation within each basin, suggesting long range dispersal, and the role of the Strait of
235
Gibraltar, and more specifically the Almería-Orán hydrological front, in restricting gene flow
236
between Atlantic and Mediterranean populations.
237
The use of multiple molecular markers with different properties can yield considerably
238
more sensitive results than a single marker (e.g., Chow and Takeyama 2000; Buonaccorsi et al.
239
2001). MtDNA has been classically used in population genetics and phylogeographic studies
240
(reviewed in Avise 2000, but see Ballard and Whitlock 2004; Hurst and Jiggins 2005). The use
241
of introns for such studies is much more recent (e.g., Villablanca et al. 1998; Daguin et al. 2001;
242
Berrebi et al. 2005). These non-coding regions are expected to evolve at a higher rate than
243
coding regions or than ribosomal genes constituting good markers for inferring recent processes
244
at the intraspecific level (but see Kreitman 1983; Villablanca et al. 1998; Zhang and Hewitt
245
2003). One of the difficulties encountered when working with nuclear DNA is the existence of
246
allele polymorphisms. In the case of Paracentrotus lividus, the presence of homozygotes in our
247
sample (19.2% of our individuals) allowed the most likely reconstruction of the allelic phase for
8
248
all our individuals following a Bayesian approach (Clark 1990; Excoffier and Slatkin 1995;
249
Stephens et al. 2001). Although some populations showed departures from Hardy-Weinberg
250
equilibrium, the biological significance of this is questionable. Most populations showing excess
251
of homozygosis had only one homozygote individual, which was enough to be significant in the
252
HWE test, due to the high expected heterozygosity in the whole sample (Table 1; Table 2 and 3
253
in Appendix I). The level of variability of ANT was probably excessive for detecting genetic
254
structure at the studied geographical scale. Therefore, most of the variability is found within
255
populations, rendering relationships between populations difficult to infer.
256
MtDNA is a single molecule that is maternally inherited without generally overcoming
257
recombination (but see Rokas et al. 2003; Tsaousis et al. 2005). Linkage of COI and 16S rRNA
258
implies a shared evolutionary history and, thus, any difference in the patterns inferred using these
259
two genes may be due to differences in mutation rates (0.5% for 16S rRNA [Chenuil and Féral
260
2003], and 1.6-3.5% for COI [Lessios et al. 1999; McCartney et al. 2000] in sea urchins),
261
ultimately determined by their functional constraints. Iuri et al. (2007) found COI to be more
262
efficient than both mitochondrial 16S rRNA and nuclear ITS-2 to detect population structure at a
263
small geographic scale in the Gulf of Naples (maximum sampled distance <80 km). On the
264
contrary, our results for 16S rRNA provided a stronger signal of population differentiation
265
between Atlantic and Mediterranean basins than that observed for COI by Duran et al. (2004) at a
266
similar geographic scale. In our study, differentiation between basins explained 11.2% of the
267
total variance while it only accounted for 1.5% in Duran et al. (2004). The 16S rRNA data reveal
268
significant differentiation between groups (basins) but differentiation is negligible among
269
populations between groups (Table 2). The reverse pattern occurred for ANT data. This
270
apparently counter-intuitive result can be explained by the high mutation rate observed for ANT,
271
that may lead to saturation and thus lack of signal when comparing the more divergent
272
populations at both sides of the Gibraltar divide, while some signal is still appreciable between
273
populations that have diverged less (i.e., within basins).
274
Following the Messinian salinity crisis, populations of marine species in the
275
Mediterranean became for the most part extinct, since this basin was reduced to an assemblage
276
of hypersaline lakes (Hsü 1972; Briggs 1974; Blondel and Aronson 1999; Duggen et al. 2003).
277
After the opening of the Strait (5.33 Mya) the Mediterranean was refilled with Atlantic waters,
278
with the concomitant entry of biota. For many species, therefore, current Mediterranean
279
populations are at most around 5 Myr old. It is unlikely that the Mediterranean hosted the
280
common sea urchin Paracentrotus lividus or its ancestors during the Messinian episode due to
281
the strict saline conditions following the closure of the Strait of Gibraltar, especially considering
282
the poor osmoregulatory abilities of echinoderms. Taking into account that the only other
9
283
species of the same genus, P. gaimardi (de Blainville, 1825), inhabits both coasts of the
284
Southern Atlantic Ocean (Mortensen 1943), the most likely hypothesis is that P. lividus
285
originated from an Atlantic stock that colonised the Mediterranean after the salinity crisis.
286
Population growth generates an excess of (recent) mutations and therefore an excess of
287
singletons (Avise et al. 1984; Watterson 1984; Tajima 1989a,b; Slatkin and Hudson 1991;
288
Ramos-Onsins and Rozas 2002). The pattern observed in our data of few frequent haplotypes
289
and many low-frequency haplotypes with few differences, similar to that observed by Duran et
290
al. (2004) and also found in other marine invertebrates (e.g., Edmands et al. 1996; Zane et al.
291
2000; Lejeusne and Chevaldonné 2006), is compatible with this prediction. Neutrality tests and
292
mismatch distributions are also sensitive to such excess of low frequency haplotypes. Ramos-
293
Onsins and Rozas (2002) suggested that Fs and R2 are more powerful tests than Tajima’s D in
294
detecting population changes, a prediction confirmed by our results (Table 3). These significant
295
values can be due to changes in demographic parameters but also a consequence of selection
296
acting upon the studied genes. In our case, the agreement between data from unlinked markers,
297
including an intron, points to demographic changes and not selection as the most likely
298
explanation for the pattern found.
