Open Access
How to translate text using browser tools
19 November 2014 Dispersal and speciation in purple swamphens (Rallidae: Porphyrio)
Juan C. Garcia-R., Steve A. Trewick
Author Affiliations +
Abstract

Dispersal, when accompanied by reduced gene flow and natural selection, influences speciation rates among groups of organisms. We used molecular phylogenetics, divergence time estimates, and population genetics to reconstruct the mode, pattern, and tempo of diversification within the wide-ranging purple swamphens (genus Porphyrio), with emphasis on the “supertramp” P. porphyrio. Our results suggest that the Porphyrio clade arose during the Middle Miocene in Africa, with a single colonization in the Americas and several other colonizations in Southeast Asia and the Indo-Pacific around 10 mya. We found that the widespread P. porphyrio is not monophyletic. Indeed, several subspecies and subspecies groups may represent species-level lineages. The P. p. melanotus lineage probably reached Australasia during the Pleistocene (600 kya), although some islands were colonized only in the past few hundred years. New Zealand, and some other islands, had previously been colonized (∼2.5 mya) by flying Porphyrio that evolved into flightless endemic species. Early and recent lineages are now sympatric. Widespread occupation of oceanic islands implies high dispersal and colonization rates, but gene flow probably occurs episodically and follows varying routes at different times. This pattern of colonization enables populations to differentiate and, ultimately, speciate.

INTRODUCTION

The ability to disperse and colonize new habitats provides organisms with ecological opportunities to harvest novel resources and establish new populations. Vagrant species that reach, via long-distance dispersal, new regions or isolated islands that are thousands of kilometers from their traditional breeding range may generate new flocks of dispersers before being totally or partially displaced by more efficient competitors (Diamond 1974, 1975). This “supertramp” strategy can also give rise to numerous, phenotypically distinct variants, which may lead to speciation (Simpson 1953, Diamond 1974, Diamond et al. 1976, Grant 1986, Whittaker 1998, Crisp et al. 2011). However, for a single species to persist, without differentiation, over a large breeding range of fragmented habitat, individuals must move between habitat patches at a rate sufficient to counter the evolutionary effects of isolation. This requirement predicts that such widespread species have a high level of gene flow among populations distributed across several landscapes. Alternatively, relatively small, isolated populations can segregate and accumulate phenotypic differences if range expansion is not constant through time or if the direction of dispersal changes. Outcomes may vary, depending on the number of independent radiations, differences in diversification rates, rate and pattern of gene flow, and rapidity of species radiation following a wave of dispersal (Mayr and Diamond 2001, Moyle et al. 2009, Cibois et al. 2011).

Inferring the geographic origin and temporal diversification of organisms is an essential part of biogeography and depends on an accurate estimate of evolutionary relationships among species (Rosen 1978, Filardi and Moyle 2005). In the case of birds, enabled by flight to disperse long distances, the spatiotemporal patterns of diversification can be challenging to analyze, especially when shallow radiations on islands generate differential morphological traits that obscure evolutionary affinities (Filardi and Moyle 2005, Irestedt et al. 2013). Dispersal and adaptation together are important drivers of insular diversification of many bird groups (Pratt 2005, Grant and Grant 2008) and account for much of the diversity that we find today in archipelagos (Trewick and Gibb 2010, Trewick 2011). Colonization of islands sometimes involves loss of the capability for further long-distance dispersal, when flight is not integral to foraging, social interaction, or predator avoidance (McNab 1994, McNab and Ellis 2006, Steadman 2006). Reduction, and even loss, of flight capacity can be an adaptive response to island life, and some speciation may occur with adaptation of flightlessness, as a result of altered selective environments (Milá et al. 2010, Sly et al. 2011, Alonso et al. 2012, Runemark et al. 2012).

Family Rallidae (Aves: Gruiformes) is diverse and cosmopolitan. It includes common species that are good dispersers, as well as regional and island endemics. Many oceanic islands that were naturally without terrestrial mammal predators appear to have favored reversion to a terrestrial lifestyle after colonization and speciation by flying ancestors (Ripley 1977, Steadman 2006). This combination of high dispersal and high endemicity associated with the loss of flight makes them interesting subjects for evolutionary analysis. In particular, the large, flamboyant purple swamphens (genus Porphyrio) demonstrate extraordinary dispersal capabilities, with evidence of multiple invasions, apparently spaced out in time, that resulted in divergences of size, color, and other traits (Ripley 1977, Remsen and Parker 1990, Trewick 1996). Seven species of purple swamphens are currently recognized, 4 of which are or were present in the Oceania region (Trewick 1996, Taylor 1998). Principal among these is the widespread “supertramp” Purple Swamphen (Porphyrio porphyrio), which occurs from Africa and the Mediterranean east to the Pacific (Ripley 1977). This taxon comprises apparently parapatric morphological variants that have, at times, been classified into ∼13 subspecies or species (Figure 1A; Ripley 1977, Taylor 1998). Although sometimes considered a reluctant flier (Craig 1977, Craig and Jamieson 1990), this taxon has nevertheless established populations on many oceanic islands, throughout the Indian and western Pacific Ocean (Mayr 1949, Ripley 1977, Trewick 1997, 2011). At least 2 colonizations of New Zealand resulted in the presence of the North Island Takahe (P. mantelli) and South Island Takahe (P. hochstetteri), endemic flightless herbivores that were sympatric with flying swamphens (Trewick 1997, Trewick and Worthy 2001). Some island populations and subspecies are known only from fossils that reveal the numerous extinctions that followed colonization of those islands by people (Steadman 1995, 2006, Steadman et al. 1999). Insular endemics have been recognized as distinct species on New Caledonia and New Zealand (Balouet and Olson 1989, Trewick and Worthy 2001) and on other Pacific islands (Steadman 1988, Kirchman and Steadman 2006).

FIGURE 1.

Geographic distribution, phylogenetic relationships, and time of divergence of purple swamphens. (A) Distributions and sampling localities of the Porphyrio porphyrio subspecies included in our study (source: IUCN). Subspecies caledonicus and vitiensis are included within samoensis (following Ripley 1997). (B) Geographic distributions of the other 6 Porphyrio species around the world: 1 = P. alleni (Africa), 2 = P. martinica (North and South America), 3 = P. flavirostris (South America), 4 = P. hochstetteri (South Island, New Zealand), 5 = P. mantelli (North Island, New Zealand), and 6 = P. albus (Lord Howe Island, Australia). (C) Chronogram based on analysis of concatenated sequences of mitochondrial and nuclear genes using a relaxed-clock Bayesian analysis in BEAST. Age constraints were based on estimates by Garcia-R. et al. (2014) of the basal divergence of Porphyrio and the split between Porphyrio and Amaurornis in a normal distribution (see text). Gray bars show estimated time of divergence and 95% HPD intervals of node ages. Support values for key clades are indicated below branches and correspond to bootstrap supports (>50%) and posterior probabilities (>0.80), respectively.

i0004-8038-132-1-140-f01.tif

We used multilocus DNA sequence data to generate a dated phylogenetic hypothesis of relationships within Porphyrio and to explore the pattern of gene flow among populations of P. porphyrio. We address the following questions to gain insights into the biogeographic origin and diversification of these birds: (1) What is the phylogenetic structure of the genus? (2) What time of diversification, and pattern of dispersal and colonization, explains current diversity? (3) Is there support for a single or multiple range expansions? (4) Are regional subspecies of P. porphyrio monophyletic, or is there a mismatch between clade structure and taxonomy?

