52
views
0
recommends
+1 Recommend
0 collections
    0
    shares
      • Record: found
      • Abstract: found
      • Article: not found
      Is Open Access

      Evolutionary origins and genetic variation of the Seychelles treefrog, Tachycnemis seychellensis (Duméril and Bibron, 1841) (Amphibia: Anura: Hyperoliidae)

      dna, frogs, heterixalus, indian ocean, phylogeography

      Read this article at

      ScienceOpenPublisherPMC
      Bookmark
          There is no author summary for this article yet. Authors can add summaries to their articles on ScienceOpen to make them more accessible to a non-specialist audience.

          Abstract

          1 Introduction Due to their isolation from potential confounding factors, remote islands have long been considered to provide important arenas for investigating evolution (Darwin, 1859). Most evolutionary studies of island biotas have focused on geologically recent volcanic island groups that have never been in contact with a large, ancient mainland, for example the Galápagos and Hawaii (e.g. Darwin, 1859; Gillespie, 2002). In contrast, the Seychelles archipelago (1600 km east of the nearest continental landmass) is formed of both granitic and coralline islands. The current granitic Seychelles are the remaining emergent part of a continental fragment, previously part of Gondwana, that was associated with India and Madagascar when they separated from Africa during the Cretaceous. At least some of the granitic Seychelles have always had some emergent land since the breakup of the Gondwanan supercontinent. Much of the continental Seychelles is currently submerged at an average depth of 55 m below sea level, forming the microcontinent ‘Seychellea’, comprising a total area of 129,650 km2 (Davies and Francis, 1964). During times of lowest stands in sea level (see Miller et al., 2005) all of the currently emergent granitic Seychelles would have been in contact. Fluctuations in sea level likely caused many episodes of dis- and reconnection among Seychelles islands, the most recent of which were within the last 10 ka (Colonna et al., 1996; Rohling et al., 1998; Siddall et al., 2003; Camoin et al., 2004; Miller et al., 2005). These fluctuations can be expected to have had a substantial impact on the amounts and spatial patterns of genetic variation of the resident biota. It is unsurprising that most remote islands lack an endemic amphibian fauna given that the osmotic properties of amphibian skin (e.g. Balinsky, 1981; Duellman and Trueb, 1986) likely reduce their dispersal capabilities over saltwater substantially (Bossuyt and Milinkovitch, 2001; Darwin, 1859; Myers, 1953; Nussbaum, 1984). The Seychelles, however, has an amphibian fauna (∼12 species) comprising both frogs (Anura) and caecilians (Gymnophiona) that are restricted solely to the granitic islands (Nussbaum, 1984). Except for the widespread frog Ptychadena mascareniensis (Duméril and Bibron, 1841; Vences et al., 2004), all Seychelles amphibians are endemic. The endemic Seychelles frogs are confined to the four largest granitic islands of Mahé, Silhouette, Praslin and La Digue (Nussbaum and Wu, 1995, 2007; Taylor et al., 2012). Except for a preliminary study of sooglossid frogs (Taylor et al., 2012) molecular analyses have not yet been conducted to determine patterns of genetic variation among and within populations on different islands. Molecular techniques have been applied to several other Seychelles organisms, and substantial spatial structuring and deep genetic splits have been revealed, indicating the presence of cryptic lineages within several currently recognised lizard species (Rocha et al., 2010a,b, 2011, 2013; Valente et al., 2014), and a freshwater crab (Daniels, 2011). Hyperoliidae is a pan-African family comprising >200 species in 17 or 18 genera (AmphibiaWeb, 2014; Frost, 2014) of small-medium sized treefrogs with representative species also found in the Seychelles and Madagascar (Vences et al., 2003a,b). The endemic Seychelles treefrog Tachycnemis seychellensis, the only species of its genus, is a sexually dimorphic, hyperoliid frog found on all four of the granitic islands of the Seychelles that support populations of frogs (Nussbaum and Wu, 1995). Like all hyperoliids T. seychellensis is an oviparous species with an aquatic larval stage, and it is