299
Mismatch distribution analyses for both markers also suggest that demographic
300
expansions took place in Paracentrotus lividus (Figure 4 and Table 4), after becoming
301
established in the Mediterranean, corroborating the data from Duran et al. (2004). Interestingly,
302
when the mismatch distribution is tested for populations one by one using 16S rRNA, only
303
Roscoff and Cádiz seem to have experienced an expansion in the Atlantic basin whereas all
304
populations in the Mediterranean show a unimodal distribution indicative of population growth
305
(data not shown). This indicates that the expansion may have been more important in the
306
Mediterranean than in the Atlantic basin. Our data suggest that expansions occurred some
307
300,000 years ago, corresponding to the Mindel-Riss interglaciary period (Kukla 2005), and
308
possibly earlier in the Mediterranean than in the Atlantic. These expansions may have been
309
determinant of the present-day genetic structure of the common sea urchin along its
310
distributional range. In particular, the lower haplotypic and nucleotidic diversity of 16S rRNA in
311
the Atlantic basin could be the result of the arrival of Mediterranean lineages into the Atlantic
312
during this expansion episode, which may have erased the genetic signal of older Atlantic
313
populations.
314
In conclusion, the origin of the populations of Paracentrotus lividus in the area studied
315
probably dates back to the end of the Messinian period and, afterwards, the interplay of glacial
316
and interglacial periods during the Quaternary and associated demographic changes would have
317
shaped the present-day distribution of the species. Restricted gene flow between the Atlantic and
10
318
the Mediterranean basins and a high connectivity within basins are nowadays the prevailing
319
processes acting upon populations of Paracentrotus lividus.
320
321
REFERENCES
322
Avise JC (2000) Phylogeography: the history and formation of species. Harvard University
323
324
325
326
327
328
329
Press. Cambridge, Ma
Avise JC, Neigel JE, Arnold J (1984) Demographic influences on mitochondrial DNA lineage
survivorship in animal populations. J Mol Evol 20, 99-105
Ballard JWO, Whitlock MC (2004) The incomplete natural history of mitochondria. Mol Ecol
13, 729-744
Bandelt HJ, Forster P, Röhl A (1999) Median-Joining networks for inferring intraspecific
phylogenies. Mol Biol Evol 16, 37-48
330
Bargelloni L, Alarcon JA, Alvarez MC, Penzo E, Magoulas A, Reis C, Patarnello T (2003)
331
Discord in the family Sparidae (Teleostei): divergent phylogeography patterns across the
332
Atlantic-Mediterranean divide. J Evol Biol 16, 1149-1158
333
334
Baus E, Darrock DJ, Bruford MW (2005) Gene-flow patterns in Atlantic and Mediterranean
populations of the Lusitanian sea star Asterina gibbosa. Mol Ecol 14, 3373-3382
335
Belkhir K, Borsa P, Chikhi L, Raufaste N, Bonhomme F (1996-2004) GENETIX 4.05, logiciel
336
sous Windows TM pour la génétique des populations. Laboratoire Génome, Populations,
337
Interactions, CNRS UMR 5171, Université de Montpellier II, Montpellier (France)
338
Benzie JAH (2000) The detection of spatial variation in widespread marine species: methods
339
and bias in the analysis of population structure in the crown of thorns starfish
340
(Echinodermata: Asteroidea). Hydrobiologia 420, 1-14
341
Berrebi P, Boissin E, Fang F, Cattaneo-Berrebi G (2005) Intron polymorphism (EPIC-PCR)
342
reveals phylogeographic structure of Zacco platypus in China: a possible target for
343
aquaculture development. Heredity 94, 589-598.
344
345
346
Blondel J, Aronson J (1999) Biology and wildlife of the Mediterranean region. Oxford
University Press, Oxford
Borsa P, Naciri M, Bahri L,Chikhi L, Garcia De Leon FJ, Kotoulas G, Bonhomme F. (1997)
347
Analyses
348
méditerranéennes (Poissons et Invertébrés). Vie Milieu 47, 295-305
349
350
351
des
donnés
génétiques
populationelles
sur
seize
espèces
Atlanto-
Boudouresque CF, Verlaque M (2001) Ecology of Paracentrotus lividus. In Edible sea urchins:
biology and ecology (eds Lawrence JM), pp. 177-512. Elseier, Tampa, FL.
Briggs JC (1974) Marine Zoogeography. McGraw-Hill, New York, NY
11
352
353
354
355
Buonaccorsi VP, McDowell JR, Graves JE (2001) Reconciling patterns of inter-ocean molecular
variance from four classes of molecular markers in blue marlin. Mol Ecol 10, 1179-1196
Clark AG (1990) Inference of haplotypes from PCR-amplified samples of diploid populations.