METHODS

Sampling

To obtain DNA for analysis, we sampled bones, toe pads, feathers, blood, and muscle tissue from specimens of the 7 known species in the genus Porphyrio (Figure 1B), including representatives of P. porphyrio subspecies from Africa, Europe, Asia, and Pacific islands (Table 1 and Figure 1A). Additionally, we sampled several populations in New Zealand and Australia separated by ∼1,500 km of sea and graded terrestrial landscapes to explore gene flow at different spatial scales.

TABLE 1.

Taxa, museum voucher numbers, locality, type of tissue, and GenBank accession numbers of data included in our study. An asterisk indicates sequences <200 bp (see Appendix B).

i0004-8038-132-1-140-t01.eps

DNA Extraction

DNA extractions from bones and toe pads were carried out in a dedicated ancient DNA (aDNA) laboratory (Ecology Group, Massey University, Palmerston North, New Zealand; see DNA Toolkit at  http://evolves.massey.ac.nz). DNA extractions from toe pad samples were performed using the QiAMP DNA Minikit (Qiagen, Valencia, California, USA), following the manufacturer's instructions and standard procedures for aDNA (Cooper and Poinar 2000, Rohland and Hofreiter 2007). DNA from bones was extracted using decalcification with EDTA and proteinase K digestion in Tris-buffered saline, followed by purification with phenol–chloroform. DNA from fresh tissues was extracted in a laboratory geographically separated from the aDNA laboratory, using either Tissue DNeasy kit (Qiagen; following the manufacturer's instructions) or incubation at 55°C with proteinase K and a CTAB buffer (2% Hexadecyl trimethyl ammonium bromide, 100 mM Tris–HCl, pH 8.0, 1.4 M NaCl, 20 mM EDTA), followed by a combined phenol–chloroform–isoamyl alcohol (25:24:1) cleanup.

Mitochondrial and Nuclear DNA Amplification

We sequenced 2 mitochondrial genes and 1 nuclear gene for population genetic analyses of P. porphyrio in Australia and New Zealand: mitochondrial control region (CR) and cytochrome oxidase b (cyt b), plus a fragment of the nuclear beta-fibrinogen intron 7 (BFG-7). For phylogenetic analysis of Porphyrio, parts of 2 additional mitochondrial genes (ribosomal RNA 12S and 16S) and 1 nuclear gene (recombination activating gene 1 [RAG-1]) were amplified from representative specimens of the currently recognized species and subspecies of P. porphyrio. Additional sequences were downloaded from GenBank (Table 4 in Appendix A). Standard polymerase chain reaction (PCR) methods (using the primers listed in Table 5 in Appendix A) were used for amplification of nuclear and mitochondrial fragments. Amplification products were purified with QIAquick PCR cleanup kit (Qiagen) or ExoI/SAP digest. For each PCR product, both strands were sequenced using Big Dye Terminator version 3.1 reagents and an ABI 3730XL automated DNA sequencer (Applied Biosystems, Foster City, California, USA). Short sequences without GenBank accession numbers are provided in Appendix B.

Phylogeny and Divergence Times

All sequences were edited, assembled, and aligned using Geneious version 6.0.5 (Drummond et al. 2012a) and checked by eye. Alignments of ribosomal genes, CR, and BFG-7 were conducted using Gblocks version 0.91b (Castresana 2000) and evaluated by eye. Cytochrome oxidase b and RAG-1 were checked for the presence of indels and stop codons. Prior to concatenated analyses, we performed individual gene tree analysis to detect spurious branch-length patterns and evidence of significant incongruence. We built a supermatrix with a six-way partition by gene: 12S, 16S, CR, cyt b, BFG-7, and RAG-1 (Wiens 2006, Holland et al. 2007, Wiens and Moen 2008, Johnson et al. 2012). Maximum likelihood (ML) trees were implemented in RAxML version 8.0.24 via the CIPRES portal (Miller et al. 2010). We used a general time-reversible model with gamma distribution (GTR + Γ), which allowed RAxML to halt bootstrap resampling automatically (bootstopping) once split support values converged (Pattengale et al. 2010). We conducted Bayesian phylogenetic analyses using MrBayes version 3.2.2 as implemented in the CIPRES portal under a GTR + Γ+ I model of evolution. The model was estimated in ModelTest version 3.7, using Akaike's Information Criterion (Posada and Crandall 1998). After performing shorter test runs, we conducted 3 parallel runs of the Metropolis Coupled Markov Chain Monte Carlo (MCMCMC) algorithm for 5 million generations each, sampling 1 tree with associated parameter values per 5,000 generations, and employing 3 heated chains and 1 cold chain. Convergence and diagnostics of the Markov process were visualized using Tracer version 1.6 (see Acknowledgments). The first half million generations (10%) were discarded as burn-in. A burn-in of 10% gave optimal results, and we obtained effective sample sizes (ESS) >200 for 95% of the parameters. The ML and Bayesian trees were viewed using FigTree version 1.4.2 (see Acknowledgments) and SplitsTree version 4.12.8 (Huson and Bryant 2006). Amaurornis flavirostra was used as an outgroup to root the tree.

Divergence times among lineages of Porphyrio were estimated using a relaxed Bayesian clock implemented in BEAST version 1.7.5 (Drummond et al. 2012b). For calibration constraints, we used the basal divergence estimate of Porphyrio with a normal distribution of 11–20 Ma (95% range) and the basal split of Amaurornis flavirostra and Porphyrio with a normal distribution of 27–35 Ma (95% range), as previously calculated from an analysis using a widely sampled dataset of mitochondrial and nuclear genes (Garcia-R. et al. 2014). We combined the results of 3 independent runs of 30 million generations to ensure ESS scores >200 for 95% of the parameters in each run. Chains were sampled every 4,000 generations, and a burn-in of 10% (3 million generations) was used. The tree, with times of divergence and highest posterior density (HPD) intervals, was visualized using FigTree.

Population Differentiation and Demographic History

Population-level analyses were carried out with 2 datasets: (1) concatenated mitochondrial loci CR and cyt b; and (2) the autosomal locus BFG-7. Sequence ambiguities at heterozygous sites in BFG-7 that indicated separate alleles were resolved using PHASE implemented in DnaSP version 5.0 (Librado and Rozas 2009) with the default parameters. To test for intralocus recombination in BFG-7, we used the PHI test (Bruen et al. 2006) implemented in SplitsTree. This is a robust test that can reliably detect recombination and report few false positives (Martin et al. 2011). We calculated the following summary statistics for genetic variation of each population in DnaSP: number of haplotypes (h), nucleotide diversity per site (π), number of segregating sites (S), Watterson's estimator of per site population mutation rate (θW), Tajima's D statistic (DT), and Ramos and Rozas's R2-test. For each population, DT was also analyzed, using 103 coalescent simulations conditioned on the sample size and the observed number of segregating sites (Hudson 1990). Demographic expansion was assessed using the R2-test implemented in DnaSP because it is the most powerful test when dealing with limited sample sizes (Rozas et al. 2003).