restricted to areas close to water bodies (Nussbaum, 1984). The abundance and type of T. seychellensis habitat varies considerably across its range (Nussbaum and Wu, 1995). The southern islands of Mahé and Silhouette are higher (up to 905 and 750 m elevation, respectively), wetter and dominated by moist–wet forests, whereas the northern islands of Praslin (up to 367 m) and La Digue (333 m) are much lower and drier. Praslin has multiple rivers and streams, but La Digue lacks constant water sources at higher altitudes, and instead T. seychellensis is restricted here to marshy areas in the low-lying plateau on the west of this small island. The sizes of the four islands vary by more than an order of magnitude, ranging from 960 ha (La Digue) to 14,480 ha (Mahé), with Silhouette (1600 ha) and Praslin (4040 ha) somewhat intermediate. Using univariate and multivariate analyses, Nussbaum and Wu (1995) discovered substantial external morphological variation among five populations of T. seychellensis from the four islands, including in adult body size and colouration, presence or absence of tubercles on various parts of the body and limbs, presence or absence of grooved digit discs, and several morphometric characters. Four morphometric characters, not dependant on sex, were found to vary significantly between all populations: internarial width, pes length, toe disc length, and length of metatarsal tubercle. An additional 10 male and two female characters varied significantly. Specimens from the more southerly islands of Mahé and Silhouette are morphologically the most similar to each other (Nussbaum and Wu, 1995). However, within Mahé (the only island for which more than one population was sampled), two populations of T. seychellensis (one marsh- and one stream-associated) only 1 km apart were as morphologically different from each other as they were to T. seychellensis on Silhouette. The populations of T. seychellensis on the more northerly islands (Praslin and La Digue) were morphometrically as distinct from each other as they were from the southern populations. Despite these large morphological differences, Nussbaum and Wu (1995) were impressed by (1) the fact that the four islands were likely connected as recently as 10 ka, (2) the intra-Mahé differences were as large as inter-island differences, (3) the substantial environmental differences across the four islands, and (4) the similar life history and bioacoustics of the different populations, and thus argued for the recognition of only a single species, one that has substantial and geographically structured morphological variation. Nussbaum & Wu’s single-species hypothesis for T. seychellensis could be challenged by high genetic diversity and/or substantial phylogeographic structure. Tachycnemis seychellensis has a complicated taxonomic history. Since Dubois (1981) the species has been included in the monotypic genus Tachycnemis Fitzinger, 1843 and the species name has been attributed to Duméril and Bibron (1841) with Tschudi’s (1838) first use of the species name considered unavailable. Fitzinger (1843) established Tachycnemis only through bibliographic reference to its single included species (as described by Tschudi, 1838) without any explanation of his biological reasons (if any) for proposing the new genus. However, it has long been considered a phenotypically rather distinct hyperoliid (e.g. Günther, 1869), and Drewes (1984) hypothesised that it is the sister group of all other extant hyperoliids. More recently, based on analysis of concatenated mitochondrial DNA (mtDNA) and nuclear DNA (nuDNA), T. seychellensis has been inferred to be most closely related to the endemic Madagascan genus Heterixalus Laurent, 1944, which has 11 currently recognised species (Frost et al., 2006; Pyron and Wiens, 2011; Richards and Moore, 1996; Vences et al., 2003a,b; Wollenberg et al. 2007). However, although Wollenberg et al.’s (2007) main analysis of concatenated data recovered Tachycnemis and Heterixalus as sister taxa, five out of six trees inferred for the individual genes placed Tachycnemis within Heterixalus, although only cox1 (mtDNA) and rho (nuDNA) did so with much support. Paraphyly of Heterixalus with respect to Tachycnemis was also found (though without strong support) by Vences et al. (2003b) in two mitochondrial gene