Mol Biol Evol 7, 111-122
356
Chenuil A, Féral JP (2003) Sequences of mitochondrial DNA suggest that Echinocardium
357
cordatum is a complex of several sympatric or hybridizing species: A pilot study. In:
358
Echinoderm Research 2001, Proceedings of the 6th European Conference on Echinoderm,
359
Banyuls-sur-Mer, France (eds Féral J-P, David B), Swets & Zeitlinger, Lisse, NL, pp. 15-
360
32
361
362
Chow S, Takeyama H (2000) Nuclear and mitochondrial DNA analyses reveal four genetically
separated breeding units of the swordfish. J Fish Biol 56, 1087-1098
363
Daguin C, Bonhomme F, Borsa P (2001) The zone of sympatry and hybridization of Mytilus
364
edulis and M. galloprovincialis, as described by intron length polymorphism at locus mac-
365
1. Heredity 86, 342-354
366
367
368
369
Duggen S, Hoernle K, van den Bogaard O, Rupke L, Phipps Morgan J (2003) Deep roots of the
Messinian salinity crisis. Nature 422, 602-606
Dupanloup I, Schneider S, Excoffier L (2002) A simulated annealing approach to define the
genetic structure of populations. Mol Ecol 11, 2571-2581
370
Duran S, Palacín C, Becerro MA, Turon X, Giribet G (2004) Genetic diversity and population
371
structure of the commercially harvested sea urchin Paracentrotus lividus (Echinodermata,
372
Echinoidea). Mol Ecol 13, 3317-3328
373
Edmands S, Moberg PE, Burton RS (1996) Allozyme and mitochondrial DNA evidence of
374
population subdivision in the purple sea urchin Strongylocentrotus purpuratus. Mar Biol
375
126, 443-450
376
377
378
379
Excoffier L, Slatkin M (1995) Maximum-likelihood estimation of molecular haplotype
frequencies in a diploid population. Mol Biol Evol 12, 921-927
Excoffier L, Laval G, Schneider S (2005) Arlequin ver. 3.0: An integrated software package for
population genetics data analysis. Evol Bioinform Online1, 47-50
380
Fenaux L, Cellario C, Etienne M (1985) Variations in the ingestion rate of algal cells with
381
morphological development of larvae of Paracentrotus lividus (Echinodermata:
382
Echinoidea). Mar Ecol Prog Ser 24, 161–165
383
384
385
386
Fu YX (1997) Statistical test of neutrality of mutations against population growth, hitchhiking
and background selection. Genetics 147, 915-925
Gosselin LA, Qian P (1996) Early postsettlement mortality of an intertidal barnacle: a critical
period for survival. Mar Ecol Prog Ser 135, 69-75
12
387
388
389
390
391
392
393
394
395
396
Gosselin LA, Qian P (1997) Juvenile mortality in benthic marine invertebrates Mar Ecol Prog
Ser 146, 265-282
Harpending HC (1994) Signature of ancient population growth in a low resolution
mitochondrial DNA mismatch distribution. Hum Biol 66, 591-600
Hart MW, Scheibling RE (1998) Heat waves, baby booms and the destruction of kelp beds by
sea urchins. Mar Biol 99, 167-176
Hellberg ME, Burton RS, Neigel JE, Palumbi SR (2002) Genetic assessment of connectivity
among marine populations. Bull Mar Sci 70, 273-290
Hereu B, Zabala M, Linares C, Sala E (2004) Temporal and spatial variability in settlement of
the sea urchin Paracentrotus lividus in the NW Mediterranean. Mar Biol 144, 1011-1018
397
Hsü KJ (1972) When the Mediterranean dried up. Sci Amer 227, 26-36
398
Hunt HL, Scheibling RE (1997) The role of early post-settlement mortality in recruitment of
399
benthic marine invertebrates: a review. Mar Ecol Prog Ser 155, 269-301
400
Hurst GDD, Jiggins FM (2005) Problems with mitochondrial DNA as a marker in population,
401
phylogeographic and phylogenetic studies: the effects of inherited symbionts. Proc R Soc
402
B 272, 1525-1534
403
Iuri V, Patti FP, Procaccini G (2007) Phylogeography of the sea urchin Paracentrotus lividus
404
(Lamarck) (Echinodermata: Echinoidea): first insights from the South Tyrrhenian Sea.
405
Hydrobiologia 580, 77-84
406
407
408
409
410
411
412
413
Jarman SN, Ward RD, Elliott NG (2002) Oligonucleotide primers for PCR amplification of
coelomate introns. Mar Biotech 4, 347-355
Kessing B, Croom H, Martin A, McIntosh C, McMillan WO, Palumbi S (1989) The simple
fool’s guide to PCR. Department of Zoology, Univ. of Hawaii, Honolulu
Kreitman M (1983) Nucleotide polymorphism at the alcohol dehydrogenase locus of Drosophila
melanogaster. Nature 304, 412-417
Kukla G (2005) Saalian supercycle, Mindel/Riss interglacial and Milankovitchs’ dating. Q Sci
Rev 24, 1573-1583
414
Lejeusne C, Chevaldonné P (2006) Brooding crustaceans in highly fragmented habitat: the
415
genetic structure of Mediterranean marine cave-dwelling mysid populations. Mol Ecol 15,
416
4123-4140
417
Lessios HA, Kessing BD, Robertson DR, Paulay G (1999) Phylogeography of the pantropical
418
sea urchin Eucidarcis in relation to land barriers and ocean currents. Evolution 53, 806-
419
817
13
420
López S, Turon X, Montero E, Palacín C, Duarte CM, Tarjuelo I. (1998) Larval abundance,
421
recruitment and early mortality in Paracentrotus lividus (Echinoidea). Interannual
422
variability and plankton-benthos coupling. Mar Ecol Prog Ser 172, 239-251
423
Lozano J, Galera J, López S, Turon X, Palacín C, Morera G (1995) Biological cycles and
424
recruitment of Paracentrotus lividus (Echinodermata: Echinoidea) in two contrasting
425
habitats. Mar Ecol Prog Ser 122, 179-191
426
Maldonado M (1985) Evolution of Mediterranean basins and a detailed reconstruction of the
427
Cenozoic paleocenography. In Western Mediterranean (eds Margalef R), Pergamon Press,
428
Oxford, pp. 17-60
429
McCartney MA, Keller G, Lessios HA (2000) Dispersal barriers in tropical oceans and
430
speciation in Atlantic and eastern Pacific sea urchins of the genus Echinometra. Mol Ecol
431
9, 1391-1400
432
433
434
435
436
437
438
Moberg PE, Burton RS (2000) Genetic heterogeneity among adult and recruit red sea urchins,
Strongylocentrotus franciscanus. Mar Biol, 136, 773-748
Mortensen T (1943). A Monograph of the Echinoidea. Camarodonta, pp.445. Copenhagen, C.A.