Lynch and Crease's pairwise FST (Lynch and Crease 1990) was calculated in DnaSP, using 5,000 replicates and a significance level of ≤0.05 to test the null hypothesis of panmixia (Raymond and Rousset 1995) between pairs of P. porphyrio populations in Oceania. The level of population genetic structure was tested using an analysis of molecular variance (AMOVA) implemented in Arlequin version 3.5 (Excoffier et al. 2005). To complement the phylogenetic inferences, haplotype networks were constructed using Network version 4.5.1.0 (Bandelt et al. 1999) with median joining to visualize the relationship between haplotypes and their geographic distribution. A Mantel test was conducted for correlation between uncorrected mitochondrial genetic distances estimated in MEGA version 5.2 (Tamura et al. 2011) and linear geographic distances (Jensen et al. 2005) using the ade4 package (Dray and Dufour 2007) in R (R Development Core Team 2008) with 10,000 permutations.

RESULTS

Phylogenetic Analyses

The present study includes wider taxonomic representation than the first molecular phylogenetic study of Porphyrio, which compared 4 taxa using a single gene, mitochondrial 12S rRNA (Trewick 1997). We have included Porphyrio alleni, P. flavirostris, and P. albus, as well as the P. porphyrio subspecies porphyrio, indicus, poliocephalus and other Australasian subspecies (note that Trewick [1997] included P. p. seistanicus from Turkey). The complete alignment of 6 gene fragments contained 4,304 base pairs [bp], comprising 816 bp of cyt b, 699 bp of CR, 728 bp of 16S, 402 bp of 12S, 868 bp of RAG-1, and 791 bp of BFG-7. No premature stop codons were detected in the 2 protein-coding genes. Individual gene trees did not reveal spurious sequences or significant conflict among individual phylogenies. Phylogenetic analyses that included just 1 representative of each species and subspecies that had the most complete gene sets yielded similar topologies (results not shown). Topologies from ML and Bayesian analyses were congruent for the concatenated dataset. The African species P. alleni is sister to the New World species pair P. martinica and P. flavirostris (Figure 1C). Porphyrio porphyrio did not form a monophyletic group. Instead, it comprised 6 distinct clades (porphyrio, indicus, madagascariensis, pulverulentus, poliocephalus [including seistanicus], and melanotus), and it was paraphyletic with respect to 3 species-level taxa: P. mantelli and P. hochstetteri from New Zealand, and P. albus from Lord Howe Island (Figure 1C, clade A). Porphyrio p. melanotus (Figure 1C, clade B) includes the parapatrically distributed subspecies bellus, caledonicus, samoensis, vitiensis, palliatus, pelewensis, melanopterus, and chathamensis (the latter was previously demonstrated to be invariant at the 12S locus by Trewick [1997]). This phylogenetic and spatial structure corresponds with significant differences in color and size that have been previously described in some detail (Mayr 1949, Ripley 1977, Simmons et al. 1980, Sangster 1998, Taylor 1998). One Indonesian specimen (P. p. indicus) did not group with other specimens from this region but instead was more closely related to the clade composed of P. p. pulverulentus and P. albus (Figure 1C). The close similarity of P. p. pulverulentus and P. albus sequences indicates a complex history of exchange, because their lineage is not recorded in islands between. Sequence data for P. albus came from old and rare museum specimens, and such aDNA sources have to be treated with caution. However, consistent results were obtained from separate samples and replicate PCRs. There is the possibility of mislabeling of museum specimens, but the white plumage characteristic of P. albus is uncommon in other populations.

Molecular Dating

Divergence time analysis suggests a Middle Miocene origin of diversification within Porphyrio, with the split between the lineages that led to the African species P. alleni and to other species occurring earliest, around 14 (19–9) mya. Splitting among P. porphyrio “subspecies” was estimated to have occurred about 6 (11–2) mya, with a likely colonization of P. p. melanotus in Australasia occurring in the late Pleistocene (600 kya). However, an earlier colonization by a flying P. porphyrio at the start of the Pleistocene (∼2.5 mya) resulted in the flightless takahe endemic to New Zealand.

Population Genetic Structure

The mitochondrial data contained 1 to 4 haplotypes (h) in each of the populations sampled (Figure 2A). Nucleotide diversity (π) at sampling localities with n ≥ 2 was variable, ranging from 0.0 in Northland and Otago, New Zealand, to 0.0059 in Victoria, Australia. Zero to 12 segregating sites (S) were observed in each population, yielding a population mutation rate per site (θW) between 0.0 and 0.0065 (Table 2). For BFG-7, two alternative haplotypes were identified for alleles possessed by heterozygous individuals. No statistically significant evidence for recombination (P = 0.06) was detected using a phi test. Between 4 and 14 inferred BFG-7 sequences were sampled per population, and 1 to 4 unique haplotypes were found (Table 2). The number of segregating sites in each population varied from 0 to 6. Aside from invariant samples from New South Wales and Western Australia (north), nucleotide diversity ranged from 0.00069 in Palmerston North, New Zealand, to 0.00440 in Victoria. This latter population also had the highest inferred population mutation rate per site (0.00414). No population showed a significantly skewed DT for mitochondrial DNA (mtDNA) or nuclear data, although power to reject the null hypothesis of neutrality may have been hampered by small sample sizes. The population-size-change (R2) test did not find evidence of demographic expansion of the populations sampled (Table 2).

FIGURE 2.

(A) Localities in Australia and New Zealand where individual Porphyrio porphyrio were sampled for population genetic analyses and haplotype networks. Colored circles are proportional to sample size at each locality. Inset images show plumage coloration of P. p. bellus (left) and P. p. melanotus (right). Median-joining haplotype networks of (B) mitochondrial DNA (mtDNA) dataset and (C) nuclear gene BFG-7 (nDNA). Circle area is proportional to the number of individuals found of each haplotype. Each line connecting haplotypes indicates a mutational step.

i0004-8038-132-1-140-f02.tif

TABLE 2.

Summary of descriptive statistics for mtDNA data (cyt b and CR) and the nuclear locus (BFG-7) used in population genetic analyses.

i0004-8038-132-1-140-t02.eps

Panmixia was evident in 2 localities sampled on the same island and separated by ∼500 km: Palmerston North and Northland on North Island, New Zealand. Nevertheless, population genetic structuring was found among other populations more geographically remote from one another, with the exception of some of the comparisons among Western Australia (north) and Victoria, which was probably due to the low sample size (Table 3). Analysis of BFG-7 showed little population genetic structuring (Table 3) among localities at which mtDNA diversity was clearly partitioned. Significantly different values were obtained for most pairwise comparisons among Victoria populations because this location had endemic haplotypes. Consistent with the population pairwise differentiation analysis of the mitochondrial data, most genetic variation was explained by differences among populations (70.5%; P < 0.0001). The AMOVA of the BFG-7 data indicated that a small but significant (14%; P < 0.01) component of variance was attributable among populations. Population geographic structure was evident in the mitochondrial and nuclear haplotype networks (Figure 2B, 2C), and the Mantel test showed a significant correlation between genetic and geographic distances among populations (r = 0.508, P < 0.05), even though only ∼26% of genetic divergence was explained by geographic distance.

TABLE 3.

