MURDOCH RESEARCH REPOSITORY

This is the author’s final version of the work, as accepted for publication following peer review but without the publisher’s layout or pagination. The definitive version is available at http://dx.doi.org/10.1007/s10641-012-0105-3

Moore, G.I. and Chaplin, J.A. (2013) Population genetic structures of three congeneric species of coastal pelagic fishes (: Arripidae) with extensive larval, post-settlement and adult movements. Environmental Biology of Fishes, 96 (9). pp. 1087-1099.

http://researchrepository.murdoch.edu.au/17252/

Copyright: © 2013 Springer Science+Business Media Dordrech.

It is posted here for your personal use. No further distribution is permitted.

1 Population genetic structures of three congeneric species of coastal pelagic fishes (Arripis:

2 Arripidae) with extensive larval, post-settlement and adult movements.

3

4 G.I. Moore1,2 and J.A. Chaplin1

5

6 1 Centre for Fish and Fisheries Research, School of Biological Sciences and Biotechnology, Murdoch

7 University, South St, Murdoch, Western Australia, Australia, 6150.

8 2 Fish Section, Department of Aquatic Zoology, Western Australian Museum, 49 Kew St., Welshpool,

9 Western Australia, Australia, 6106.

10

11 Author to whom correspondence should be addressed:

12 G.I. Moore

13 e-mail: [email protected]

14 phone: +61 08 9212 3744 15

1 16 Abstract

17

18 The population genetic structures of three congeneric coastal pelagic marine fishes (, A.

19 truttaceus and A. georgianus) were investigated to determine whether these structures were consistent

20 with the apparently high gene flow life histories of these species. This investigation used fragment

21 length polymorphisms at two to four nDNA intron loci (amplified by EPIC-PCR) in samples of each

22 species collected from across their entire Australian distributions. The results revealed no evidence of

23 genetic subdivision in any species based on an analysis of variation either across samples (with

24 multilocus FST values ranging from effectively zero to 0.005), or between pairs of samples. These

25 findings, when considered in combination with other available evidence, are consistent with the view

26 that each of these species represents a rare example of a coastal fish species that is genetically

27 homogeneous over a vast area, including the entire geographic range (A. truttaceus and A. georgianus)

28 or Australian distribution (A. trutta).

29

30 Keywords: EPIC-PCR, Leeuwin Current, East Australian Current, migratory, high gene-flow,

31 Australian Salmon

2 32 Introduction

33 It was often traditionally assumed that coastal pelagic marine fishes exhibit genetic homogeneity, and

34 possibly even panmixia (random mating), over their geographic distributions (see DeLigny &

35 Pantelouris 1973; Waples 1987; Ward et al. 1994; Hauser & Ward 1998 for discussion). However,

36 most genetic studies of coastal fishes with a seemingly highly dispersive life history have detected the

37 presence of population subdivision (e.g. Gold et al. 2002; Knutsen et al. 2003; Rohfritsch & Borsa

38 2005; Ruzzante et al. 2006; Atarhouch et al. 2007; Chairi et al. 2007; Chlaida et al. 2008; Wang et al.

39 2008; Andre´ et al. 2011, but see Hoolihan et al. 2006; Cárdenas et al. 2009; Palm et al. 2009). Indeed,

40 in view of the results of recent genetic studies, it has been argued that panmixia is unexpected even in

41 widespread fish species that are distributed across different environments (Als et al. 2011).

42

43 Pelagic species with a life history that involves high dispersal potential at every life stage (larval,

44 juvenile and adult) could be expected to be the most likely candidates to exhibit extensive gene flow,

45 generating a considerable homogenising effect, if not panmixia (Waples 1987; Ward et al. 1994; Siegel

46 et al. 2003). Although this type of life history is uncommon among coastal fishes (Gillanders et al.

47 2003), members of the family Arripidae (Australian Salmon) exhibit impressive vagility throughout

48 their lives. In fact, few coastal marine fishes appear to share the dispersal potential of Arripis species

49 (Ward et al. 1994; Gillanders et al. 2003) and A. truttaceus, in particular, may well have the greatest

50 documented migration from juvenile to adult distribution of any marine fish (Gillanders et al. 2003).

51

52 The family Arripidae is endemic to temperate Australasian seas and contains a single genus Arripis,

53 and four currently recognised species. In Australian mainland waters, the family is represented by the

54 Eastern Australian Salmon Arripis trutta (Bloch & Schneider 1801), the Western Australian Salmon

55 Arripis truttaceus (Cuvier 1829) and the Australian Herring Arripis georgianus (Valenciennes 1831).

56 The latter two species are endemic to these waters, while the former also occurs in New Zealand waters

57 (Paulin 1993). A fourth species, the Northern Kahawai Arripis xylabion Paulin 1993 occurs around

58 islands north of New Zealand (and possibly also in New Zealand waters), but is not included in the

59 present study. All species are targeted in long-term, low financial value recreational and commercial

60 fisheries and there is a long research history for each in the context of fisheries management (e.g.

61 Fairbridge 1950; Malcolm 1960; Nicholls 1971; Lenanton & Hall 1976; Stanley 1988a; Cappo et al.

62 2000; Fairclough et al. 2000a; Stewart et al. 2011).

3 63

64 The three Australian species of Arripis have complex life histories involving impressive migrations

65 (see Fig. 1). For example, adult A. truttaceus are distributed around the south-west corner of Australia.

66 Breeding adults migrate to, and concentrate in, the western-most parts of this area to spawn (Malcolm

67 1960; Stanley 1988b). The planktonic larvae are then transported hundreds or thousands of kilometres

68 to the south-east by the Leeuwin Current and wind-generated water flow (Petrusevics & Bye 1995;

69 Caputi et al. 1996). Post-settlement and first-year A. truttaceus occur along much of the southern coast

70 of Australia but are most abundant in south-eastern Australian waters (Malcolm 1960). Juvenile and

71 sub-adult fish migrate progressively westward, and enter the adult distribution in their fourth year of

72 life (Malcolm 1960; Hoedt & Dimmlich 1994). The migratory life history of A. trutta reflects that of

73 A. truttaceus, except that the individuals follow a north-south route along the east coast of Australia

74 (Stanley & Malcolm 1977; Stanley 1988c; Stewart et al. 2011; Fig. 1). Likewise, the migratory life

75 history of A. georgianus is very similar to that of A. truttaceus (Fig. 1), except that A. georgianus may

76 have more local larval retention (Lenanton 1982; Fairclough et al. 2000b), and individuals enter the

77 adult distribution at two years of age (Fairclough et al. 2000a).

78

79 On the basis of possessing a high dispersal potential at every life stage, it is tempting to hypothesize

80 that the species of Arripis provide some rare examples of coastal pelagic species that are genetically

81 homogeneous over their entire range (see Waples 1987; Ward et al. 1994; Siegel et al. 2003). The only

82 published test of this hypothesis was an allozyme-based study of the population structure of A.