trees. In contrast, these authors found that when three genes were concatenated, but using only two Heterixalus species, Tachycnemis and Heterixalus were sister taxa. Using multiple Heterixalus species in their analyses, Richards and Moore (1996), Frost et al. (2006), and Pyron and Wiens (2011) also recovered a Tachycnemis–Heterixalus sister-group relationship. Where trees for individual loci are discordant, coalescence-based methods can be expected to yield more accurate species phylogenies than multilocus concatenation (e.g. Edwards et al., 2007; Heled and Drummond, 2010; Kubatko and Degnan, 2007; Maddison and Knowles, 2006), but this latter approach has yet to be implemented in the case of Tachycnemis and Heterixalus. Based on a sister-group relationship with Heterixalus and molecular dating analyses, the presence of T. seychellensis in the Seychelles is considered to originate from an overseas dispersal, with Tachycnemis diverging from its closest African/Madagascan relative an estimated 9.79–35.34 Ma (Crottini et al., 2012). This is in contrast to the sooglossid frogs that, as with the Seychelles caecilians, have probably been resident at least since Seychellea (the Seychelles microcontinent) was last part of Gondwana (Nussbaum, 1984). Here we report phylogenetic analyses of mtDNA and nuDNA data (3228 base pairs (bp)) to (1) test the hypothesised sister-group relationship between Heterixalus and Tachycnemis and monophyly of the former genus, and (2) assess genetic variation within T. seychellensis across its range and test the hypothesis that it is a single, morphologically highly variable species. 2 Methods 2.1 Taxon sampling Tachycnemis seychellensis tissue samples (liver, heart and muscle, frozen and stored at −80 °C) were obtained from 52 voucher specimens from the Seychelles islands of Mahé (15 samples), Silhouette (15 samples), Praslin (15 samples) and La Digue (7 samples) between 1988 and 1991; these correspond to four of the five populations sampled by Nussbaum and Wu (1995) (tissues of only a single Mahé population from Mare aux Cochons were available). Vouchers and tissues are deposited in the University of Michigan Museum of Zoology, USA (UMMZ) (see Appendix for details). 2.2 Laboratory protocols Genomic DNA was extracted from liver, heart and muscle samples from the 52 T. seychellensis samples using the Qiagen DNeasy™ Tissue Kit. Three mitochondrial gene fragments were sequenced for all samples: cytochrome b (cytb), cytochrome oxidase subunit 1 (cox1) and 16S rRNA (16s). Four nuclear loci were also sequenced: rhodopsin exon 1 (rho), recombination activating gene 1 (rag1), tyronsinase precursor (tyr) and pro-opiomelanocortin (pomc). The rag1 and tyr sequences showed no variation within T. seychellensis, and thus only a subset of individuals from each locality were included in the analyses of the relationships between Tachycnemis and Heterixalus. Primer information is given in Table 1. Sequences were amplified using the polymerase chain reaction (PCR) with a total reaction volume of 15 μl: 1.5 μl of Bioline Buffer, 0.75 μl of MgCl2, 0.15 μl of dNTPs, 0.15 μl of Taq, 06 μl of both the forward and reverse primers, 0.6 μl of template DNA, and 10.65 μl ddH2O. Cycling conditions were: denature at 94 °C for 60s; followed by 35 (16s, cytb) or 40 (cox1, tyr, pomc, rag1) cycles of denaturing at 94 °C for 30 s, annealing at 48 °C (cox1), 50 °C (16s), 52 °C (cytb), 56 °C (rag1), 60 °C (rho), or 62 °C (tyr, pomc) for 30 s, and extending at 72 °C for 30 s; and a final extending step of 72 °C for 5 min. 2.3 Genetic variation within T. seychellensis Sequences were proof-read using Sequencher v.4.8 and initially aligned using ClustalX v.2.0 (Larkin et al., 2007) using default settings before being checked by eye. All genes except the non-protein-coding 16s were checked for pseudogenes and insertions by searching for stop codons and indels (e.g. Zhang and Hewitt, 1996) in MEGA5 (Tamura et al., 2011). The program DAMBE (Xia and Xie, 2001) was used to test for saturation using the test of Xia et al. (2003) across the different codon positions and the combined dataset. To infer the phylogenetic relationships within T. seychellensis for the mitochondrial locus, we used Bayesian inference (BI) implemented in BEAST v.1.7.4 (Drummond et al., 2012). No outgroup taxa were used