Pairwise comparisons between populations using mtDNA data (above diagonal) and the nuclear locus BFG-7 (below diagonal). Negative values represent a program idiosyncrasy due to the small sample size and are effectively zero. Significant values are in bold (*0.01 < P < 0.05; **0.001 < P < 0.01).

i0004-8038-132-1-140-t03.eps

DISCUSSION

Biogeography and Evolution of Swamphens

Our phylogenetic analyses and molecular dating support independent and temporally nonoverlapping colonization events among Porphyrio species. This interpretation is, however, based on surviving or recently extinct lineages only; other colonizations are represented by fossils on Oceanic islands (Steadman 1988, 2006, Steadman et al. 1999) or have left no trace at all. The most likely area of origin of Porphyrio is Africa, with colonization westward into the Americas and several other colonizations northeastward (Europe, Asia, and Oceania) during the Miocene and Pleistocene. The oldest split among the currently recognized P. porphyrio lineage (Figure 1C, clade A) occurred in the Late Miocene (∼6 mya), giving rise to P. p. porphyrio on the Mediterranean coast of Europe and P. p. indicus in Indonesia. Further diversification took place during the Pliocene, giving rise to P. p. madagascariensis in Africa and a radiation into Oceania.

The unique sequence obtained from the extinct P. albus of Lord Howe Island suggests a close affinity to Philippine P. p. pulverulentus, indicating that it was perhaps a white color variant founded from P. p. pulverulentus migrants. The flightless status of P. albus appears to be equivocal, and the population seems to have been polymorphic for plumage, with a high frequency of white individuals (White 1790, Hindwood 1940, Greenway 1967). Aberrations in color have been found in some insular populations, caused perhaps by an allele fixed through a founder effect (Cunningham 1955, Steadman 2006, Uy et al. 2009). White Porphyrio occur intermittently, and recent observations include an individual P. p. melanotus in Otago, New Zealand (Trewick and Morgan-Richards 2014). The Lord Howe population may have been established from a small number of colonizing individuals from the Philippines during the late Pleistocene (∼500 kya), but this would have involved dispersal from the Philippines to Lord Howe Island over other islands. We remain cautious about the short DNA sequence obtained from P. albus.

Despite the appearance that flight is used only infrequently among P. porphyrio subspecies, the lineage has dispersed, colonized, and established populations multiple times across open expanses of water. Haplotypes from Indonesia (a specimen from Java is closely related to a specimen from the Philippines) and New Caledonia (specimens from this locality are closely related to specimens from localities as far away as Sulawesi and Palau) support the inference of exchange (Figure 1C). Although rare misplaced haplotypes of this sort might be evidence of ongoing exchange among island populations, they could also be the product of incomplete lineage sorting or past migration events.

Porphyrio p. melanotus (Figure 1C, clade B) appears to have entered Australasia within the past 600,000 yr, but bone deposits show a more recent arrival on some remote islands (Millener 1981, Taylor 1998, Steadman 2006). This includes New Zealand, where deposits indicate colonization ∼500 yr ago, after Polynesian settlement (Trewick and Worthy 2001). This dating is much more recent than the estimated divergence of P. porphyrio and takahe lineages (P. hochstetteri and P. mantelli) that must represent an earlier, separate colonization. Within the Australia–New Zealand geographic region, the distribution of genetic variability (phylogenetic analysis, FST, AMOVA, haplotype networks, and Mantel test) indicates that the genetic structure of P. p. melanotus populations is not homogeneous. This lineage may have originated in Wallacea, and eustatic sea-level changes could have aided colonization by reducing overwater dispersal distances. Lowered sea level during glacial phases of the Pleistocene reduced the distance between some land areas, including between Papua New Guinea and Australia (Voris 2000, Hall 2009, Jønsson et al. 2010, Wurster et al. 2010, Lohman et al. 2011, Condamine et al. 2013, Irestedt et al. 2013). They did not, however, significantly alter the overwater distance between Australia and New Zealand (Graham 2008). This colonization pattern has created the allopatric distribution currently shown in Oceania. We note that higher genetic diversity in Australia than across the Tasman Sea reflects a recent arrival in New Zealand and indicates an influence of distribution due to persistence of habitat rather than geographic distance correlated with human settlement and clearance of forest.

Pairwise differences among Australasian populations were higher for the mtDNA data than for the nuclear DNA gene. This difference may be explained by the higher mutation rate and lack of recombination in the mitochondrial genome (Neiman and Taylor 2009). However, different population genetic (e.g., background selection), demographic (e.g., effective sex ratio and/or male-biased migration rates), or natural selection also must be considered for those higher mtDNA FST estimates (Palumbi and Baker 1994, Charlesworth 1998, 2009, Stinchcombe and Hoekstra 2008, Muir et al. 2012). Despite the indication that the populations studied have not undergone recent demographic changes, with a marked reduction of genetic variation within populations and increasing genetic differentiation, the swamphens within P. p. melanotus of south Western Australia (Figure 1C, clade B) show exceptional and not subtle differentiation in plumage color pattern (Whittell 1934). The current nominate subspecies P. p. bellus in south Western Australia has a prominently brighter blue breast and throat color than P. p. melanotus (see images in Figure 2A). Differences in color, size, and other traits are evident among other lineages within the melanotus clade (Ripley 1977) and in other clades. For instance, within the poliocephalus clade, the Middle Eastern “seistanicus” population is grayish compared with individuals from India (see color photographs in Figure 1C). The mismatch between plumage patterns and the distribution of neutral population genetic markers suggests that differentiation in color and other traits have arisen rapidly in Porphyrio and are subject to selection in local environments (Mayr 1954, Nosil et al. 2009, Feder et al. 2012) or to stochastic genetic drift (Clegg et al. 2002a, 2002b). The lack of sorting at the BFG-7 locus suggests that fixed plumage-color differences among populations are not the result of drift but are perhaps better explained by “purifying” sexual selection.

Selection that results in character divergence among populations can occur without being detectable by neutral genetic markers (Charlesworth and Charlesworth 2009, Nosil et al. 2009). There may be lineages with genomic region(s) involved in adaptive divergence, and these regions may respond independently to environment pressures via selection (Schneider et al. 1999, Clegg et al. 2002a, 2002b, Schluter 2009, Via 2009, Cooke et al. 2012). Appearance is a trait that is important in assortative mating by individuals within a population (Schluter 2009, Maan and Seehausen 2011) and is strongly implicated in the behavior of communally breeding P. p. melanotus (Jamieson 1988). As such, it may drive monomorphism in local populations. Selection on mate choice and kin fitness likely maintains local population phenotypes in stable frequencies (Andersson et al. 1998, Eaton 2005, Johnsen et al. 2006, Pryke and Griffith 2006, Murphy 2008).

Although P. p. melanotus is not the lineage that gave rise to the flightless insular species of New Zealand (P. mantelli and P. hochstetteri), the recent success of P. p. melanotus in reaching several remote islands in the Pacific is testimony to the high success rate of dispersal and colonization. However, the restriction of gene flow evident in FST values suggests that range expansion is probably episodic, and this enhances the opportunity for speciation and establishment of reproductive barriers. Sexual selection could help drive locus-specific evolution without being evident in genes that are (with respect to those traits) neutral, and reproductive isolation could evolve as a consequence of local adaptation and selection on characters involved in mate choice and inclusive fitness by way of mating behavior.