83 georgianus that was conducted by Ayvazian et al. (2004), as part of a multi-disciplinary stock

84 assessment of this species. Ayvazian et al. (2004) found no evidence of population genetic subdivision

85 in this species, but based on a test with very limited power due to a very low level of allozyme

86 polymorphism. There is clearly a need to find more polymorphic markers to improve the power of the

87 analyses and to study the population genetics of the other Arripis species.

88

89 The present study used nDNA markers in the form of fragment length polymorphism at several intron

90 loci to investigate the population genetic structures of Arripis species. These polymorphisms are

91 assayed using ‘universal’ primers in exon-primed, intron-crossing PCR (EPIC-PCR, Palumbi and

92 Baker 1994) and have been used to investigate population and evolutionary questions in other fishes

93 and also invertebrate species (e.g. Bierne et al. 2000; Berrebi et al. 2005; Atarhouch et al. 2007;

4 94 Fauvelot et al. 2007; Rolland et al. 2007; Li et al. 2010; Boissin et al. 2011). This type of nDNA

95 marker was used because it provides many of the advantageous qualities of microsatellite markers (e.g.

96 highly polymorphic, single locus, co-dominant markers; e.g. Li et al. 2002), but avoids the time and

97 cost required to developed customised primers, as would have been the case for a microsatellite-based

98 study. This latter point is particularly relevant in the context of this study because the species of

99 Arripis are of limited financial value and have a relatively low profile from a conservation perspective,

100 and hence attracting funding for molecular work is difficult. In addition, there is currently no published

101 information on microsatellite loci in the Arripidae and, given the uncertainty of the systematic position

102 of this family (e.g. Johnson and Fritzche 1989; Yagishita et al. 2009), there are no obviously related

103 groups to target for cross-species transfer of microsatellite primers (see Selkoe & Toonen 2006).

104 Furthermore, transferability of microsatellite loci between fish families is generally poor (e.g. Barbará

105 et al. 2007).

106

107 The objective of this study was to use length frequency polymorphisms at intron loci to investigate the

108 population genetic structures of Arripis trutta, A. truttaceus and A. georgianus, across their entire

109 ranges in Australian mainland waters. This was done to determine whether these structures are

110 consistent with the hypothesized high gene-flow life histories of these species.

111

112 Materials and Methods

113

114 Sample Collection

115

116 For each species, the collection locations were distributed across their Australian range (Fig. 1). For A.

117 trutta, 29 – 30 individuals were sampled from each of three locations in 2007. For A truttaceus, 13 –

118 45 individuals were sampled from each of five locations in 2007 or 2008, while for A. georgianus 9 –

119 44 individuals were collected from each of seven locations in 2006 or 2008 (see Table 2). For each

120 species, a combination of juveniles and adults was sampled.

121

122 Genetic Assays

123 Total genomic DNA was isolated from approximately 5 mg of muscle tissue from Arripis individuals

124 using a MasterPureTM extraction kit according to the manufacturer’s instructions.

5 125

126 Fragment length polymorphism at two to four intron loci, depending on the species, was assayed. The

127 loci were: Aldolase B, intron 4 (AldoB4); Calmodulin, intron 4 (CaM4); S7 ribosomal protein gene,

128 intron 2 (S72), and Creatin kinase (CK7). The loci were amplified by EPIC-PCR using published

129 primers (from Chow & Hazama 1998; Chow & Takeyama 1998; Hassan et al. 2002). These four loci

130 were chosen from a pool of 28 intron loci because, in preliminary trials, they were the only loci that

131 amplified and exhibited polymorphism in at least one species of Arripis (Moore 2012).

132

133 PCR reactions were carried out in a GeneAmp 9700 Thermal Cycler (Applied Biosystems) with: (i) an

134 initial denaturation phase of 5 min at 94˚C, (ii) 38 amplification cycles, with each cycle consisting of

135 30 s of denaturation at 94°C, 30 s of annealing at optimised temperatures (Table 1), 30 s of extension at

136 72°C, and (iii) a final 20 min extension at 72°C (to standardise poly-A additions). The only exceptions

137 were: (i) the locus CK7 in A. georgianus, which was amplified using a touch-down PCR of 57

138 amplification cycles, with annealing starting at 58°C for 60 s and decreasing by 0.3°C and 1 s each

139 cycle; and (ii) the locus CaM4 in A. georgianus, which was amplified using a touch-down PCR of 50

140 amplification cycles, with annealing starting at 65°C for 60 s and decreasing by 0.4°C each cycle

141 (Table 1). The PCR reaction mixture contained 0.1 mM of each dNTP (Promega), 0.25U Taq DNA

142 polymerase (Roche), 1.5 µl Taq Buffer (100 mM Tris-HCl, 15 mM MgCl2, 500 mM KCl, pH 8.3;

143 Roche), 0.04-0.06 µM of each primer (Table 1; forward primers were fluorescently labelled), and

144 approximately 10 ng of DNA template, adjusted to a volume of 15 µl with PCR grade H2O.

145

146 PCR products were assayed either singularly or multiplexed for one large and one small locus, using

147 optimised volumes for each locus (Table 1). Fragment length was determined with an ABI 3730

148 Automated Sequencer (Applied Biosystems) using LIZ1200 (Applied Biosystems) as the size standard.

149 Allele peaks were manually scored to one decimal place using the software GENEMARKER v. 1.71

150 (SoftGenetics Inc.). The data points for the allele scores of all individuals of all species at a locus

151 grouped into a series of discrete clusters, with variability around each allele and gaps between each

152 allele cluster, as is typical for automated sizing of DNA fragments (see Idury & Cardon 1997; Anon.

153 2004). Each allele was named by the modal integer of the associated cluster of data points. It is

154 possible that some similar but different sized alleles have inadvertently been placed into the same allele

6 155 cluster. Both positive (individuals for which the genotype was already known) and negative (no added

156 DNA template) controls were included in assays.

157

158 Data Analyses

159 The program MICRO-CHECKER (v. 2.2.3; van Oosterhout et al. 2004) was used to test for the

160 presence of null alleles at each locus in each sample of each species. For each species, the number of

161 alleles per locus (NA), observed heterozygosity (HO) and expected heterozygosity (HE, Nei 1987) were

162 calculated for each sample and over all samples. Exact tests were used to assess departures from

163 Hardy-Weinberg Equilibrium (HWE) conditions at each locus in each sample. Fisher’s exact

164 probability tests were used to test for departures from linkage disequilibrium for each pairwise

165 combination of loci, in each sample of each species. An exact chi-square test with Fisher’s combined

166 probability method was used as a global test of linkage disequilibrium for all samples within a species.

167 The above exact tests were conducted in GENEPOP v. 4.0 (Raymond & Rousset 1995) with 10,000

168 dememorisation steps, 1,000 batches and 10,000 iterations. Where appropriate, a sequential Bonferroni

169 correction (Rice 1989) was applied to control for group-wide Type I error.