because BEAST estimates the position of the root in the tree assuming a molecular clock (Heled and Drummond, 2010). Input XML files were generated for BEAST analyses using BEAUti v.1.7.4. We selected best partitioning strategies and BEAST-compatible substitution models using PartionFinder (Lanfear et al., 2012). The coalescent tree prior with exponential growth was used in BEAST based on the assumption that, after an initial colonisation, T. seychellensis likely expanded its range. Following the results of initial runs, uncorrelated relaxed clocks were rejected for all partitions and a strict clock implemented because constant rates could not be rejected. Two MCMC chains were run for 1 x 108 generations for each partitioning strategy, with trees sampled every 10,000 generations to ensure convergence; this gave a total of 10,000 output trees per run. Convergence was checked by manual observation of the trace plots and ESS scores using Tracer v.1.5 (Rambaut and Drummond, 2009). All BEAST analyses were performed using the CIPRES Science Gateway v.3.1 (Miller et al., 2010). To infer allelic phases from polymorphic sites in the nuDNA, the program PHASE v.2.1 (Stephens et al., 2001; Stephens and Scheet, 2005) was used, with input files created using seqPHASE (Flot, 2010). Haplotype networks under the median-joining algorithm (Bandelt et al., 1999) were produced to display intraspecific variation for T. seychellensis for the pomc and rho loci using the program NETWORK v.4.611 (fluxus-engineering.com). Tajima’s D (Tajima, 1989) and Fu’s Fs (Fu, 1997) neutrality tests were used to investigate historical demographic properties in each island population of T. seychellensis. Negative values indicate a recent population expansion, values close to zero indicate a stable population, and any values considerably over zero indicate a recent population decrease. Both D and Fs were calculated using Arlequin v.3.5.1.3 (Excoffier and Lischer, 2010). Input files for Arlequin were created using PGDSpider v.2.0.3.0 (Lischer and Excoffier, 2012). 2.4 Testing monophyly of Heterixalus The multispecies coalescent method as implemented in ∗BEAST (Heled and Drummond, 2010) was used to infer the species trees for Tachycnemis and Heterixalus spp., treating mtDNA (cytb, cox1 and 16s), tyr, rag1 and rho as four separately evolving loci. Sequence data for Heterixalus spp. were previously published (Wollenberg et al., 2007) and obtained from GenBank, and those for Tachycnemis were newly generated. It is recommended to include a minimum of two specimens per species for ∗BEAST analyses so that there is a coalescent event with which to estimate population size (Heled and Drummond, 2010), but this was not possible for all species of Heterixalus because of inadequate specimen and/or character sampling in GenBank. For this reason H. alboguttatus, H. boettgeri and H. carbonei were excluded from these analyses. The remaining taxa nonetheless included representatives of all five Heterixalus species groups identified by Wollenberg et al. (2007). Many studies using multilocus datasets do not partition by codon position, but we ran two sets of analyses in order to test for discrepancies between this ad hoc method and the optimal partitioning strategy identified by PartitionFinder (Table 3). Due to over parameterization, convergence was not reached in the identified optimal partitioning scheme in further analyses and therefore only the results of the locus partitioned analysis are used. Preliminary analyses of the locus-partitioned dataset suggested that strict clocks be implemented for the mtDNA and rhod partitions and an uncorrelated lognormal relaxed clock for the tyr and rag1. Rates for molecular clocks were initially set at default (1.0) for all partitions and estimated relative to the mtDNA partition. Two MCMC chains were run for 2 × 108 generations, with trees sampled every 10,000 generations, to ensure convergence was reached the first 5% were discarded as burn-in, although convergence was reached prior to this cut-off. The species-tree Yule-process prior was used with the piecewise linear and constant-root population-size model. Convergence of all parameters was verified using Tracer v.1.5 (Rambaut and Drummond, 2009). 