ACKNOWLEDGMENTS

The sampling of Porphyrio species and the extensive sampling of P. porphyrio in Africa, Europe, Asia, and Oceania were supplemented by contributions from museum collections, researchers, and institutions. We are grateful to the following individuals and institutions who provided tissues necessary for our study: W. Boles and J. Sladek at the Australian Museum (Australia); L. Joseph and R. Palmer at the Australian National Wildlife Collection (Australia); J. Summer at the Museum Victoria (Australia); C. Stevenson at the Western Australian Museum (Australia); C. Miskelly and A. Tennyson at the Te Papa Museum (New Zealand); C. Fisher and T. Parker at the Liverpool Museum (England); R. Prys-Jones at the Tring Museum of Natural History (England); R. T. Brumfield at the Louisiana University Museum of Natural Science (USA); D. Willard at the Field Museum of Natural History (USA); P. Sweet and C. Filardi at the American Museum of Natural History (USA); J. Fjeldså and J. Kristensen at the Natural History Museum of Denmark (Denmark); S. Birks at the Burke Museum (USA); E. Garcia Franquesa at the Barcelona Museum (Spain); M. Marquez at the Banco de Tejidos Animales de Cataluya (Spain); G. Davies at the Ditsong Museum (South Africa); M. Trotter, R. Adams, and S. Pilkington at Fish and Game (New Zealand); T. Burns, M.Sc. student at Massey University (New Zealand); A. Gamauf at the Vienna Museum (Austria); C. M. Milensky at the Smithsonian Institution (USA); R. Urriza at the Philippines National Museum (Philippines); R. Hinlo (Philippines); D. Philippe (France); and S. A. Subrata (Indonesia). For helpful advice on the analysis, we thank B. Holland (University of Tasmania). Images of birds used in figures were generously provided by J. Dolphijn, L. and V. Dunis, A. d'Affonseca, C. Gouraud, J. Sagardia, H. Weerman, J. Babbington, M. Montoro, V. Moreno, S. A. Subrata, N. Ramesh, P. Bourdin, and D. Gambas Izquierdo. We are grateful to I. Castro and F. Sheldon for helpful comments that greatly improved the manuscript. J.C.G.-R thanks the Massey Doctoral Scholarship, New Zealand International Doctoral Scholarship, A. M. Soto, and T. Garcia. Tracer software is available at  http://tree.bio.ed.ac.uk/software/tracer/. FigTree software is available at  http://tree.bio.ed.ac.uk/software/figtree/.

LITERATURE CITED

1.

R Alonso A. J Crawfordand E Bermingham (2012). Molecular phylogeny of an endemic radiation of Cuban toads (Bufonidae: Peltophryne) based on mitochondrial and nuclear genes. Journal of Biogeography 39:434–451. Google Scholar

2.

S Andersson J Örnborgand M Andersson (1998). Ultraviolet sexual dimorphism and assortative mating in Blue Tits. Proceedings of the Royal Society of London, Series B 265:445–450. Google Scholar

3.

J. C Balouetand S. L Olson (1989). Fossil birds from Late Quaternary deposits in New Caledonia. Smithsonian Contributions to Zoology 469. Google Scholar

4.

H.-J Bandelt P Forsterand A Röhl (1999). Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 16:37–48. Google Scholar

5.

T. C Bruen H Philippeand D Bryant (2006). A simple and robust statistical test for detecting the presence of recombination. Genetics 172:2665–2681. Google Scholar

6.

J Castresana (2000). Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Molecular Biology and Evolution 17:540–552. Google Scholar

7.

G Chambersand E MacAvoy (1999). Molecular genetic analysis of hybridisation. Science for Conservation 29. Wellington: Department of Conservation. Google Scholar

8.

B Charlesworth (1998). Measures of divergence between populations and the effect of forces that reduce variability. Molecular Biology and Evolution 15:538–543. Google Scholar

9.

B Charlesworth (2009). Effective population size and patterns of molecular evolution and variation. Nature Reviews Genetics 10:195–205. Google Scholar

10.

B Charlesworthand D Charlesworth (2009). Darwin and genetics. Genetics 183:757–766. Google Scholar

11.

A Cibois J. S Beadell G. R Graves E Pasquet B Slikas S. A Sonsthagen J.-C Thibaultand R. C Fleischer (2011). Charting the course of reed-warblers across the Pacific islands. Journal of Biogeography 38:1963–1975. Google Scholar

12.

S. M Clegg S. M Degnan J Kikkawa C Moritz A Estoupand I. P. F Owens (2002a). Genetic consequences of sequential founder events by an island-colonizing bird. Proceedings of the National Academy of Sciences USA 99:8127–8132. Google Scholar

13.

S. M Clegg S. M Degnan C Moritz A Estoup J Kikkawaand I. P. F Owens (2002b). Microevolution in island forms: The roles of drift and directional selection in morphological divergence of a passerine bird. Evolution 56:2090–2099. Google Scholar

14.

F. L Condamine E. F. A Toussaint A. M Cotton G. S Genson F. A. H Sperlingand G. J Kergoat (2013). Fine-scale biogeographical and temporal diversification processes of Peacock Swallowtails (Papilio subgenus Achillides) in the Indo-Australian Archipelago. Cladistics 29:88–111. Google Scholar

15.

G. M Cooke N. L Chaoand L. B Beheregaray (2012). Divergent natural selection with gene flow along major environmental gradients in Amazonia: Insights from genome scans, population genetics and phylogeography of the characin fish Triportheus albus. Molecular Ecology 21:2410–2427. Google Scholar

16.

A Cooper (1994). Molecular evolutionary studies of New Zealand birds. School of Biological Sciences. Victoria University, New Zealand. Google Scholar

17.

A Cooperand H. N Poinar (2000). Ancient DNA: Do it right or not at all. Science 289:1139. Google Scholar

18.

A Cooper C Mourer-Chauviré G Chambers A von Haeseler A Wilsonand S Pääbo (1992). Independent origins of New Zealand moas and kiwis. Proceedings of the National Academy of Sciences USA 89:8741–8744. Google Scholar

19.

J. L Craig (1977). The behaviour of the pukeko, Porphyrio porphyrio melanotus. New Zealand Journal of Zoology 4:413–433. Google Scholar

20.

J. L Craigand I. G Jamieson (1990). Pukeko: Different approaches and some different answers. In Cooperative Breeding in Birds: Long-term Studies of Ecology and Behaviour ( P. B Staceyand W. D Koenig Editors). Cambridge University Press, Cambridge, UK. pp. 387–412. Google Scholar

21.

M. D Crisp S. A Trewickand L. G Cook (2011). Hypothesis testing in biogeography. Trends in Ecology & Evolution 26:66–72. Google Scholar

22.

J. M Cunningham (1955). A colony of Buff-coloured Pukeko. Notornis 6(3):83–84. Google Scholar

23.

J. M Diamond (1974). Colonization of exploded volcanic islands by birds: The supertramp strategy. Science 184:803–806. Google Scholar

24.

J. M Diamond (1975). Assembly of species communities. In Ecology and Evolution of Communities ( M. L Codyand J. M Diamond Editors). Harvard University Press, Cambridge, MA, USA. pp. 342–444. Google Scholar

25.

J. M Diamond M. E Gilpinand E Mayr (1976). Species–distance relation for birds of the Solomon Archipelago, and the paradox of the great speciators. Proceedings of the National Academy of Sciences USA 73:2160–2164. Google Scholar

26.