170

171 For each species, single-locus FST estimates were made using FSTAT v. 2.9.3.2 (Goudet 1995),

172 assuming samples in HWE and using 10,000 permutations to test statistical significance. The value of

173 FST over all loci was calculated using the software GENETIX v. 4.05.2 (Belkhir et al. 2004) and again

174 using 10,000 permutations to test the statistical significance. Pairwise FST tests between samples

175 within each species were conducted in FSTAT, with statistical significance determined as above.

176 Finally, the factorial correspondence analysis (FCA) of GENETIX was used to search for structure in

177 the multilocus genotype data, based on individuals rather than samples, following Guinand (1996).

178

179 A retrospective power analysis was conducted with 1000 runs using the method described by Theisen

180 et al. (2008) and the software POWSIM v. 4.0 (Ryman & Palm 2006). Briefly, this involved simulated

181 populations that were allowed to drift to various levels of genetic divergence (quantified by FST, as a

182 product of the number of generations, t, and population size, Ne). Power was determined by finding

183 the least number of generations of drift that could be detected among the sampled populations of each

184 species with a power ≥ 95%, while holding the population size constant at 1000. This analysis only

185 included the individuals of a species that were typed for all polymorphic loci for that species (A. trutta

7 186 N = 82, A. truttaceus N = 119, A. georgianus N = 188), because this software cannot accept different

187 sample sizes for different loci within a sample.

188

189 Results

190

191 The AldoB4 and S72 loci were polymorphic in all three species, while CK7 was polymorphic only in

192 Arripis trutta and A. georgianus and CaM4 was polymorphic only in A. georgianus (Table 2). The

193 alleles ranged in size from 169-191 bp at locus AldoB4, from 469-504 bp at CaM4, 844-984 bp at CK7

194 and 816-931 bp at S72 (Table 2; details in Moore 2012). The overall level of polymorphism was

195 greatest in the samples of A. georgianus (four polymorphic loci, with a total of 32 different alleles),

196 followed by those of A. trutta (three polymorphic loci, with a total of 19 different alleles), while those

197 of A. truttaceus contained only two polymorphic loci, with 7 different alleles (Table 2). Mean expected

198 heterozygosity per sample per locus, when the locus was polymorphic in the sample, was typically low

199 to moderate (e.g. between 0.083 and 0.655; Table 2). The results from MICRO-CHECKER revealed

200 no evidence for the presence of null alleles or large allele dropout at any locus in any sample in any

201 species. In fact, only a single departure from HWE was detected (locus CK7, HO sample of A.

202 georgianus) and even this was not statistically significant following Bonferroni correction (Table 2).

203 No cases of significant linkage disequilibrium were detected, after Bonferroni correction, for any pair

204 of loci in any sample of any species, suggesting independent loci (data not presented).

205

206 Within each species, single locus estimates of (global) FST were consistently low (Table 3).

207 Accordingly, the (global) multi-locus values of FST for each species were also low (e.g. ≤ 0.005) and

208 not significantly different from that expected by chance (Table 3). Similarly, the values of FST

209 between each pair of samples within each of the three study species were invariably low (e.g.

210 effectively zero for A. trutta, FST < 0.018 for A. truttaceus, and FST < 0.030 for A. georgianus) and

211 none were significantly different from that expected by chance (Table 4). For each species, no spatial

212 patterns were discernable in the FCA, as there was little to no clustering of individuals from the same

213 collection site in any species and the distribution of the multi-locus genotype data was not strongly

214 explained by either axis (Fig. 2).

215

8 216 Power analyses suggested that genetic divergence would be detected with ≥ 95% confidence, at least at

217 FST ≥ 0.034 for A. trutta (Ne = 1000, t = 70), FST ≥ 0.039 for A. truttaceus (Ne = 1000, t = 80) and FST

218 ≥ 0.009 for A. georgianus (Ne = 1000, t = 20).

219

220 Discussion

221

222 This research did not reveal any evidence of genetic structuring in any of Arripis trutta, A. truttaceus or

223 A. georgianus in Australian waters. For one or more species, this could potentially be due to: (i) errors

224 in the genetic data; (ii) inadequate sampling; and/or (iii) the effects of selection and/or (iv) inadequate

225 genetic data or (v) no genetic structure. Each of these alternatives is evaluated below. As is explained,

226 when the genetic data are considered along with other evidence, alternative (v) seems particularly

227 likely.

228

229 Errors in the genetic data. There was no evidence of methodological errors, such as PCR artefacts,

230 impacting on the Arripis genetic data sets and regardless most such errors typically produce

231 heterogeneity rather than homogeneity in genetic data (Pompanon et al. 2005). The conservative

232 approach to allele scoring (see Methods) could conceivably have masked some variation through a

233 failure to discriminate alleles of similar but different size. However, the adopted approach was

234 preferable to other methods of allele scoring of fragment length variation at intron loci where, for

235 example, arbitrary size categories have been imposed on the data (e.g. Gomulski et al. 1998; Berrebi et

236 al. 2005; Hubert et al. 2006; Atarhouch et al. 2007).

237

238 Inadequate sampling. The results were based on samples obtained from across virtually the entire

239 Australian distribution of all three study species. Thus, it seems highly unlikely that any genetic

240 differentiation in any of the species in Australian waters was missed due to, for example, sampling all

241 individuals from a single, localised breeding unit within a subdivided species (see Page and Holmes

242 1998).

243

244 Extensive sampling for temporal or cohort effects (see Johnson and Black 1984; Hedgecock 1994) was

245 not undertaken, partly because preliminary analyses revealed no evidence of significant spatial genetic

9 246 heterogeneity in any of the three study species. In addition, limited testing found no evidence that

247 these factors had any effect on genetic variation in A. georgianus (Moore 2012).

248

249 Selection. Although uniform selection can maintain genetic homogeneity across a species’ range (see

250 Ehrlich and Raven 1969; Lamy et al. 2011), it is unlikely to account for the absence of genetic structure

251 at multiple, independent intron loci (see Avise 2000; Hare 2001; but see Amos and Harwood 1998), as

252 was observed in the Arripis species.

253

254 Inadequate genetic data.

255 Resolving the finer details of population genetic structure in high gene-flow situations, such as in the

256 arripids, is difficult (see Shaklee & Currens 2003). Evidence in support of the hypothesis of genetic

257 homogeneity is relatively strong for A. georgianus, where, for example, the results of the power

258 analysis indicate that the current data set could detect modest levels of genetic subdivision (FST ≥

259 0.009). Furthermore, a prior allozyme study, based on four loci, similarly failed to detect any evidence

260 of genetic subdivision in this species (based on 646 individuals sampled from 11 sites between Victoria

261 and Western Australia; Ayvazian et al. 2004), although the levels of polymorphism at the allozyme loci

262 were admittedly very low. The number of polymorphic intron loci assayed for A. trutta and A.

263 truttaceus was too small to provide compelling evidence of a lack of structure for either species.