3 Results 3.1 Monophyly of Heterixalus? We aligned 1500 bp of mtDNA (consisting of 424 variable sites (v.s.), of which 391 were parsimony informative (p.s.)), 760 bp of rag1 (62 v.s., 48 p.s.), 357 bp of rho (30 v.s., 15 p.s.), and 611 bp of tyr (56 v.s., 45 p.s.) giving a total sequence length of 3228 bp. Partitioning schemes and nucleotide models used in analyses are presented in Table 3. Partitioning the ∗BEAST dataset by linked loci provided evidence for T. seychellensis being the sister taxon to a monophyletic Heterixalus; a sister-group relationship between H. madagascariensis and H. punctatus; for H. andrakata being the sister taxon to H. tricolor + H. variabilis; and for H. betsileo being the sister taxon to the andrakata–tricolor–variabilis clade (Fig. 1). The relationships of the remaining two species (H. rutenbergi and H. luteostriatus) are unresolved (Fig. 1). Among Heterixalus species and clades, mean p-distances for cytb range from 3% to 19.5%, with the p-distance between Tachycnemis and Heterixalus being 21% (Fig. 1). Given an approximate rate of 0.6–1% per million years for cytb in amphibians (see Elmer et al., 2007), this marker indicates that T. seychellensis diverged from its closest sampled relative in the region of 11.5–19.2 Ma. 3.2 Genetic variation within T. seychellensis We aligned three mitochondrial genes for the 52 specimens: 16s consisted of 599 bp with three variable sites (v.s.), all of which were parsimony-informative (p.s.); cytb 763 bp (32 v.s., 30 p.s.); and cox1 786 bp (16 v.s., 13 p.s.). The dataset was almost complete, with very little missing sequence data across all genes and no genes missing for any individual. No saturation was detected. The rag1 and tyr data were constant in 20 and 19 samples sampled across all populations, respectively. The pomc data consisted of 629 bp (7 v.s., 6 p.s.) for 50 specimens; and rho 337 bp (2 v.s., 2 p.s.) for 25 specimens. The sample size for pomc and rho was reduced because of a shorter amplified sequence length of some samples and because of a lack of confidence in the accuracy of the PHASE calling of the small number of variable sites. The best partitioning strategies and models as determined by PartitionFinder for the mtDNA analyses were the same under AIC and AICc but different under BIC (Table 2). Thus, two BEAST analyses were run under these alternatives, and the resulting tree topologies were identical and support values nearly so (Fig. 2). The mtDNA has a maximum p-distance of 1.5% between any of the seven haplotype groups, and no simple geographic structure is observed in the mtDNA tree (Fig. 3), with samples from all islands except La Digue comprising two haplotype clades that are not sister groups, although not all relationships are well supported. The mean p-distance for cytb among the main mtDNA haplotype clades ranges from 0.4% to 1.5% (Fig. 2). Given an approximate rate of 0.6–1% per million years for cytb in amphibians (see Elmer et al., 2007), this marker suggests that extant mtDNA haplotype lineages of T. seychellensis began diverging approximately 0.75–1.25 Ma. The two variable nuclear genes yielded networks with a general lack of geographic structure (Fig. 3). For pomc (Fig. 3a) there is a small amount of population structuring, with endemic haplotypes shared by multiple individuals within Praslin and Mahé. For rho (Fig. 3b) each of the four haplotypes are found on all islands except La Digue (two haplotypes). Fu’s Fs results indicate recent expansions for all of the island populations with maximal significance (Table 4). Tajima’s D values suggest an opposite trend, with positive values indicating either a population size decrease or balancing selection, but Tajima’s D results are not significant for any island (Table 4). 4 Discussion 4.1 Monophyly of Heterixalus? It has been estimated that T. seychellensis diverged from its closest living relative in Madagascar 9.79–35.34 Ma (Crottini et al., 2012), which implies transoceanic dispersal to the Seychelles given that this microcontinent split from Madagascar approximately 84 Ma (Ali and Aitchison, 2008; Plummer and Belle, 1995) and India by 64 Ma (McKenzie and Sclater, 1971; Norton and Sclater, 1979). Transoceanic dispersal remains a rarely documented phenomenon in amphibians, but see Hedges et al. (1992), Measey et al. (2007), Vences et al. (2003a,b, 2004). Our analyses using the multispecies coalescent support previous