Reitzel
Nilsson T (1982) The Pleistocene: Geology and Life in the Quaternary Age. D. Ridel Publishing
Co., Dordrecht, Holland
Palacín C, Giribet G, Carner S, Dantart L, Turon X (1998) Low density of sea urchins influence
439
the structure of algal assemblages in the western Mediterranean. J Sea Res 39, 281-290
440
Palumbi SR (2004) Marine reserves and ocean neighbourhoods: The spatial scale of marine
441
442
443
444
445
446
447
448
449
450
451
452
453
populations and their management. Annu Rev Environ Resour 29, 31-68
Patarnello T, Volckaert FAMJ, Castilho R (2007) Pillars of Hercules: is the AtlanticMediterranean transition a phylogeographical break? Mol Ecol 16, 4426-4444
Pedrotti ML (1993) Spatial and temporal distribution and recruitment of echinoderm larvae in
the Ligurian Sea. J Mar Biol Assoc UK 73, 513–530
Pérès JM (1989) Historia de la biota mediterránea y la colonización de las profundidades. El
Mediterráneo Occidental (eds Margalef R), Omega S.A, Barcelona, pp. 200-235
Ramos-Onsins S, Rozas J (2002) Statistical properties of new neutrality tests against population
growth. Mol Biol Evol 19, 2092-2100
Rogers AR, Harpending H (1992) Population growth waves in the distribution of pairwise
genetic differences. Mol Biol Evol 9, 552-569
Rokas A, Ladoukakis E, Zouros E (2003) Animal mitochondrial DNA recombination revisited.
Trends Ecol Evol 18, 411-417
14
454
455
456
457
Rozas J, Sánchez del Barrio JC, Messeguer X, Rozas R (2003) DnaSP, DNA polymorphism
analyses by the coalescent and other methods. Bioinformatics 19, 2496-2497
Sala E, Boudouresque CF, Harmelin-Vivien M (1998) Fishing, trophies and the structure of
algal assemblages: evaluation of an old but untested paradigm. Oikos 82, 425-439
458
Savy S (1987) Les prédateurs de Paracentrotus lividus (Echinodermata). In: Colloque
459
International sur Paracentrotus lividus et les Oursins Comestibles (eds Boudouresque CF),
460
GIS Posidonie Publ, Marseille, pp.413-423
461
462
463
464
Sivertsen K (2006) Overgrazing of kelp beds along the coast of Norway. J Appl Phyc 18, 599610
Slatkin M, Hudson RR (1991) Pairwise comparisons of mitochondrial DNA sequences in stable
and exponentially growing populations. Genetics 129, 555-562
465
Smith AB, Pisani D, Mackenzie-Dodds JA et al. (2006) Testing the molecular clock: molecular
466
and paleontological estimates of divergence times in the Echinoidea (Echinodermata). Mol
467
Biol Evol 23, 1832-1851
468
Sponaugle S, Cowen RK, Shanks A, Morgan SG, Leis JM, Pineda J, Boehlert GW, Kingsford
469
MJ, Lindeman KC, Grimes C, Munro JL (2002) Predicting self-recruitment in marine
470
populations: biophysical correlates and mechanisms. Bull Mar Sci 70, 341-375
471
472
473
474
475
476
477
478
Stephens M, Smith NJ, Donnelly P (2001) A new statistical method for haplotype reconstruction
from population data. Am J Hum Genet 68, 978-989
Stephens M, Scheet P (2005) Accounting for decay of linkage disequilibrium in haplotype
inference and missing data imputation. Am J Hum Genet 76, 449-462
Tajima F (1989a) Statistical method for testing the neutral mutation hypothesis by DNA
polymorphism. Genetics 123, 585-595
Tajima F (1989b) The effect of change in population size on DNA polymorphism. Genetics 123,
597-601
479
Templeton AR, Boerwinkle E, Sing CF (1987) A cladistic analysis of phenotypic associations
480
with haplotypes inferred from restriction endonuclease mapping. I. Basic theory and an
481
analysis of alcohol dehydrogenase activity in Drosophila. Genetics 117, 343-351
482
Templeton AR, Sing CF (1993) A cladistic analysis of phenotypic association with haplotypes
483
inferred from restriction endonuclease mapping. IV. Nested Analyses with cladogram
484
uncertainty and recombination. Genetics 134, 659-669
485
486
487
488
Thomas FIM (1994) Physical properties of gametes in three sea urchin species. J Exp Biol 194,
263-284
Tintore J, La Violette PE, Blade I, Cruzado A (1998) A study of an intense density front in the
eastern Alboran Sea: the Almeria-Oran front. J Phys Oceanogr 18, 1384-1397
15
489
490
Tomas F, Romero J, Turon X (2004) Settlement and recruitment of the sea urchin Paracentrotus
lividus in two contrasting habitats in the Mediterranean. Mar Ecol Prog Ser 282, 173-184
491
Tomas F, Romero X, Turon X (2006) Experimental evidence that intra-specific competition in
492
seagrass meadows reduces reproductive potential in the sea urchin Paracentrotus lividus
493
(Lamarck). Scientia Mar 69, 475-484
494
495
Tsaousis AD, Martin DP, Ladoukakis ED, Posada D, Zouros E (2005) Widespread
recombination in published animal mtDNA sequences. Mol Biol Evol 22, 925-933
496
Turon X, Giribet G, López S, Palacín C (1995) Growth and population structure of
497
Paracentrotus lividus (Echinodermata: Echinoidea) in two contrasting habitats. Mar Ecol
498
Prog Ser 122, 193-204
499
500
Villablanca FX, Roderick GK, Palumbi SR (1998) Invasion genetics of the Mediterranean fruit
fly: variation in multiple nuclear introns. Mol Ecol 7, 547-560
501
Waelbroeck C, Labeurie L, Michel E, Duplessy JC, McManus JF, Lambeck K, Balbon E,
502
Labracherie M (2002) Sea-level and deep water temperature changes derived from benthic
503
foraminifera isotopic records. Quaternary Sci Rev 21: 295-305.