S Drayand A.-B Dufour (2007). The ade4 package: Implementing the duality diagram for ecologists. Journal of Statistical Software 22:1–20. Google Scholar

27.

A. J Drummond B Ashton S Buxton M Cheung A Cooper C Duran M Field J Heled M Kearse S Markowitz R Moir S Stones-Havas et al . (2012a). Geneious version 6.05.  http://www.geneious.com Google Scholar

28.

A. J Drummond M. A Suchard D Xieand A Rambaut (2012b). Bayesian phylogenetics with BEAUti and the BEAST 1.7. Molecular Biology and Evolution 29:1969–1973. Google Scholar

29.

M. D Eaton (2005). Human vision fails to distinguish widespread sexual dichromatism among sexually “monochromatic” birds. Proceedings of the National Academy of Sciences USA 102:10942–10946. Google Scholar

30.

L Excoffier G Lavaland S Schneider (2005). Arlequin (version 3.0): An integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online 1:47–50. Google Scholar

31.

J. L Feder S. P Eganand P Nosil (2012). The genomics of speciation-with-gene-flow. Trends in Genetics 28:342–350. Google Scholar

32.

C. E Filardiand R. G Moyle (2005). Single origin of a pan-Pacific bird group and upstream colonization of Australasia. Nature 438:216–219. Google Scholar

33.

J. C Garcia-R G. C Gibband S. A Trewick (2014). Deep global evolutionary radiation in birds: Diversification and trait evolution in the cosmopolitan bird family Rallidae. Molecular Phylogenetics and Evolution 80:96–108. Google Scholar

34.

I. J Graham (Editor) (2008). A Continent on the Move: New Zealand Geoscience into the 21st Century. Geological Society of New Zealand in association with GNS Science, Wellington, New Zealand. Google Scholar

35.

P. R Grant (1986). Ecology and Evolution of Darwin's Finches. Princeton University Press, Princeton, NJ, USA. Google Scholar

36.

P. R Grantand B. R Grant (2008). How and Why Species Multiply: The Radiation of Darwin's Finches. Princeton University Press, Princeton, NJ, USA. Google Scholar

37.

J. C Greenway (1967). Extinct and Vanishing Birds of the World, second edition. Dover, New York, NY, USA. Google Scholar

38.

J Grothand G Barrowclough (1999). Basal divergences in birds and the phylogenetic utility of the nuclear RAG-1 gene. Molecular Phylogenetics and Evolution 12:115–123. Google Scholar

39.

R Hall (2009). Southeast Asia's changing palaeogeography. Blumea 54:148–161. Google Scholar

40.

K. A Hindwood (1940). The birds of Lord Howe Island. The Emu 40:1–86. Google Scholar

41.

B Holland G Conner K Huberand V Moulton (2007). Imputing supertrees and supernetworks from quartets. Systematic Biology 56:57–67. Google Scholar

42.

R. R Hudson (1990). Gene genealogies and the coalescent process. In Oxford Surveys in Evolutionary Biology, vol. 7 ( D Futuymaand J Antonovics Editors). Oxford University Press, Oxford, UK. pp. 1–44. Google Scholar

43.

D. H Husonand D Bryant (2006). Application of phylogenetic networks in evolutionary studies. Molecular Biology and Evolution 23:254–267. Google Scholar

44.

M Irestedt P.-H Fabre H Batalha-Filho K. A Jønsson C. S Roselaar G Sangsterand P. G. P Ericson (2013). The spatio-temporal colonization and diversification across the Indo-Pacific by a ‘great speciator' (Aves, Erythropitta erythrogaster). Proceedings of the Royal Society of London, Series B 280:20130309. Google Scholar

45.

M Irestedt U Johansson T Parsonsand P Ericson (2001). Phylogeny of major lineages of suboscines (Passeriformes) analysed by nuclear DNA sequence data. Journal of Avian Biology 32:15–25. Google Scholar

46.

I. G Jamieson (1988). Provisioning behaviour in a communal breeder: An epigenetic approach to the study of individual variation in behaviour. Behaviour 104:262–280. Google Scholar

47.

J. L Jensen A. J Bohonakand S. T Kelley (2005). Isolation by distance, web service. BMC Genetics 6:13. Google Scholar

48.

A Johnsen S Andersson J. G Fernandez B Kempenaers V Pavel S Questiau M Raess E Rindaland J. T Lifjeld (2006). Molecular and phenotypic divergence in the Bluethroat (Luscinia svecica) subspecies complex. Molecular Ecology 15:4033–4047. Google Scholar

49.

K. A Johnson B. R Holland M. M Heslewoodand D. M Crayn (2012). Supermatrices, supertrees and serendipitous scaffolding: Inferring a well-resolved, genus-level phylogeny of Styphelioideae (Ericaceae) despite missing data. Molecular Phylogenetics and Evolution 62:146–158. Google Scholar

50.

K. A Jønsson R. C. K Bowie R. G Moyle L Christidis J. A Norman B. W Benzand J Fjeldså (2010). Historical biogeography of an Indo-Pacific passerine bird family (Pachycephalidae): Different colonization patterns in the Indonesian and Melanesian archipelagos. Journal of Biogeography 37:245–257. Google Scholar

51.

R. T Kimball E. L Braun F. K Barker R. C. K Bowie M. J Braun J. L Chojnowski S. J Hackett et al . (2009). A well-tested set of primers to amplify regions spread across the avian genome. Molecular Phylogenetics and Evolution 50:654–660. Google Scholar

52.

J Kirchman (2012). Speciation of flightless rails on islands: A DNA-based phylogeny of the typical rails of the Pacific. The Auk 129:56–69. Google Scholar

53.

J. J Kirchmanand D. W Steadman (2006). New species of rails (Aves: Rallidae) from an archaeological site on Huahine, Society Islands. Pacific Science 60:281–297. Google Scholar

54.

T Kocher W Thomas A Meyer S Edwards S Pääbo F Villablancaand A Wilson (1989). Dynamics of mitochondrial DNA evolution in animals: Amplification and sequencing with conserved primers. Proceedings of the National Academy of Sciences USA 86:6196–6200. Google Scholar

55.

P Libradoand J Rozas (2009). DnaSP v5: A software for comprehensive analysis of DNA polymorphism data. Bioinformatics 25:1451–1452. Google Scholar

56.

B. C Livezey (1998). A phylogenetic analysis of the Gruiformes (Aves) based on morphological characters, with an emphasis on the rails (Rallidae). Philosophical Transactions of the Royal Society of London, Series B 353:2077–2151. Google Scholar

57.

D. J Lohman M de Bruyn T Page K von Rintelen R Hall P. K. L Ng H.-T Shih G. R Carvalhoand T von Rintelen (2011). Biogeography of the Indo-Australian Archipelago. Annual Review of Ecology, Evolution, and Systematics 42:205–226. Google Scholar

58.

M Lynchand T. J Crease (1990). The analysis of population survey data on DNA sequence variation. Molecular Biology and Evolution 7:377–394. Google Scholar

59.

M. E Maanand O Seehausen (2011). Ecology, sexual selection and speciation. Ecology Letters 14:591–602. Google Scholar

60.

D. P Martin P Lemeyand D Posada (2011). Analysing recombination in nucleotide sequences. Molecular Ecology Resources 11:943–955. Google Scholar

61.

E Mayr (1949). Notes on the birds of Northern Melanesia. 2. American Museum Novitates 1417. Google Scholar

62.