264 However, in this context, it is worth noting that an unpublished study of variation at the esterase locus

265 revealed no evidence of subdivision in the latter species (based 580 individuals from nine sites between

266 Victoria and Western Australia; MacDonald 1980). There are no prior population genetic studies for A.

267 trutta in Australian waters. Furthermore, all Arripis species, particularly A. truttaceus, exhibit

268 unusually low levels of polymorphism genome wide (Ayvazian et al. 2004; MacDonald 1980; Moore

269 2012), making it unusually difficult to find sufficient polymorphic markers to test a hypothesis of

270 genetic homogeneity with a high degree of power (see Moore 2012).

271

272 No genetic structure.

273 The results of this study are consistent with the hypothesis that the amount of gene flow across the

274 entire and vast Australian distribution of each Arripis species is sufficient to prevent genetic

275 structuring. This was hypothesised in view of the high migratory life-history of these species (see

276 Introduction). While additional genetic data are required to confirm this interpretation (see above), the

10 277 genetic results are consistent with the results of studies based on other types of evidence, specifically

278 otolith microchemistry and tagging, as follows.

279

280 Otolith microchemistry analyses have provided evidence of extensive movements and mixing of

281 individuals in both A. trutta and A. georgianus (Ayvazian et al. 2004; Hughes et al. 2011). Using

282 analysis of δ18O and δ13C from 559 fish across 10 sites between Adelaide (South Australia) and

283 Dongara (Western Australia), Ayvazian et al. (2004) found that, in general, the sampled individuals of

284 A. georgianus had not spent their entire lives in the waters in which they were captured. This is in

285 contrast to a range of other species with similar geographic distributions (e.g. Edmonds and Fletcher

286 1997; Edmonds et al. 1999). Similar conclusions were reached for A. trutta by Hughes et al. (2011)

287 based on 115 individuals sampled across eight sites between northern New South Wales and Tasmania.

288 Elemental signatures (Li, Na, Mg, Mn, Sr, Ba) were strongly influenced by age, reflecting the extensive

289 movements and geographical age stratification of maturing fish. The data from both studies supported

290 the idea that A. trutta comprised a single, well-mixed population. Both studies also reported limited

291 evidence of partial residency in some populations, but not in sexually mature fish. No otolith

292 microchemistry studies have been conducted on A. truttaceus.

293

294 Tagging studies have also demonstrated a progressive age-based movement of maturing fish from

295 nursery areas in the south-east to respective adult distributions along the east coast (A. trutta) and the

296 south-west coast (A. truttaceus and A. georgianus) of Australia. These studies, which were based on

297 hundreds of tagged fish, also showed that adults of each species make multi-directional movements

298 within their adult distributions with no evidence for site-fidelity or a return to nursery areas (e.g.

299 Stanley 1986; 1988a; 1988b; 1988c; Ayvazian et al. 2004). The conclusions from all of these studies

300 were that each of the species represents a single homogenous population.

301

302 The results of this study add to a small list of coastal pelagic species that could be genetically

303 homogeneous and possibly even panmictic over a vast distance (e.g. see Hoolihan et al. 2006; Cárdenas

304 et al. 2009; Palm et al. 2009). It is also noteworthy that the results were effectively the same for the

305 eastern and western sibling-species of Australian Salmon (A. trutta and A. truttaceus, respectively),

306 indicating that environmental conditions on both the east and west coasts of Australia are similarly

307 conducive to high gene flow in these species. Three factors appear to be critical in the high gene flow

11 308 life history of Arripis. Firstly, almost all of the spawning individuals of each species converge on a

309 relatively spatially-restricted area over a short time-frame, which will facilitate random mating (see

310 also White et al. 2009). Secondly, pelagic spawning is closely coincident (and possibly linked) with

311 oceanographic conditions, especially major currents. Specifically, spawning in both A. truttaceus and

312 A. georgianus occurs in south-western Australia, peaking in April and May when the south-east-

313 flowing Leeuwin Current is at its strongest (Caputi et al. 1996), while spawning in A. trutta occurs in

314 eastern Australia, peaking in January when the south-flowing East Australian Current is strongest

315 (Stewart et al. 2011). Thirdly, there is a progressive migration of early life history stages back to the

316 adult distribution, which presumably provides an extensive homogenising effect in all species of

317 Arripis (see also Ward et al. 2003).

318

319 Conclusions.

320 Firmly proving or disproving a hypothesis of genetic homogeneity in a marine species is difficult, as is

321 exemplified by a series of studies of the European Eel (Anguilla anguilla), which have variously

322 proposed and rejected a model of homogeneity (see DeLigny & Pantelouris 1973; Daemen et al. 2001;

323 Wirth & Bernatchez 2001; Maes & Volckaert 2002; Lintas et al. 1998; Dannewitz et al. 2005; Palm et

324 al. 2009; Als et al. 2011). This study provides the strongest evidence to date that, as predicted, each of

325 Arripis trutta, A. truttaceus and A. georgianus is genetically homogenous in Australian waters.

326 Information from additional genetic markers is, however, required to improve the power of the analysis

327 of the population genetic structures of these species.

328

329 Acknowledgements

330

331 We are grateful to E. Sezmiş, N. Philips, F. Brigg and M. Hale for valuable advice on many aspects of

332 this work. We thank S. Hoeksema, B. Chuwen, D. Wheatcroft, K. Armstrong, J. Stewart, J. Hughes

333 and P. & M. Selim (Selim Fisheries) for assistance in sourcing specimens. This work complies with

334 the current laws of Australia and was conducted under Murdoch University Ethics Approval

335 W1199/2006. Funding was provided by Murdoch University and the Centre for Fish and Fisheries

336 Research, Murdoch University.

337

338 References

12 339

340 Als, T. D., Hansen, M. M., Maes, G. E., Castonguay, M., Riemann, L., Aarestrup, K., Munk, P.,

341 Sparholt, H., Hanel, R., Bernatchez, L. (2011). All roads lead to home: panmixia of European

342 eel in the Sargasso Sea. Mol Ecol 20, 1333-1346

343 Amos, W., Harwood, J. (1998). Factors affecting levels of genetic diversity in natural populations. Phil

344 Trans R Soc Lond, Ser B 353, 177-186

345 Andre´, C., Larsson, L. C., Laikre, L., Bekkevold, D., Brigham, J., Carvalho, G. R., Dahlgren, T. G.,

346 Hutchinson, W. F., Mariani, S., Mudde, K., Ruzzante, D. E., Ryman, N. (2011). Detecting

347 population structure in a high gene-flow species, Atlantic herring (Clupea harengus): direct,

348 simultaneous evaluation of neutral vs putatively selected loci. Heredity 106, 270-280

349 Anon. (2004). Microsatellite Analysis on the Applied Biosystems 3130 Series Genetic Analyzers.

350 Applied Biosystems

351 Atarhouch, T., Rami, M., Naciri, M., Dakkak, A. (2007). Genetic population structure of sardine

352 (Sardina pilchardus) off Morocco detected with intron polymorphism (EPIC-PCR). Mar Biol