504
Watterson GA (1984) Allele frequencies after a bottleneck. Theor Pop Biol 26, 387-407
505
Zane L, Ostellari L, Maccatrozzo L Bargelloni L, Cuzin-Roudy J, Buchholz F, Patarnello T
506
(2000) Genetic differentiation in a pelagic crustacean (Meganyctiphanes norvegica,
507
Euphausiacea) from the North East Atlantic and the Mediterranean Sea. Mar Biol 136:
508
191-199
509
510
Zhang D, Hewitt GM (2003) Nuclear DNA analyses in genetic studies of populations: practice,
problems and prospects. Mol Ecol 12, 563-584
511
512
Acknowledgements
513
We thank S. Duran, S. López-Legentil, J. Sánchez-Fontenla, M. Rius, C. Palacín, A. Blanquer,
514
A. Corbacho, E. Cebrián and E. Macpherson for providing samples for this study. Two
515
anonymous reviewers and Associate Editor Thorsten Reusch provided comments that helped to
516
improve this manuscript. This study was funded by projects CTM2004-05265 and CTM2007-
517
66635 of the Spanish Government and by internal funds from the Museum of Comparative
518
Zoology, Harvard University. We declare that all the experiments done comply with current
519
Spanish laws.
520
16
521
Figure captions
522
Figure 1. Sampling scheme for Paracentrotus lividus. Codes for populations: 1) Ros: Roscoff;
523
2) San: Santander; 3) Fer: Ferrol; 4) Lis: Lisbon; 5) Cdz: Cádiz; 6) Tfe: Tenerife; 7) Ceu: Ceuta;
524
8) Tar: Tarifa; 9) Gata: Gata; 10) Nao: Nao; 11) Tos: Tossa de Mar; 13) Cad: Cadaqués; 14)
525
Cab: Cabrera; 15) Cor: Corsica; 16); Gre: Greece.
526
Figure 2. Multidimensional scaling (MDS) for 16S rRNA (a) and ANT (b). Open circles
527
represent Atlantic populations (including Ceuta and Tarifa) whereas filled circles represent
528
Mediterranean populations.
529
Figure 3. Minimum Spanning Network for 16S rRNA. Circles represent the 38 haplotypes
530
observed in our sample. Areas of the circles are proportional to the number of sampled
531
individuals. Partitions inside the circles represent the proportion of each population within each
532
haplotype. Red dots represent missing, probably unsampled haplotypes or extinct sequences.
533
Figure 4. Mismatch distribution for 16S rRNA and ANT for each basin. The dashed line
534
represents the expected distribution under a sudden expansion model; the solid line represents
535
the observed distribution. Ceuta and Melilla are considered as belonging to the Atlantic
536
group of populations.
537
17
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
Ros
45ºN
Fer
San
Cor
Cad
Tos
Col
Lis
Nao
Tar
Cdz
Gata
Ceu
Cab
Gre
Tfe
0º
15ºE
18
Pairwise nucleotide differences
(a)
(b)
568
569
19
570
571
572
573
574
575
576
577
578
579
580
Roscoff
Santander
Ferrol
Lisbon
Cádiz
Tenerife
Ceuta
Tarifa
Gata
Nao
Columbretes
Tossa
Cadaqués
Cabrera
Corsica
Greece
20
581
16S (Atlantic)
1200
Observed
Expected
1000
Frequency
800
600
400
200
0
0
1
2
3
4
5
6
7
Pairwise differences
582
16S (Mediterranean)
1000
Observed
Expected
800
Frequency
600
400
200
0
0
2
4
6
8
10
Pairwise differences
583
ANT (Atlantic)
1800
Observed
Expected
1600
1400
Frequency
1200
1000
800
600
400
200
0
0
584
2
4
6
8
10
12
14
Pairwise differences
21
585
ANT (Mediterranean)
2000
Observed
Expected
Frequency
1500
1000
500
0
0
586
587
2
4
6
8
10
12
14
16
Pairwise differences
22
588
589
590
591
592
593
594
Table 1. Diversity measures for the populations of Paracentrotus lividus studied. Number of individuals per
population (N). Number of haplotypes per population (Nh), out of which the number of private alleles is shown in
brackets. Two haplotypes in the Atlantic and 5 in the Mediterranean are exclusive from each basin, but shared by
more than one population. Haplotypic (H) and nucleotidic (π) diversity, with standard deviations presented in
brackets. Inbreeding coefficient for ANT. Asterisks represent significant coefficients at p<0.05. H exp represents the
expected heterozygosity and Hobs represents the observed heterozygosity. Ceuta and Tarifa were considered in the
Atlantic group of populations following results from SAMOVA (see text).