E Mayr (1954). Change of genetic environment and evolution. In Evolution as a Process ( J Huxley A. C Hardyand E. B Ford Editors). Allen and Unwin, London, UK. pp. 157–180. Google Scholar

63.

E Mayrand J Diamond (2001). The Birds of Northern Melanesia: Speciation, Ecology, and Biogeography. Oxford University Press, New York, NY, USA. Google Scholar

64.

B. K McNab (1994). Energy conservation and the evolution of flightlessness in birds. The American Naturalist 144:628–642. Google Scholar

65.

B. K McNaband H. I Ellis (2006). Flightless rails endemic to islands have lower energy expenditures and clutch sizes than flighted rails on islands and continents. Comparative Biochemistry and Physiology A 145:295–311. Google Scholar

66.

B Milá B. H Warren P Heeband C Thébaud (2010). The geographic scale of diversification on islands: Genetic and morphological divergence at a very small spatial scale in the Mascarene Grey White-eye (Aves: Zosterops borbonicus). BMC Evolutionary Biology 10:158. Google Scholar

67.

P. R Millener (1981). The Quaternary avifauna of the North Island, New Zealand. Ph.D. dissertation, University of Auckland, Auckland, New Zealand. Google Scholar

68.

M. A Miller W Pfeifferand T Schwartz (2010). Creating the CIPRES Science Gateway for inference of large phylogenetic trees. In Proceedings of the Gateway Computing Environments Workshop (GCE 2010). Institute of Electrical and Electronics Engineers, New York, NY, USA. pp. 1–8. Google Scholar

69.

M Morgan-Richards S. A Trewick A Bartosch-Härlid O Kardailsky M. J Phillips P. A McLenachanand D Penny (2008). Bird evolution: Testing the Metaves clade with six new mitochondrial genomes. BMC Evolutionary Biology 8:1–12. Google Scholar

70.

R. G Moyle C. E Filardi C. E Smithand J Diamond (2009). Explosive Pleistocene diversification and hemispheric expansion of a “great speciator.” Proceedings of the National Academy of Sciences USA 106:1863–1868. Google Scholar

71.

G Muir C. J Dixon A. L Harperand D. A Filatov (2012). Dynamics of drift, gene flow, and selection during speciation in Silene. Evolution 66:1447–1458. Google Scholar

72.

T. G Murphy (2008). Lack of assortative mating for tail, body size, or condition in the elaborate monomorphic Turquoise-browed Motmot (Eumomota superciliosa). The Auk 125:11–19. Google Scholar

73.

M Neimanand D. R Taylor (2009). The causes of mutation accumulation in mitochondrial genomes. Proceedings of the Royal Society of London, Series B 276:1201–1209. Google Scholar

74.

P Nosil D. J Funkand D Ortiz-Barrientos (2009). Divergent selection and heterogeneous genomic divergence. Molecular Ecology 18:375–402. Google Scholar

75.

S. R Palumbiand C. S Baker (1994). Contrasting population structure from nuclear intron sequences and mtDNA of humpback whales. Molecular Biology and Evolution 11:426–435. Google Scholar

76.

N. D Pattengale M Alipour O. R. P Bininda-Emonds B. M. E Moretand A Stamatakis (2010). How many bootstrap replicates are necessary? Journal of Computational Biology 17:337–354. Google Scholar

77.

D Posadaand K. A Crandall (1998). Modeltest: Testing the model of DNA substitution. Bioinformatics 14:817–818. Google Scholar

78.

H. D Pratt (2005). The Hawaiian Honeycreepers: Drepanidinae (Bird Families of the World). Oxford University Press, Oxford, UK. Google Scholar

79.

T. M Prychitkoand W. S Moore (1997). The utility of DNA sequences of an intron from the beta-Fibrinogen gene in phylogenetic analysis of woodpeckers (Aves: Picidae). Molecular Phylogenetics and Evolution 8:193–204. Google Scholar

80.

S. R Prykeand S. C Griffith (2006). Red dominates black: agonistic signalling among head morphs in the colour polymorphic Gouldian Finch. Proceedings of the Royal Society of London, Series B 273:949–957. Google Scholar

81.

R Development Core Team (2008). R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria. Google Scholar

82.

M Raymondand F Rousset (1995). An exact test for population differentiation. Evolution 49:1280–1283. Google Scholar

83.

J. V Remsen Jr and T. A. Parker III (1990). Seasonal distribution of the Azure Gallinule (Porphyrula flavirostris), with comments on vagrancy in rails and gallinules. The Wilson Bulletin 102:380–399. Google Scholar

84.

S. D Ripley (1977). Rails of the World: A Monograph of the Family Rallidae. David R. Godine, Boston, MA. Google Scholar

85.

N Rohlandand M Hofreiter (2007). Ancient DNA extraction from bones and teeth. Nature Protocols 2:1756–1762. Google Scholar

86.

D. E Rosen (1978). Vicariant patterns and historical explanation in biogeography. Systematic Biology 27:159–188. Google Scholar

87.

J Rozas J. C Sánchez-DelBarrio X Messeguerand R Rozas (2003). DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics 19:2496–2497. Google Scholar

88.

A Runemark J Hey B Hanssonand E. I Svensson (2012). Vicariance divergence and gene flow among islet populations of an endemic lizard. Molecular Ecology 21:117–129. Google Scholar

89.

G Sangster (1998). Purple swamphen is a complex of species. Dutch Birding 20:13–22. Google Scholar

90.

D Schluter (2009). Evidence for ecological speciation and its alternative. Science 323:737–741. Google Scholar

91.

C. J Schneider T. B Smith B Larisonand C Moritz (1999). A test of alternative models of diversification in tropical rainforests: Ecological gradients vs. rainforest refugia. Proceedings of the National Academy of Sciences USA 96:13869–13873. Google Scholar

92.

K. E. L Simmons R Gillmor P. A. D Hollom R Hudson E. M Nicholson M. A Ogilvie P. J. S Olney C. S Roselaar K. H Voous D. I. M Wallaceand J Wattel (1980). Handbook of the Birds of Europe, the Middle East and North Africa. The Birds of the Western Palearctic. Oxford University Press, Oxford, UK. Google Scholar

93.

G. G Simpson (1953). The Major Features of Evolution. Columbia University Press, New York, NY, USA. Google Scholar

94.

N. D Sly A. K Townsend C. C Rimmer J. M Townsend S. C Lattaand I. J Lovette (2011). Ancient islands and modern invasions: Disparate phylogeographic histories among Hispaniola's endemic birds. Molecular Ecology 20:5012–5024. Google Scholar

95.

D. W Steadman (1988). A new species of Porphyrio (Aves: Rallidae) from archaeological sites in the Marquesas Islands. Proceedings of the Biological Society of Washington 101:162–170. Google Scholar

96.

D. W Steadman (1995). Prehistoric extinctions of Pacific island birds: Biodiversity meets zooarchaeology. Science 267:1123–1131. Google Scholar

97.

D. W Steadman (2006). Extinction and Biogeography of Tropical Pacific Birds. University of Chicago Press, Chicago, IL, USA. Google Scholar

98.

D. W Steadman J. P Whiteand J Allen (1999). Prehistoric birds from New Ireland, Papua New Guinea: Extinctions on a large Melanesian island. Proceedings of the National Academy of Sciences USA 96:2563–2568. Google Scholar

99.