353 150, 521-528

354 Avise, J. C. (2000). Phylogeography. The History and Formation of Species. Cambridge,

355 Massachusetts: Harvard University Press

356 Ayvazian, S. G., Bastow, T. P., Edmonds, J. S., How, J., Nowara, G. B. (2004). Stock structure of

357 Australian Herring (Arripis georgiana) in southwestern Australia. Fish Res 67, 39-53

358 Barbará, T., Palma-Silva, C., Paggi, G.M., Bered, F., Fay, M.F., Lexer, C. (2007). Cross-species

359 transfer of nuclear microsatellite markers: potential and limitations. Mol Ecol 16, 3759–3767

360 Belkhir, K., Borsa, P., Goudet, J., Chikhi, L., Bonhomme, F. (2004). GENETIX, logiciel sous

361 Windows™ pour la génétiques des populations. Montpellier, France: Laboratoire Génome et

362 Populations, Université Montpellier II

363 Berrebi, P., Boissin, E., Fang, F., Cattaneo-Berrebi, G. (2005). Intron polymorphism (EPIC-PCR)

364 reveals phylogeographic structure of Zacco platypus in China: a possible target for aquaculture

365 development. Heredity 94, 589-598

366 Bierne, N., Lehnert, S. A., Bédier, E., Bonhomme, F., Moore, S. S. (2000). Screening for intron-length

367 polymorphisms in penaeid shrimps using exon-primed intron-crossing (EPIC)-PCR. Mol Ecol

368 9, 233-235

13 369 Boissin, E., Hoareau, T. B., Berrebi, P. (2011). Effects of current and historic fragmentation on the

370 genetic structure of the sand goby Pomatoschistus minutus (Osteichthys, Gobiidae). Biol J Linn

371 Soc 102, 175-198

372 Cappo, M., Walters, C. J., Lenanton, R. J. (2000). Estimation of rates of migration, exploitation and

373 survival using tag recovery data for western Australian "salmon" (Arripis truttaceus: Arripidae:

374 Percoidei). Fish Res 44, 207-217

375 Caputi, N., Fletcher, W. J., Pearce, A., Chubb, C. F. (1996). Effect of the Leeuwin Current on the

376 recruitment of fish and invertebrates along the Western Australian coast. Mar Freshw Res 47,

377 147-155

378 Cárdenas, L., Silva, A. X., Magoulas, A., Cabezas, J., Poulin, E., Ojeda, F. P. (2009). Genetic

379 population structure in the Chilean jack mackerel, Trachurus murphyi (Nichols) across the

380 South-eastern Pacific Ocean. Fish Res 100, 109-115

381 Chairi, H., Idaomar, M., Rebordinos, L. (2007). Mitochondrial DNA analysis of the European Anchovy

382 in the southern Mediterranean and northern Atlantic coasts. J Fish Aq Sc 2, 206-215

383 Chlaida, M., Laurent, V., Kifani, S., Benazzou, T., Jaziri, H., Planes, S. (2008). Evidence of a genetic

384 cline for Sardina pilchardus along the Northwest African coast. ICES J Mar Sci 66, 264-271

385 Chow, S., Hazama, K. (1998). Universal PCR primers for S7 ribosomal protein gene introns in fish.

386 Mol Ecol 7, 1247-1263

387 Chow, S., Takeyama, H. (1998). Intron length variation observed in the creatine kinase and ribosomal

388 protein genes of the Swordfish Xiphias gladius. Fish Sc 64, 397-402

389 Chow, S., Takeyama, H. (2000). Nuclear and mitochondrial DNA analyses reveal four genetically

390 separated breeding units of the swordfish. J Fish Biol 56, 1087-1098

391 Daemen, E., Cross, T., Ollevier, F., Volckaert, F. A. M. (2001). Analysis of the genetic structure of

392 European eel (Anguilla anguilla) using microsatellite DNA and mtDNA markers. Mar Biol 139,

393 755-764

394 Dannewitz, J., Maes, G. E., Johansson, L., Wickström, H., Volckaert, F. A. M., Järvi, T. (2005).

395 Panmixia in the European eel: a matter of time... Proc Roy Soc B 272, 1129-1137

396 DeLigny, W., Pantelouris, E. M. (1973). Origin of the European eel. Nature 246, 518-519

397 Epicentre (2003). MasterPureTM DNA Purification Kit: Instructions for Use

398 Ehrlich, P. R., Raven, P. H. (1969). Differentiation of populations. Science 165: 1228-1232

14 399 Fairbridge, W. S. (1950). Some populations of the Australian "Salmon". In Indo-Pacific Fisheries

400 Council Proceedings, pp. 80-85. Cronulla, NSW

401 Fairclough, D. V., Dimmlich, W. F., Potter, I. C. (2000a). Reproductive biology of the Australian

402 herring Arripis georgiana. Mar Freshw Res 51, 619-630

403 Fairclough, D. V., Dimmlich, W. F., Potter, I. C. (2000b). Length and age compositions and growth

404 rates of the Australian herring Arripis georgiana in different regions. Mar Freshw Res 51, 631-

405 640

406 Fauvelot, C., Lemaire, C., Planes, S., Bonhomme, F. (2007). Inferring gene flow in coral reef fishes

407 from different molecular markers: which loci to trust? Heredity 99, 331-339

408 Gillanders, B. M., Able, K. W., Brown, J. A., Egglestone, D. B., Sheridan, P. F. (2003). Evidence of

409 connectivity between juvenile and adult habitats for mobile marine fauna: an important

410 component of nurseries. Mar Ecol Prog Ser 247, 281-295

411 Gold, J. R., Pak, E., DeVries, D. A. (2002). Population structure of king mackerel (Scomberomorus

412 cavalla) around peninsular Florida, as revealed by microsatellite DNA. Fish Bull 100, 491-509

413 Gomulski, L. M., Bourtzis, K., Brogna, S., Morandi, P. A., Bonvicini, C., Sebastiani, F., Torti, C.,

414 Guglielmino, C. R., Savakis, C., Gasperi, G., Malacrida, A. R. (1998). Intron size

415 polymorphism of the Adh1 gene parallels the worldwide colonization history of the

416 Mediterranean fruit fly, Ceratitis capitata. Mol Ecol 7, 1729-1741

417 Goudet, J. (1995). FSTAT (Version 1.2): A computer program to calculate F-statistics. Heredity 86,

418 485-486

419 Guinand, B. (1996). Use of a multivariate model using allele frequency distributions to analyse patterns

420 of genetic differentiation among populations. Biol J Linn Soc 58, 173-195

421 Haney, R. A., Silliman, B. R., Rand, D. M. (2007). A multi-locus assessment of connectivity and

422 historical demography in the bluehead wrasse (Thalassoma bifasciatum). Heredity 98, 294-302