studies based on concatenated multilocus DNA sequence data (e.g. Pyron and Wiens, 2011; Vences et al., 2003a,b; Wollenberg et al. 2007) that have hypothesised T. seychellensis to be the sister taxon to Heterixalus. Translation of cytb p-distances to divergence times among lineages produces estimates that fall within Crottini et al.’s (2012) estimate of 9.79–35.34 Ma for the divergence between T. seychellensis and its closest living relative. The cytb p-distances between Tachycnemis and Heterixalus spp. (Fig. 1) are clearly more in agreement with overseas dispersal than Seychelles-Africa or Seychelles-Madagascar Cretaceous vicariance as an explanation for the origin of Tachycnemis in the Seychelles. 4.2 Genetic variation within T. seychellensis The results of our genetic analyses are consistent with Nussbaum and Wu’s (1995) interpretation that the Seychelles treefrog represents a single species. The low levels of genetic diversity within T. seychellensis and lack of notable phylogeographic structure can be explained by a rapid range expansion (supported by results for Fu’s Fs) and/or by multiple admixture events possibly during eustatic sea-level fluctuations. The latter is plausible given that all island populations, apart from La Digue, have multiple mtDNA haplotype clades and that nuDNA haplotypes show no clear geographic structuring. The relatively low levels of genetic variation within the Seychelles treefrog are comparable with several other Seychelles taxa such as Drosophila flies (Legrand et al., 2009) and freshwater turtles (Silva et al., 2010), although the turtles are probably a recent human introduction (Fritz et al., 2013). Conversely, studies of other taxa including lizards (Rocha et al., 2010a,b, 2011, 2013; Valente et al., 2014), a freshwater crab (Daniels, 2011), and a sooglossid frog (Taylor et al., 2012) have revealed much higher levels of inter-island genetic variation. It remains to be fully assessed whether differences in patterns of genetic variation among Seychelles organisms can be explained by ecology (and dispersal ability) and/or duration of residency. The presence of low genetic diversity and little phylogeographic structure but high morphological variation as is observed in T. seychellensis is unusual in (at least adult) anurans and it is difficult to find any examples in the literature (though see e.g., Gvoždík et al., 2008, 2010). The combination of low levels of genetic diversity (and little phylogeographic structure) within T. seychellensis yet substantial morphological variation is perhaps best explained by rapid local adaptation to different environmental settings, ecophenotypic plasticity, or from previous genetic bottlenecks and/or continuing small population sizes (see also Nussbaum and Wu, 1995). The latter explanation seems unlikely on the islands of Mahé and Praslin where T. seychellensis is abundant in the sampled populations (STM, RAN, DJG pers. obs.). These explanations could be tested using population-genetic analyses of data from more rapidly evolving nuclear markers. Genetic (cytb p-) distances between populations of T. seychellensis on different islands (see Fig. 2) provide no evidence for admixture between the islands after 200–333 ka. This might suggest that during the most recent sea-level fluctuations (∼10 ka), where all islands would have been connected (Colonna et al., 1996; Rohling et al., 1998; Siddall et al., 2003; Camoin et al., 2004; Miller et al., 2005), little migration occurred or, if migration did occur, mitochondrial haplotypes did not become fixed. 5 Conclusions We find support for the sister-group relationship between T. seychellensis and a monophyletic Heterixalus. There is little genetic variation within T. seychellensis, even among populations on the four different islands within its range, and the variation is not strongly spatially structured. This is consistent with Nussbaum and Wu’s (1995) interpretation that there is a single species of Seychelles treefrog. The patterns of genetic variation that we have discovered do not allow us to reject Nussbaum and Wu’s (1995) proposal that substantial morphological variation within T. seychellensis is the result of local ecological adaptation and/or small population sizes now and/or in the past, though ecophenotypic plasticity might also be considered.