J. R Stinchcombeand H. E Hoekstra (2008). Combining population genomics and quantitative genetics: Finding the genes underlying ecologically important traits. Heredity 100:158–170. Google Scholar

100.

K Tamura D Peterson N Peterson G Stecher M Neiand S Kumar (2011). MEGA5: Molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Molecular Biology and Evolution 28:2731–2739. Google Scholar

101.

B Taylor (1998). Rails: A Guide to the Rails, Crakes, Gallinules and Coots of the World. Yale University Press, New Haven, CT, USA. Google Scholar

102.

R Thomas W Schaffner A Wilsonand S Pääbo (1989). DNA phylogeny of the extinct marsupial wolf. Nature 340:465–467. Google Scholar

103.

S. A Trewick (1996). Morphology and evolution of two takahe: Flightless rails of New Zealand. Journal of Zoology 238:221–237. Google Scholar

104.

S. A Trewick (1997). Flightlessness and phylogeny amongst endemic rails (Aves: Rallidae) of the New Zealand region. Philosophical Transactions of the Royal Society of London, Series B 352:429–446. Google Scholar

105.

S. [A. Trewick (2011). Vicars & vagrants. Australasian Science 32 (September):24–27. Google Scholar

106.

S. A Trewickand G. C Gibb (2010). Vicars, tramps and assembly of the New Zealand avifauna: A review of molecular phylogenetic evidence. Ibis 152:226–253. Google Scholar

107.

S. [A Trewick and M Morgan-Richards (2014). NZ Wild Life: Introducing the Weird and Wonderful Character of Natural New Zealand. Penguin Books, Melbourne, Australia. Google Scholar

108.

S. A Trewickand T. H Worthy (2001). Origins and prehistoric ecology of takahe based on morphometric, molecular, and fossil data. In The Takahe: Fifty Years of Conservation Management and Research ( W. G Leeand I. G Jamieson Editors). University of Otago Press, Dunedin, New Zealand. pp. 31–48. Google Scholar

109.

J. A. C Uy R. G Moyle C. E Filardiand Z. A Cheviron (2009). Difference in plumage color used in species recognition between incipient species is linked to a single amino acid substitution in the melanocortin-1 receptor. The American Naturalist 174:244–254. Google Scholar

110.

S Via (2009). Natural selection in action during speciation. Proceedings of the National Academy of Sciences USA 106:9939–9946. Google Scholar

111.

H. K Voris (2000). Maps of Pleistocene sea levels in Southeast Asia: Shorelines, river systems and time durations. Journal of Biogeography 27:1153–1167. Google Scholar

112.

J White (1790). Journal of a Voyage to New South Wales with sixty-five plates of non descript animals, birds, lizards, serpents, curious cones of trees and other natural productions. J. Debrett, London, UK. Google Scholar

113.

R. J Whittaker (1998). Island Biogeography: Ecology, Evolution, and Conservation. Oxford University Press, Oxford, UK. Google Scholar

114.

H. M Whittell (1934). The swamp-hen (Porphyrio) in Western Australia. The Emu 34:85–95. Google Scholar

115.

J. J Wiens (2006). Missing data and the design of phylogenetic analyses. Journal of Biomedical Informatics 39:34–42. Google Scholar

116.

J. J Wiensand D. S Moen (2008). Missing data and the accuracy of Bayesian phylogenetics. Journal of Systematics and Evolution 46:307–314. Google Scholar

117.

C. M Wurster M. I Bird I. D Bull F Creed C Bryant J. A. J Dungaitand V Paz (2010). Forest contraction in north equatorial Southeast Asia during the Last Glacial Period. Proceedings of the National Academy of Sciences 107:15508–15511. Google Scholar

Appendices

APPENDIX A

TABLE 4.

Taxa, GenBank accession numbers, and original source of data for additional DNA sequences included in our study.

i0004-8038-132-1-140-t04.eps

TABLE 5.

Primers for PCR and DNA sequencing employed in our study. Asterisks denote primers taken from the primer database of the Allan Wilson Centre for Molecular Ecology and Evolution, Massey University.

i0004-8038-132-1-140-t05.eps

APPENDIX B

Short Sequences

These sequences are <200 bp, too short to be submitted to GenBank. Specimen name is followed by museum voucher when available. Acronyms for museums are the same as in Table 1.

12S rRNA

>Porphyrio porphyrio porphyrio_BM 93-0242-T

AGTACCCGCCTGAGAACTACGAGCACAAACGCTTAAAACTCTAAGGACTTGGCGGTGC TCCAAACCCACCTAGAGGAGCCTGTTCTGTAATCGATAACCCACGATATACCCAACCCCTTCTCGCCCAAAGCAGC

>Porphyrio porphyrio pelewensis_LIV T9774

AACTGGGATTAGATACCCCACTATGCTTGGCCCTAAATCCAGATACTCACCACCACT AGAGTATCCGCCTGAGAACTACGAGCACAAACGCTTAAAACTCTAAGGACTTGGCGGTGCC CCAAACCCACCTAGAGGAGCCTGTTCTGTAATCGATAACCCACGATATACCCAACCCCTT CTTGCCCAAAGCAGC

>Porphyrio porphyrio palliatus_LIV T9048

AACTGGGATTAGATACCCCACTATGCTTGGCCCTAAATCCAGATACTCACTACCACTAG AGTATCCGCCTGAGAACTACGAGCACAAACGCTTAAAACTCTAAGGACTTGGCGGTGCCCCA AACCCACCTAGAGGAGCCTGTTCTGTAATCGATAACCCACGATATACCCAACCCCTTCTTGC CCAAAGCAGC

>Porphyrio porphyrio poliocephalus_AMNH DOT17002

CGATATACCCAACCCCTTCTTGCCCAAAGCAGCCTACATACCGCCGTCCCCAGCTCACCTCC CCTGAGAGCCTAAATAGTGAGCACAACAACACCTCGCTAATAAGACAGGTCAAGGTATAGCCCAT GAAGGGGTAGAAATGGGCTACATTTTCTAAAATAGAAA

>Porphyrio porphyrio caledonicus

CGATATACCCAACCCCTTCTTGCCCAAAGCAGCCTACATACCGCCGTCCCCAGCTCACCT CCCCTGAGAGCCTAAATAGTGAGCACAACAACACCTCGCTAATAAGACAGGTCAAGGTATAGC CCATGAAGGGGTAGAAATGGGCTACATTTTCTAAAATAGAAA

Cyt b

>Porphyrio mantelli_NMNZ DM7930

GGATCATACCTCTATAAAGAAACCTGAAACACAGGAATCATCCTACTACTCACCCTAATA GCCACTGCCTTCGTAGGCTATGTCCTACCATGAGGACAAATATCCTTCTGAGGCGCTACAGTC ATTACAAACCTATTCTCAGCCATC

Juan C. Garcia-R. and Steve A. Trewick "Dispersal and speciation in purple swamphens (Rallidae: Porphyrio)," The Auk 132(1), 140-155, (19 November 2014). https://doi.org/10.1642/AUK-14-114.1
Received: 16 May 2014; Accepted: 1 September 2014; Published: 19 November 2014
KEYWORDS
biogeography
dispersal
phylogeny
speciation
Back to Top