Identify whether hybridization would likely lead to speciation or homogenization in each scenario.

Hybridization among diverging lineages is common in nature. Genomic data provide a special opportunity to characterize the history of hybridization and the genetic basis of speciation. We review existing methods and empirical studies to identify recent advances in the genomics of hybridization, as well as issues that need to be addressed. Notable progress has been made in the development of methods for detecting hybridization and inferring individual ancestries. However, few approaches reconstruct the magnitude and timing of gene flow, estimate the fitness of hybrids or incorporate knowledge of recombination rate. Empirical studies indicate that the genomic consequences of hybridization are complex, including a highly heterogeneous landscape of differentiation. Inferred characteristics of hybridization differ substantially among species groups. Loci showing unusual patterns – which may contribute to reproductive barriers – are usually scattered throughout the genome, with potential enrichment in sex chromosomes and regions of reduced recombination. We caution against the growing trend of interpreting genomic variation in summary statistics across genomes as evidence of differential gene flow. We argue that converting genomic patterns into useful inferences about hybridization will ultimately require models and methods that directly incorporate key ingredients of speciation, including the dynamic nature of gene flow, selection acting in hybrid populations and recombination rate variation.

Keywords: evolutionary genomics, gene flow, hybridization, introgression, reproductive barriers, speciation

The scope of hybridization plays a special role in understanding how new species are born. The capacity of lineages to hybridize is often used as a basis for species identification (Dobzhansky 1937; Mayr 1963). More generally, the dynamics of gene flow between diverging populations describe key elements of the speciation process.

Hybridization leaves detectable footprints in genomes, raising the prospect that it can be characterized from genomic patterns of variation. In particular, genomic data provide avenues to answer two basic questions about speciation: (i) what is the history of gene flow between nascent species and (ii) what genetic barriers maintain their integrity? Although these questions were addressed before the genomic era, the ability to rapidly and affordably survey genomic variation among individuals and populations has ushered in new hope of answering them (Sousa & Hey 2013; Seehausen et al. 2014). Because of recombination and independent assortment during meiosis, diversity patterns at the very large number of loci sampled by genomic data sets can be viewed (ideally) as replicated outcomes of hybridization history, enabling increasingly accurate reconstruction of that history. At the same time, inspection of the genomic distribution of variation can pinpoint specific regions with unusual patterns that might contribute to reproductive isolation or adaptive introgression.

Genomic studies of hybridization generally follow a recipe with two steps. First, genomewide patterns of variation are catalogued. If each of the sampled populations is suspected or determined to belong to one diverging lineage (e.g. species) or another, differentiation between these populations is quantified. If the sampled populations likely reflect recent hybridization (e.g. in hybrid zones), other measures of admixture – such as geographic clines in allele frequency across transects – might be used. The second step is to compare observed genomic distributions to those expected under one or more models to draw inferences about hybridization and speciation.

Genomewide portraits of differentiation and admixture are shaped by a suite of evolutionary processes. Parameters of particular interest from the speciation perspective include the form, frequency, magnitude and timing of hybridization, as well as the fitness of hybrids. But several other factors complicate interpretations. Natural selection acting before or after hybridization affects differentiation at targeted mutations and at nearby neutral variants (Wright 1969; Charlesworth et al. 1997; Nordborg 1997), thereby inflating heterogeneity in differentiation among loci, even in the absence of gene flow (Noor & Bennett 2009; Nachman & Payseur 2012; Cruickshank & Hahn 2014; Burri et al. 2015). Recombination rate – a parameter that varies both across the genome and among species (Smukowski & Noor 2011) – determines the genomic scale over which patterns of differentiation and admixture vary. Shared variation among populations may reflect unsorted ancestral polymorphism (incomplete lineage sorting) rather than hybridization (Clark 1997; Wakeley & Hey 1997).

In this paper, we synthesize emerging trends in the genomics of hybridization between diverging lineages. After reviewing analytical methods that can be used to characterize hybridization, we summarize findings from empirical studies conducted on the genomic scale. Based on this survey, we prioritize future research to maximize insights into the process of speciation.

Translating genomic data sets into useful inferences about hybridization and speciation is a formidable task. Recognizing the increasing practicality of collecting high-quality genomic data, we focus here on analytical prospects and challenges. We refer the reader to recent reviews for guidance on important issues related to sequencing, genotyping and variant calling (Nielsen et al. 2011; Altmann et al. 2012; Arnold et al. 2013; Buerkle & Gompert 2013; Mardis 2013; Narum et al. 2013; Schlötterer et al. 2014; Olson et al. 2015).

We surveyed the literature for analytical methods that have either been applied in genomic studies of hybridization between diverging lineages or have the potential to be applied in such studies. Both methods developed specifically to understand speciation and methods developed for other purposes were included.

A wide range of methods is available (Table 1). The collection of approaches samples major descriptors of genetic variation: levels of diversity and divergence, site frequency spectra, haplotype variation and phylogenetic relationships. Methods vary in their characterization of hybridization. Defining gene flow as the movement of alleles between populations, different methods provide access to contrasting levels of gene flow. For example, coalescent approaches usually assume low rates, whereas analyses of clines consider rates high enough to interfere with selection. Some strategies aim to simply detect gene flow, whereas others strive to reconstruct its magnitude and/or timing. Depending on the approach, estimated rates of gene flow are genomewide averages or locus-specific values. Hybridization is modelled as an instantaneous event or as a continuous phenomenon. For several parametric approaches, there exists a trade-off in computational cost between increasing the number of individuals and increasing the number of loci sampled. For example, methods that fit the isolation with migration model using Markov chain Monte Carlo handle many individuals but are slow with a large number of loci (Hey & Nielsen 2004, 2007; Hey 2010; Sethuraman & Hey 2015), whereas analytical likelihood methods applied to this model can use entire genomes but are restricted to a few individuals (Lohse et al. 2011, 2015, 2016). Principal components analysis (PCA) (a nonparametric approach) can be efficiently applied to whole-genome data sets from large numbers of individuals (Patterson et al. 2006; Price et al. 2006), but it provides no direct insights into the causes of resulting patterns.