          Related collections

          Most cited references50

          • Record: found
          • Abstract: found
          • Article: not found

          The Phanerozoic record of global sea-level change.

          K. Miller (2005)
          We review Phanerozoic sea-level changes [543 million years ago (Ma) to the present] on various time scales and present a new sea-level record for the past 100 million years (My). Long-term sea level peaked at 100 +/- 50 meters during the Cretaceous, implying that ocean-crust production rates were much lower than previously inferred. Sea level mirrors oxygen isotope variations, reflecting ice-volume change on the 10(4)- to 10(6)-year scale, but a link between oxygen isotope and sea level on the 10(7)-year scale must be due to temperature changes that we attribute to tectonically controlled carbon dioxide variations. Sea-level change has influenced phytoplankton evolution, ocean chemistry, and the loci of carbonate, organic carbon, and siliciclastic sediment burial. Over the past 100 My, sea-level changes reflect global climate evolution from a time of ephemeral Antarctic ice sheets (100 to 33 Ma), through a time of large ice sheets primarily in Antarctica (33 to 2.5 Ma), to a world with large Antarctic and large, variable Northern Hemisphere ice sheets (2.5 Ma to the present).
            Bookmark
            • Record: found
            • Abstract: found
            • Article: not found

            Accounting for decay of linkage disequilibrium in haplotype inference and missing-data imputation.

            Although many algorithms exist for estimating haplotypes from genotype data, none of them take full account of both the decay of linkage disequilibrium (LD) with distance and the order and spacing of genotyped markers. Here, we describe an algorithm that does take these factors into account, using a flexible model for the decay of LD with distance that can handle both "blocklike" and "nonblocklike" patterns of LD. We compare the accuracy of this approach with a range of other available algorithms in three ways: for reconstruction of randomly paired, molecularly determined male X chromosome haplotypes; for reconstruction of haplotypes obtained from trios in an autosomal region; and for estimation of missing genotypes in 50 autosomal genes that have been completely resequenced in 24 African Americans and 23 individuals of European descent. For the autosomal data sets, our new approach clearly outperforms the best available methods, whereas its accuracy in inferring the X chromosome haplotypes is only slightly superior. For estimation of missing genotypes, our method performed slightly better when the two subsamples were combined than when they were analyzed separately, which illustrates its robustness to population stratification. Our method is implemented in the software package PHASE (v2.1.1), available from the Stephens Lab Web site.
              Bookmark
              • Record: found
              • Abstract: found
              • Article: not found

              Dynamics of mitochondrial DNA evolution in animals: amplification and sequencing with conserved primers.

              With a standard set of primers directed toward conserved regions, we have used the polymerase chain reaction to amplify homologous segments of mtDNA from more than 100 animal species, including mammals, birds, amphibians, fishes, and some invertebrates. Amplification and direct sequencing were possible using unpurified mtDNA from nanogram samples of fresh specimens and microgram amounts of tissues preserved for months in alcohol or decades in the dry state. The bird and fish sequences evolve with the same strong bias toward transitions that holds for mammals. However, because the light strand of birds is deficient in thymine, thymine to cytosine transitions are less common than in other taxa. Amino acid replacement in a segment of the cytochrome b gene is faster in mammals and birds than in fishes and the pattern of replacements fits the structural hypothesis for cytochrome b. The unexpectedly wide taxonomic utility of these primers offers opportunities for phylogenetic and population research.
                Bookmark

                Author and article information

                Journal
                24555995
                4101239
                10.1016/j.ympev.2014.02.004
                http://creativecommons.org/licenses/by/3.0/

                Evolutionary Biology
                dna,frogs,heterixalus,indian ocean,phylogeography
                Evolutionary Biology
                dna, frogs, heterixalus, indian ocean, phylogeography

                Comments

                Comment on this article