Population
16S
N
H
0.563
(±0.062)
0.839
Roscoff
9
(±0.110)
0.378
Santander
10
(±0.181)
0.733
Ferrol
10
(±0.101)
0.564
Lisbon
11
(±0.134)
0.818
Cádiz
11
(±0.119)
0.250
Tenerife
8
(±0.18)
0.533
Ceuta
10
(±0.180)
0.200
Tarifa
10
(±0.154)
0.839
79
MED
(±0.020)
0.667
Gata
10
(±0.163)
0.952
Nao
7
(±0.96)
0.894
Columbretes 12
(±0.63)
0.885
Tossa
13
(±0.064)
0.929
Cadaqués
8
(±0.084)
0.944
Cabrera
9
(±0.070)
0.929
Corsica
8
(±0.84)
0.970
Greece
12
(±0.044)
158 0.785
Total
(±0.032)
ATL
79
Π
0.00189
(±0.00027)
0.00239
(±0.00057)
0.00072
(±0.00037)
0.00259
(±0.00057)
0.00184
(±0.00049)
0.00262
(±0.00067)
0.00043
(±0.00031)
0.00179
(±0.00075)
0.0076
(±0.00058)
0.00407
(±0.00028)
0.00261
(±0.00102)
0.00375
(±0.00080)
0.00390
(±0.00052)
0.00406
(±0.00064)
0.00413
(±0.00070)
0.00466
(±0.00076)
0.00429
(±0.00068)
0.00374
(±0.00066)
0.00333
(±0.00025)
ANT
Nh
(private)
N
H
Π
16 (9)
75
4 (1)
9
4
9
7 (2)
10
4
10
7 (3)
11
3
8
6
10
3 (1)
8
29 (22)
76
7 (4)
10
7
6
7 (1)
11
7 (4)
13
6 (2)
8
7 (2)
9
6
8
10 (4)
11
38
151
0.975
(±0.006)
0.961
(±0.034)
0.928
(±0.040)
0.968
(±0.028)
1
(±0.016)
0.961
(±0.028)
0.933
(±0.48)
0.995
(±0.018)
0.992
(±0.025)
0.959
(±0.011)
0.953
(±0.028)
0.939
(±0.058)
0.931
(±0.046)
0.975
(±0.021)
0.958
(±0.036)
0.958
(±0.033)
0.942
(±0.048)
0.961
(±0.024)
0.9672
(±0.0060)
0.01397
(±0.00058)
0.01360
(±0.00191)
0.01267
(±0.00185)
0.01367
(±0.00162)
0.01623
(±0.00149)
0.01330
(±0.00118)
0.01026
(±0.00112)
0.01486
(±0.00131)
0.01442
(±0.00193)
0.01230
(±0.00064)
0.01014
(±0.00141)
0.00577
(±0.00077)
0.01072
(±0.00180)
0.01259
(±0.00131)
0.01164
(±0.00148)
0.01409
(±0.00189)
0.01099
(±0.00110)
0.01608
(±0.00168)
0.01316
(±0.00044)
Nh
Inbreeding
H exp.
(private) Coefficient
87 (60)
0.0621*
0.913
H obs.
0.8108
14 (5)
0.200*
0.9074
0.778
11 (6)
0.045
0.8765
0.889
16 (5)
0.182*
0.920
0.800
20 (12)
-0.006
0.945
1.000
16 (8)
0.349*
0.917
0.636
11 (4)
0.611*
0.867
0.375
19 (10)
-0.006
0.945
1.000
15 (4)
-0.0091
0.930
1.000
82 (56)
0.098*
0,901
0.803
13 (3)
0.383*
0.905
0.600
9 (2)
-0.071
0.861
1.000
16 (8)
0.024
0.888
0.909
21 (12)
0.143*
0.944
0.846
12 (3)
0.229*
0.898
0.750
15 (8)
0.069
0.915
0.900
12 (9)
0.075
0.883
0.875
15 (10)
0.444*
0.917
0.546
142
0.080*
0.907
0.806
595
596
23
597
598
599
Table 2. Analyses of Molecular Variance for 16S rRNA (a) and ANT (b). Ceuta and Tarifa were considered as
600
(a) 16S
Source of variation
belonging to the Atlantic group of populations. Asterisks represent significant tests at p<0.05 after 10,000
permutations. Va, Vb and Vc are the associate covariance components. FSC, FST and FCT are the F-statistics.
Df
Sum of
squares
Variance
components
% of
variation
Fixation
indices
AMOVA between basins
Among groups
1
10.646
0.12237 Va
11.20*
FCT: 0.00086
Among populations within groups
14
13.689
0.00084 Vb
0.08
FSC: 0.11275
Within populations
142
137.678
0.96956 Vc
88.72*
FST: 0.11198
Total
157
162.013
1.09277
Among populations without groups
15
24.335
0.06622 Va
6.39*
FST: 0.06393
Within populations
142
137.678
0.96956 Vb
93.61
Total
157
162.013
1.03578
AMOVA without groups
601
602
(b) ANT
Source of variation
Df
Sum of
squares
Variance
components
% of
variation
Fixation
indices
Among groups
1
3.561
0.00207 Va
0.10
FCT: 0.02848
Among populations within groups
14
44.924
0.06050 Vb
2.84*
FSC: 0.02942
Within populations
288
594.455
2.06408 Vc
97.06*
FST: 0.00098
Total
303
642.941
2.12665
Among populations without groups
15
47.260
0.05706 Va
2.68*
FST: 0.02684
Within populations
288
595.782
2.06869 Vb
97.32
Total
303
643.043
2.12575
AMOVA between basins
AMOVA without groups
603
24
604
605
Table 3. Tests of neutrality for 16S rRNA and ANT. Ceuta and Tarifa were considered as belonging to the Atlantic
group of populations. Asterisks represent significant results: * p<0.05; ** p<0.01; *** p<0.002.
16S
Atlantic
Mediterranean Total
Tajima’s D
-1.73683
-1.65461
-1.90111*
Fu’s Fs
-7.010***
-18.888***
-30.171***
R2
0.042*
0.043*
0.029*
ANT
Atlantic
Mediterranean Total
Tajima’s D
-1.47953
-1.89368
-1.85294*
Fu’s Fs
-123.389***
-116.093***
-246.361***
R2
0.0501
0.0395*
0.0369*
606
607
25
608
609
610
611
Table 4. Parameters of population expansion for both markers, for each basin separately and for the whole sample.
Ceuta and Tarifa were considered as belonging to the Atlantic group of populations. Parameters are estimated with a
confidence interval of 0.01 for 10,000 bootstrap replicates. SDD (sum of square deviations) with its respective p
values in brackets.