423 Hare, M. P. (2001). Prospects for nuclear gene phylogeography. Trends Ecol and Evol 16, 700-706

424 Hassan, M., Lemaire, C., Fauvelot, C., Bonhomme, F. (2002). Seventeen new exon-primed intron-

425 crossing polymerase chain reaction amplifiable introns in fish. Mol Ecol Notes 2, 334-340

426 Hauser, L., Ward, R. (1998). Population identification in pelagic fish: the limits of molecular markers.

427 In Advances in Molecular Ecology (Carvalho, G. R., ed.), pp. 191-224. Amsterdam: IOS Press

15 428 Hedgecock, D. (1994). Does variance in reproductive success limit effective population sizes of marine

429 organisms? In Genetics and evolution of aquatic organisms (Beaumont, A. R., ed.), pp. 122–

430 134: Chapman and Hall, London

431 Hoedt, F. E., Dimmlich, W. F. (1994). Diet of subadult Australian Salmon, Arripis truttaceus, in

432 Western Port, Victoria. Aust J Mar Freshw Res 45, 617-623

433 Hoolihan, J. P., Anandh, P., van Herwerden, L. (2006). Mitochondrial DNA analyses of narrow-barred

434 Spanish mackerel (Scomberomorus commerson) suggest a single genetic stock in the ROPME

435 sea area (Arabian Gulf, Gulf of Oman, and Arabian Sea). ICES J Mar Sci 63, 1066-1074

436 Hubert, N., Duponchelle, F., Nuñez, J., Rivera, R., Renno, J.-F. (2006). Evidence of reproductive

437 isolation among closely related species of Serrasalmus (Ostariophysii, Characidae) from the

438 Upper Madeira River, Amazon, Bolivia. J Fish Biol 69, 31-51

439 Hughes, J. M., Stewart, J., Gillanders, B. M., Suthers, I. M. (2011). An examination of the population

440 structure of the eastern Australian salmon (Arripis trutta) stock in south-eastern Australia.

441 Proceedings of the International Symposium on Advances in Fish Tagging and Marking

442 Technology. February 2008, Auckland, New Zealand. Trans Am Fish Soc 76

443 Idury, R. M., Cardon, L. R. (1997). A simple method for automated allele binning in microsatellite

444 markers. Genome Res 7, 1104-1107

445 Johnson, G. D., Fritzsche, R. A. (1989). Graus nigra, an omnivorous girellid, with a comparative

446 osteology and comments on relationships of the Girellidae (Pisces: ). Proc Acad Nat

447 Sc Philadelphia 141: 1-27

448 Johnson, M. S., Black, R. (1984). Pattern beneath the chaos: the effect of recruitment on genetic

449 patchiness in an intertidal limpet. Evolution 38, 1371-1383

450 Knutsen, H., Jorde, P. E., André, C., Stenseth, N. C. (2003). Fine-scaled geographical population

451 structuring in a highly mobile marine species: the Atlantic cod. Mol Ecol 12, 385-394

452 Lamy, J-B., Bouffier, L., Burlett, R., Plomion, C., Cochard, H., Delzon, S. (2011). Uniform selection as

453 a primary force reducing population genetic differentiation of cavitation resistance across a

454 species range. PLoS ONE 6(8): e23476

455 Lenanton, R. (1982). Alternative non-estuarine nursery habitats for some commercially and

456 recreationally important fish species of south-western Australia. Aust J Mar Freshw Res 33,

457 881-900

16 458 Lenanton, R. C. J., Hall, N. G. (1976). The Western Australian amateur fishery for Australian Herring

459 (Arripis georgianus): results of the 1973 creel census. 59 pp.: Department of Fisheries and

460 Wildlife, Western Australia

461 Li, C., Riethoven, J-J. M., Lingbo, M. (2010). Exon-primed intron-crossing (EPIC) markers for non-

462 model teleost fishes. BMC Evol. Biol. 10, 90

463 Li, Y.-C., Korol, A. B., Fahima, T., Beiles, A., Nevo, E. (2002). Microsatellites: genomic distribution,

464 putative functions and mutational mechanisms: a review. Mol Ecol 11, 2453-2465

465 Lintas, C., Hirano, J., Archer, S. (1998). Genetic variation of the European eel (Anguilla anguilla). Mol

466 Mar Biol Biotech 7, 263-269

467 MacDonald, C. M. (1980). Population structure, biochemical adaptation and systematics in temperate

468 marine fishes of the genera Arripis and Chrysophrys (Pisces: Perciformes). Dissertation, 336pp,

469 Australian National University, Canberra

470 Maes, G. E., Volckaert, F. A. M. (2002). Clinal genetic variation and isolation by distance in the

471 European eel Anguilla anguilla (L.). Biol J Linn Soc 77, 509-521

472 Malcolm, W. B. (1960). Area of distribution, and movement of the western subspecies of the

473 Australian "Salmon", Arripis trutta esper Whitley. Aust J Mar Freshw Res 11, 282-325

474 Moore, G. I. (2012). Aspects of the evolutionary history of a pair of fish species (Arripidae: Arripis) on

475 either side of a biogeographic barrier in southern Australian seas. Dissertation, 237pp. Centre

476 for Fish and Fisheries Research, School of Biological Sciences and Biotechnology, Murdoch

477 University, Western Australia

478 Nei, M. (1987). Molecular Evolutionary Genetics. New York: Columbia University Press

479 Nicholls, A. G. (1971). A monitoring programme for Australian "Salmon" in Western Australia. 40 pp.:

480 CSIRO Division of Fisheries and Oceanography, Cronulla, NSW

481 Page, R. D. M., Holmes, E. C. (1998). Molecular Evolution. A Phylogenetic Approach. Oxford:

482 Blackwell Science

483 Palm, S., Dannewitz, J., Prestegaard, T., Wickström, H. (2009). Panmixia in European eel revisited: no

484 genetic difference between maturing adults from southern and northern Europe. Heredity 103,

485 82-89

486 Palumbi, S. R., Baker, C. S. (1994). Contrasting population structure from nuclear intron sequences and

487 mtDNA of Humpback Whales. Mol Biol Evol 11, 426-435

17 488 Petrusevics, P., Bye, J. A. T. (1995). The Leeuwin Current, wind, mean sea level and variability in

489 Australian salmon larvae recruitment. In Proc Ocean Atm Pac Intl Conf (Aung, T. H., ed.), pp.

490 288-292. Adelaide, South Australia

491 Pompanon, F., Bonin, A., Bellemain, E., Taberlet, P. (2005). Genotyping errors: causes, consequences

492 and solutions. Nat Rev Genet 6, 847-859

493 Raymond, M., Rousset, F. (1995). GENEPOP: population genetics software for exact tests and

494 ecumenicism. (vers. 4.0). J Heredity 86, 248-249

495 Rice, W. R. (1989). Analyzing tables of statistical tests. Evolution 43, 223-225

496 Rohfritsch, A., Borsa, P. (2005). Genetic structure of Indian scad mackerel Decapterus russelli:

497 Pleistocene vicariance and secondary contact in the central Indo-West Pacific seas. Heredity 95,

498 315-322

499 Rolland, J.-L., Bonhomme, F., Lagardère, F., Hassan, M., Guinand, B. (2007). Population structure of

500 the common sole (Solea solea) in the Northeastern Atlantic and the Mediterranean Sea:

501 revisiting the divide with EPIC markers. Mar Biol 151, 327-341

502 Ruzzante, D. E., Mariani, S., Bekkevold, D., Andre, C., Mosegaard, H., Clausen, L. A. W., Dahlgren,

503 T. G., Hutchinson, W. F., Hatfield, E. M. C., Torstensen, E., Brigham, J., Simmonds, E. J.,

504 Laikre, L., Larsson, L. C., Stet, R. J. M., Ryman, N., Carvalho, G. R. (2006). Biocomplexity in

505 a highly migratory pelagic marine fish, Atlantic herring. Proc Roy Soc B 273, 1459-1464

506 Ryman, N., Palm, S. (2006). POWSIM: a computer program for assessing statistical power when

507 testing for genetic differentiation. Mol Ecol 6, 600-602

508 Sale, P. F., Cowen, R. K., Danilowicz, B. S., Jones, G. P., Kritzer, J. P., Lindeman, K. C., Planes, S.,

509 Polunin, N. V. C., Russ, G. R., Sadovy, Y. J., Steneck, R. S. (2005). Critical science gaps

510 impede use of no-take fishery reserves. Trends Ecol and Evol 20, 74-80

511 Selkoe, K. A., Toonen, R. J. (2006). Microsatellites for ecologists: a practical guide to using and

512 evaluating microsatellite markers. Ecology Letters 9, 615-629

513 Shaklee, J. B., Currens, K. P. (2003). Genetic stock identification and risk assessment. In Population

514 Genetics: Principles and Applications for Fisheries Scientists (Hallerman, E. M., ed.), pp. 291-

515 328. Bethesda: American Fisheries Society

516 Siegel, D., Kinlan, B. P., Gaylord, B., Gaines, S. D. (2003). Lagrangian descriptions of marine larval

517 dispersion. Mar Ecol Prog Ser 260, 83-96

18 518 Stanley, C. A. (1986). Tagging experiments on Australian Salmon (Arripis trutta esper): Recapture

519 data for South Australian releases, 1952 to 1976. CSIRO Marine Laboratories, Hobart,

520 Tasmania

521 Stanley, C. A. (1988a). Tagging experiments on Australian Salmon (Arripis trutta): Recapture data for

522 New South Wales releases, 1939 to 1965, and Victorian releases, 1939 to 1964. CSIRO Marine

523 Laboratories, Hobart, Tasmania, No. 197

524 Stanley, C. A. (1988b). Tagging experiments on Australian Salmon (Arripis trutta esper): Recapture

525 data for Western Australian releases, 1945 to 1976. CSIRO Marine Laboratories, Hobart,

526 Tasmania, No. 196

527 Stanley, C. A. (1988c). Tagging experiments on Australian Salmon (Arripis trutta): Recapture data for

528 Tasmanian releases, 1949 to 1964. CSIRO Marine Laboratories, Hobart, Tasmania, No. 193

529 Stanley, C. A., Malcolm, W. B. (1977). Reproductive cycles in the eastern subspecies of the Australian

530 Salmon, Arripis trutta marginata (Cuvier, Valenciennes). Aust J Mar Freshw Res 28, 287-301

531 Stewart, J., Hughes, J. M., McAllister, J., Lyle, J. M., MacDonald, C. M. (2011). Australian salmon

532 (Arripis trutta): Population structure, reproduction, diet and composition of commercial and

533 recreational catches. 259 pp. Cronulla: Investment and Industry NSW, Cronulla

534 Theisen, T. C., Bowen, B. W., Lanier, W., Baldwin, J. D. (2008). High connectivity on a global scale in

535 the pelagic wahoo, Acanthocybium solandri (tuna family Scombridae). Mol Ecol 17, 4233-4247

536 van Oosterhout, C., Hutchinson, W. F., Wills, D. P. M., Shipley, P. (2004). MICRO-CHECKER:

537 software for identifying and correcting genotyping errors in microsatellite data. Mol Ecol Notes

538 4, 535-538

539 Wang, M., Zhang, X., Yang, T., Han, Z., Yanagimoto, T., Gao, T. (2008). Genetic diversity in the

540 mtDNA control region and population structure in the Sardinella zunasi Bleeker. Afr J Biotech

541 7, 4384-4392

542 Waples, R. S. (1987). A multispecies approach to the analysis of gene flow in marine shore fishes.

543 Evolution 41, 385-400

544 Ward, R. D., Woodwark, M., Skibinski, D. O. F. (1994). A comparison of genetic diversity levels in

545 marine, freshwater, and anadromous fishes. J Fish Biol 44, 213-232

546 Ward, T. M., Staunton-Smith, J., Hoyle, S., Halliday, I. A. (2003). Spawning patterns of four species of

547 predominantly temperate pelagic fishes in the sub-tropical waters of southern Queensland. Est

548 Coast Shelf Sci 56, 1125-1140

19 549 White, T. A., Stefanni, S., Stamford, J., Hoelzel, A. R. (2009). Unexpected panmixia in a long-lived,

550 deep-sea fish with well-defined spawning habitat and relatively low fecundity. Mol Ecol 18,

551 2563-2573

552 Wirth, T., Bernatchez, L. (2001). Genetic evidence against panmixia in the European eel. Nature 409,

553 1037-1040

554 Yagishita, N., Miya, M., Yamanoue, Y., Shirai, S. M., Nakayama, K., Suzuki, N., Satoh, T. P.,

555 Mabuchi, K., Nishida, M., Nakabo, T. (2009). Mitogenomic evaluation of the unique facial

556 nerve pattern as a phylogenetic marker within the perciform fishes (Teleostei: Percomorpha).

557 Mol Phyl Evol 53: 258-266

20 558 Fig. 1 Sampling sites and generalised map of the distribution and movements of Arripis trutta

559 (dashed arrows), A. truttaceus (black arrows), and A. georgianus (grey arrows) in Australian waters.

560 ‘I’ indicates presumed larval dispersal, including some local retention; ‘II’ represents direction of

561 movement of juvenile fish; ‘III’ represents movements of sub-adult fish; and ‘IV’ represents adult

562 distribution and movements. Sample sites are indicated with a species prefix: A. trutta (E), A.

563 truttaceus (W) and, A. georgianus (H), and location suffix: Fremantle (F), Canal Rocks (C), Augusta

564 (A), Yeagarup (Y), Peaceful Bay (P), Irwin Inlet (I), Oyster Harbour (O), Albany (L), Wellstead

565 Estuary (W), in Western Australia; The Coorong, South Australia (S); Flinders Island, Tasmania (T);

566 Lakes Entrance, Victoria (V); Port Stephens, New South Wales (N) 567

21

568 Fig. 2 Factorial correspondence analysis (FCA) illustrating the lack of a strong relationship

569 between the multi-locus genotypes and collection sites of individuals of Arripis trutta (top), A.

570 truttaceus (middle) and A. georgianus (bottom). The total variation explained by each axis is given as

571 a percentage. Sample codes are as per Fig. 1. For scaling purposes, outliers have been shifted and their

572 correct position indicated by an arrow and coordinates. Note, many points overlap and are obscured 573

22

574 Table 1 PCR and plate loading conditions for assays for polymorphic intron loci in the three

575 species of Arripis. In each case, the optimised annealing temperature and primer concentration used in

576 PCR reactions is given, along with the volume of PCR product loaded onto plates for fragment

577 analysis. A dash (-) indicates that the locus was monomorphic for a species, and not included in the

578 present study. An asterisk (*) indicates touchdown PCR

579

Arripis trutta Arripis truttaceus Arripis georgianus Anneal. Primer Load Anneal. Primer Load Anneal. Primer Load Locus temp. conc. vol. temp. conc. vol. temp. conc. vol. (°C) (µM) (µl) (°C) (µM) (µl) (°C) (µM) (µl)

AldoB4 56 0.04 2.0 56 0.04 2.0 57 0.04 0.5

Cam4 - - 65-45* 0.04 1.0

CK7 53 0.06 3.5 - 58-41* 0.04 2.0

S72 62 0.06 2.0 63 0.06 2.5 64 0.06 2.5 580

23 581 Table 2 Characteristics of four intron loci for samples of N individuals of three species of Arripis.

582 NA is the number of alleles for each sample and for the species (in parentheses). The range of allele

583 sizes for each locus is given for each species (in bp). HE and HO are expected and observed

584 heterozygosity, respectively. HWE(p) is the probability of a Type 1 error in exact tests for departures

585 from Hardy-Weinberg equilibrium expectations. Collection sites are ordered from the most northerly

586 to the most south-easterly (see Fig. 1). A dash (-) indicates that the locus was monomorphic for a

587 sample. An asterisk (*) indicates not significant after Bonferroni correction. ‘n/a’ indicates that a p-

588 value could not be calculated because only 2 alleles were present in the sample, one of which was

589 represented by a single copy

24 Arripis trutta Arripis truttaceus Arripis georgianus EN EV ET WA WY WL WW WS HF HC HP HI HO HW HS AldoB4 (170-191 bp) (178-182 bp) (170-182 bp) N 29 30 30 30 15 39 9 37 44 39 36 23 28 9 41

N 5 3 3 (6) 2 2 2 2 2 (2) 5 2 4 4 3 3 4 (5) A HE 0.419 0.347 0.326 0.480 0.391 0.474 0.345 0.382 0.316 0.120 0.305 0.361 0.194 0.291 0.159

HO 0.414 0.433 0.333 0.667 0.400 0.513 0.222 0.405 0.295 0.128 0.250 0.348 0.214 0.333 0.171 HWE(p) 0.33 0.45 1.00 0.06 1.00 0.74 0.34 1.00 0.50 1.00 0.43 0.71 1.00 1.00 1.00

CaM4 (477 bp) (477 bp) (469-504 bp) N monomorphic monomorphic 42 39 36 23 28 8 40

NA 5 5 3 3 3 2 6 (8)

HE 0.177 0.146 0.107 0.083 0.223 0.116 0.228

HO 0.190 0.154 0.111 0.087 0.179 0.125 0.200 HWE(p) 1.00 1.00 1.00 1.00 0.35 n/a 0.48

CK7 (943-984 bp) (952 bp) (844-976 bp) N 28 28 29 monomorphic 42 38 35 23 26 8 39

NA 3 4 3 (4) 6 7 4 3 2 1 6 (11)

HE 0.283 0.197 0.273 0.278 0.174 0.209 0.198 0.142 - 0.279

HO 0.321 0.214 0.321 0.214 0.184 0.229 0.217 0.154 - 0.231 HWE(p) 1.00 1.00 1.00 0.20 1.00 1.00 1.00 1.00 - 0.13

S72 (816-931 bp) (892-931 bp) (871-919 bp) N 29 25 28 27 15 38 13 45 32 36 34 23 26 8 36

NA 6 7 4 (9) 4 3 4 4 5 (5) 6 5 4 3 3 3 5 (8)

HE 0.655 0.594 0.535 0.444 0.558 0.465 0.426 0.579 0.466 0.450 0.286 0.294 0.272 0.531 0.336

HO 0.552 0.480 0.679 0.333 0.467 0.474 0.462 0.667 0.406 0.528 0.324 0.261 0.154 0.750 0.333 HWE(p) 0.31 0.31 0.51 0.12 0.56 1.00 0.43 0.24 0.44 1.00 1.00 0.52 0.01* 0.70 0.63

25 590 Table 3 Single locus and multilocus tests of population differentiation in each Arripis species.

591 FST values are given along with the probability of obtaining, by chance, a value as large as or larger

592 than the observed FST from 10,000 permutations. A dash (-) indicates that the locus was monomorphic

593 for a species

594

Arripis trutta Arripis truttaceus Arripis georgianus FST p-value FST p-value FST p-value (i) Single-locus AldoB4 -0.010 0.34 0.011 0.23 0.007 0.14 CaM4 - - - - -0.004 0.98 CK7 -0.006 0.52 - - -0.009 0.26 S72 -0.002 0.63 0.001 0.42 0.005 0.48

(ii) Overall -0.005 0.73 0.005 0.22 0.005 0.10 595

26 596 Table 4 Pairwise FST (below diagonal) and associated p-values (above diagonal) for samples of

597 Arripis trutta, A. truttaceus and A. georgianus. p-values are the probability of obtaining, by chance, a

598 value as large as or larger than the observed FST from 10,000 permutations. Sample codes are as per

599 Fig. 1

600

Arripis trutta EN EV ET EN 0.27 0.32 EV -0.004 0.97 ET -0.000 -0.013

Arripis truttaceus WA WY WL WW WS WA 0.40 0.83 0.49 0.25 WY 0.009 0.11 0.22 0.79 WL -0.011 0.012 0.76 0.08 WW 0.008 -0.011 0.000 0.32 WS 0.018 -0.019 0.018 -0.006

Arripis georgianus HF HC HP HI HO HW HS HF 0.35 0.90 0.34 0.37 0.87 0.72 HC 0.003 0.39 0.14 0.38 0.61 0.90 HP 0.005 0.012 0.92 0.68 0.24 0.73 HI 0.010 0.019 -0.013 0.30 0.07 0.28 HO 0.010 0.005 -0.005 -0.000 0.19 0.54 HW -0.016 0.000 0.029 0.030 0.026 0.42 HS 0.004 -0.002 0.002 0.009 -0.007 0.018 601

27