16S
Parameters
ANT
Atlantic
Mediterranean
Total
Atlantic
Mediterranean
Total
Τ
1.581
2.097
2.043
4.544
4.176
4.510
Θo
0.113
0.217
0.159
0.643
0.580
0.560
Θ1
2031.363
4330.372
3687.971
1379.746
1149.083
1210.135
0.003005
(0.14)
0.000574
(0.12)
0.00057
(0.080)
0.002188
(0.29)
0.0016
(0.34)
0.001098
(0.61)
0.0013
(054)
0.001279
(0.42)
0.0017
(0.39)
Goodness-of-fit test (SDD)
Demographic
expansion
Spatial
expansion
0.00108
(0.54)
0.000982
(0.60)
-------
612
26
613
APPENDIX
614
Table 1. Absolute haplotype frequencies for 16S. Population codes correspond to codes in Figure 1.
615
Hapl Ros San Fer Lis Cdz Tfe Ceu Tar ATL Gata Nao Col Tos Cad Cab Cor Gre MED
1
4
4
3
2
2
1
1
3
2
4
1
3
4
5
5
3
8
36
1
6
4
3
1
1
2
2
1
1
2
1
1
2
2
1
12
8
1
1
9
1
1
1
1
4
1
2
1
12
1
1
12
1
1
13
1
1
14
1
1
1
1
2
1
5
1
1
1
1
1
1
1
1
1
1
3
11
1
1
3
1
1
4
2
2
14
1
1
17
1
1
1
2
4
2
19
1
1
2
2
20
1
18
2
10
1
16
1
4
7
15
1
1
5
10
4
2
2
10
3
1
1
6
21
1
1
22
1
1
3
1
23
2
2
24
1
1
25
1
1
26
1
1
27
1
1
28
1
1
29
1
1
30
1
1
31
1
1
32
1
1
33
1
1
34
1
1
35
1
1
36
1
1
27
37
1
1
38
1
1
Total
9
10
11 10 11
8
10
10
79
10
7
12 13
8
9
8
12
79
616
617
28
618
619
Table 2. Haplotypic phase for each individual form results obtained with PHASE. Homozygotic individuals are
marked with an asterisk. Population codes correspond to codes in Figure 1.
ATLANTIC
Individual
Haplotye
Individual
Haplotye
ROSCOFF
CÁDIZ
H1, H2
H8, H49
Ros1
Cdz1
MEDITERRANEAN
Individual
Haplotye
Individual
Haplotye
GATA
CADAQUÉS
H134, H135 Cad2
H7, H85
Gata2
H3, H4
Ros2
H5, H6
Ros3
H7, H7
Ros4*
H8, H9
Ros5
H10, H10
Ros6*
H11, H12
Ros8
H7, H13
Ros9
H10, H14
Ros10
SANTANDER
H1, H15
San2
H7, H10
San3
H16, H17
San4
H18, H18
San5*
H19, H20
San7
H21, 22
San8
H1, H7
San9
H7, H17
San10
H7, H17
San11
FERROL
H1, H23
Fer1
H7, H7
Fer2*
H24, H25
Fer3
H11, H26
Fer4
H1, H1
Fer5*
H27, H28
Fer9
H7, H17
Fer11
H29, H30
Fer13
H31, H32
Fer14
H33, H34
Fer17
LISBON
H7, H27
Lis2
H35, H36
Lis3
H10, H37
Lis6
H34, H38
Lis7
H39, H40
Lis8
H41, H42
Lis9
H8, H42
Lis10
H43, H44
Lis15
H45, H46
Lis16
H47, H48
Lis17
Gata3
Gata4
Gata10*
Gata11*
Gata13
Gata15
Gata16
Gata17*
Gata18
H8, H50
H26, H51
H7, H7
H52, H52
H53, H53
H8, H8
H54, H55
H56, H57
H10, H58
H29, H59
TENERIFE
H7, H60
Tfe1
H12, H61
Tfe3
H10, H62
Tfe4
H1, H1
Tfe6*
H63, H63
Tfe7*
H1, H1
Tfe10*
H64, H64
Tfe11*
H65, H65
Tfe13*
CEUTA
H66, H67
Ceu1
H28, H68
Ceu2
H69, H70
Ceu4
H17, H71
Ceu6
H72, H73
Ceu8
H7, H74
Ceu11
H75, H76
Ceu12
H12, H77
Ceu13
H49, H78
Ceu14
H7, H79
Ceu15
TARIFA
H1, H7
Tar1
H7, H80
Tar2
H28, H68
Tar3
H81, H82
Tar4
H83, H84
Tar5
H12, H49
Tar7
H10, H85
Tar9
H29, H86
Tar14
Cdz2
Cdz3
Cdz4*
Cdz5*
Cdz6*
Cdz7*
Cdz8
Cdz9
Cdz10
Cdz11
H68, H136
H137, H137
H68, H68
H1, H1
H7, H138
H7, H139
H7, H140
H141, H141
H2, H142
NAO
H7, H85
Nao5
H96, H97
Nao6
H31, H85
Nao11
H10, H37
Nao12
H17, H83
Nao14
H10, H85
Nao20
COLUMBRETES
H7, H8
Col1
H7, H91
Col3
H92, H93
Col4
H7, H87
Col8
H31, H88
Col9
H31, H85
Col10
H89, H90
Col12
H10, H94
Col15
H72, H95
Col16
H7, H7
Col17*
H7, H28
Col18
TOSSA
H85, H98
Tos1
H92, H92
Tos2*
H8, H12
Tos4
H7, H99
Tos5
H100, H101
Tos6
H1, H102
Tos7
H103, H104
Tos8
H105, H106
Tos9
H8, H84
Tos10
H22, H53
Tos11
H17, H107
Tos18
H7, H7
Tos33*
H108, H109
Tos34
H1, H110
Cad4
H17, H17
Cad10*
H8, H51
Cad15
H7, H7
Cad17*
H13, H51
Cad18
H111, H112
Cad19
H9, H61
Cad23
CABRERA
H7, H83
Cab19
H113, H114
Cab22
H7, H9
Cab23
H38, H115
Cab24
H116, H117
Cab25
H118, H119
Cab26
H7, H120
Cab27
H7, H8
Cab34
H72, H72
Cab40*
CORSICA
H7, H7
Cor5*
H7, H8
Cor6
H92, H131
Cor12
H31, H85
Cor13
H132, H133
Cor17
H8, H30
Cor18
H1, H2
Cor19
H7, H17
Cor20
GREECE
H7, H7
Gre1*
H121, H121
Gre2*
H122, H123
Gre5
H7,H60
Gre7
H124, H125
Gre8
H1, H1
Gre9*
H126, H127
Gre10
H128, H128
Gre12*
H129, H129
Gre14*
H34, H73
Gre16
H1, H130
Gre17
620
29
621
622
623
624
625
626
627
628
Table 3. Absolute frequencies of each ANT haplotype per basin.
Haplotype ATL
MED
Total
Haplotype ATL
MED Total Haplotype ATL MED To
11
8
19
1
1
2
1
1
H1
H37
H73
1
2
3
1
1
2
1
H2
H38
H74
1
1
1
1
1
H3
H39
H75
1
1
1
1
1
H4
H40
H76
1
1
1
1
1
H5
H41
H77
1
1
2
2
1
H6
H42
H78
18
27
45
1
1
1
H7
H43
H79
6
7
13
1
1
1
H8
H44
H80
1
2
3
1
1
1
H9
H45
H81
8
3
11
1
1
1
H10
H46
H82
2
2
1
1
1
2
H11
H47
H83
4
1
5
1
1
1
1
H12
H48
H84
1
1
2
3
3
1
7
H13
H49
H85
1
1
1
1
1
H14
H50
H86
1
1
1
2
3
1
H15
H51
H87
1
1
2
2
1
H16
H52
H88
5
5
10
2
1
3
1
H17
H53
H89
2
2
1
1
1
H18
H54
H90
1
1
1
1
1
H19
H55
H91
1
1
1
1
4
H20
H56
H92
1
1
1
1
1
H21
H57
H93
1
1
2
1
1
1
H22
H58
H94
1
1
1
1
1
H23
H59
H95
1
1
1
1
2
1
H24
H60
H96
1
1
1
1
2
1
H25
H61
H97
2
2
1
1
1
H26
H62
H98
2
2
2
2
1
H27
H63
H99
3
1
4
2
2
1
H28
H64
H100
3
3
2
2
1
H29
H65
H101
1
1
2
1
1
1
H30
H66
H102
1
4
5
1
1
1
H31
H67
H103
1
1
2
3
5
1
H32
H68
H104
1
1
1
1
1
H33
H69
H105
2
1
3
1
1
1
H34
H70
H106
1
1
1
1
1
H35
H71
H107
1
1
1
3
4
1
H36
H72
H108
Table 4. FST values for 16S (lower diagonal) and for ANT (upper diagonal) based on pairwise nucleotide
differences. Population codes correspond to codes in Figure 1. Values in bold represent significant comparisons
for p<0.05. The empty diagonal represents FST=0. Note that, for 16S (lower diagonal) more than 50% of
comparisons among basins are significant whereas only when comparing Ferrol to Santander and Tarifa in the
Atlantic FST values were significant within basins. As for ANT, Nao shows a striking difference with all other
populations, all comparisons being significant.
-0.00235 -0.01301 -0.00096
0.04941
-0.02116 0.03136
0.00810 0.16986
0.01153
-0.07614 0.06913 -0.02859
-0.02640 -0.01190 0.07542 -0.00100
0.06756
0.03762
0.06147
0.08728
-0.00596
0.00192
-0.02138
0.00801
0.07553
0.01111 -0.02075
0.01399 -0.01079
0.01868 -0.02204
0.04769 0.01527
0.08201 0.05606
0.04582
0.01231
0.02500
0.07672
0.04519
-0.01875 0.
-0.00307 -0.0
0.01228 -0.0
0.04081 0.0
0.04909 0.0
0.10380
0.14511
0.17063
0.15730
0.17424
30
0.02391 -0.07781
-0.01734 -0.00168
0.05868 0.06504
0.05262 0.02597
0.07612 0.13354
0.12152 0.10075
0.08484 0.11534
0.17250 0.21468
0.16156 0.16475
0.15419 0.15304
0.02811 0.03854
Ros
San
0.15995
0.07407
0.22087
0.15500
0.19000
0.19393
0.17737
0.25810
0.22565
0.21641
0.12745
Fer
0.04469 -0.06833
0.04238 0.00362 0.03855 0.24229 0.03381 0.0
-0.02882 -0.03894 -0.04876
-0.02963 0.08477 0.14719 0.02407 0.0
0.07231 -0.00458 -0.02132 0.00383
0.05282
0.13716 0.00021 -0.0
0.06420 0.01650 -0.01112 -0.00473 0.01961
0.17311 0.03043 0.0
0.10487 0.03466 0.07082 0.02015 0.09386 -0.04184
0.06629 0.1
0.09510
0.07812
0.08016
0.10733
-0.03670
-0.05160
0.13220
-0.0
0.10061 0.07158 0.06797 0.04523 0.06030 -0.02893 -0.03872 0.00282
0.20163 0.17081 0.19481 0.12778 0.22381 0.01189 -0.02028 -0.00445 -0.0
0.17570 0.13886 0.14214 0.13680 0.18024 0.01538 -0.02448 -0.04575 0.0
0.16740 0.12880 0.13393 0.09005 0.17764 -0.04002 -0.04965 -0.07413 -0.0
0.04849 0.02787 0.00836 -0.01103 0.03115 -0.05117 0.00128 0.02273 -0.0
Lis
Cdz
Tfe
Ceu
Tar
Gata
Nao
Col
T
629
630
631
31
Download