Methods differ in the frequency of usage by researchers studying the genomics of speciation (Table 2). The most popular strategy interprets heterogeneity in summary statistics of between-population differentiation – usually FST – as evidence of differential gene flow across the genome. Another commonly used class of approaches searches for signs of gene flow in genomic patterns of shared-derived variants between populations or species. In genomic studies of hybridization, the ‘ABBA-BABA test’ (Green et al. 2010; Durand et al. 2011) is a widespread application of such phylogenetic thinking. Interest in this test and its derivatives (Green et al. 2010; Durand et al. 2011; Patterson et al. 2012; Martin et al. 2015b; Pease & Hahn 2015), referred to as ‘D-statistics’, was stimulated by its original application (Green et al. 2010), which revealed that ancestors of modern humans hybridized with Neanderthals. Several approaches commonly used to infer demographic history within species (especially humans) – including diffusion-based analysis of the site frequency spectrum (Gutenkunst et al. 2009) and a host of methods based on individual haplotypes (Pool & Nielsen 2009; Gravel 2012; Mailund et al. 2012; Harris & Nielsen 2013) – have seen limited application in the genomic characterization of gene flow between diverging lineages.

Inferences about hybridization and speciation from genomic data

Species groupSamplingGenomic dataGenomic patterns of differentiation/introgressionAnalytical method for detecting hybridizationEstimate rate or timing of gene flow?Inferences about hybridization and speciation from authorsReferences
Butterflies81 individuals across Heliconius melpomene, H. timareta, H. elevatusWhole-genome sequencing RAD sequencingHigher gene flow (2–5% of genome) between sympatric than between allopatric H. melpomene vs. H timareta; elevated gene flow in B/D colour regionD-statisticsNoExchange of protective colour pattern genes between species; H. elevatus formed by hybridizationHeliconius Genome Consortium (2012)
Butterflies31 individuals (total) from 5 H. melpomene populations, 1 Heliconius cydno population, 1 H. timareta populationWhole-genome sequencingPervasive phylogenetic discordance, structured by geography; fraction of genome introgressed higher in older comparisons; reduced differentiation between sympatric pairs; higher ratio of sympatric/allopatric FST on Z chromosomePhylogenetic discordance D-statistics FST outliersNoLarge-scale gene flow between H. melpomene and H. cydno/timareta lineage throughout much of their divergence historyMartin et al. (2013)
Butterflies51 individuals (total) from H. melpomene, H. cydno, H. timareta, Heliconius heurippaRAD sequencingAdmixture up to 13% in some individuals; FST highly correlated across loci among sympatric pairs but not parapatric pairs; some differentiated regions associated with geography rather than taxonStructure FST outliersNoGene flow between sympatric H. melpomene and H. cydno/timareta; possible hybrid origin for H. heurippaNadeau et al. (2013)
Butterflies116 Lycaides idas, 76 Lycaides melissa, 186 putative hybridsRAD sequencingAmong hybrids, outlier FST loci enriched for L. idas ancestry; differentiation correlated with genomic cline parametersFST outliers Genomic clinesNoAdaptive divergence and genetic drift in allopatry contribute to reproductive isolation; complex genetic basis to reproductive isolationGompert et al. (2012)
Chipmunks4 Tamias ruficaudus 4 Tamias amoenusExome sequencingGene genealogies show species monophylyPhylogenetic discordance D-statisticsNoLittle to no role for hybridization between these sympatric species, despite earlier introgression of mtDNAGood et al. (2015)
Cichlids10 individuals from each of Pundamilia nyererei, P. pundamilia, P. sp. ‘pink and fin’, Mbipia mbipi, M. LuteaRAD sequencingWhen SNPs are partitioned by levels of differentiation, species relationships change (e.g. high-differentiation SNPs group by colour)FST outliers StructureNoAt least two intergeneric hybridization events; hybridization generates new species and contributes to adaptive radiation; selection plays an important role in speciationKeller et al. (2013)
Cichlids24 species in four sympatric crater lake radiations and 29 outgroupsRAD sequencingWeak support for monophyly of sympatric radiations; species within a crater vary in shared ancestry with outgroupsStructure PCA TreemixNoComplex colonization histories within crater lakes; gene flow with neighbouring rivers after initial colonization of lakes; questions status as compelling model for sympatric speciationMartin et al. (2015a)
Cut-throat trout94 individuals from 5 populations of Onchorhynchus clarkia lewisi, westslope cut-throat troutRAD sequencing1–10% population admixture; several superinvasive alleles (at much higher frequencies than genomic background)Frequency of introgressed alleles at diagnostic lociNoCandidate invasiveness genesHohenlohe et al. 2013;
Drosophila3 Drosophila pseudoobscura (1 each for 3 subspecies), 1 Drosophila persimilis, 1 D. mirandaPartial genome sequencingElevated sequence divergence near inversions; diversity correlated with divergence correlated between D. pseudoobscura and D. persimilis but not with divergence between D. pseudoobscura and D. mirandaDxyNoInversions crucial for persistence of species in the face of recent gene flow; secondary contact between D. persimilis and D. pseduoobscura homogenized collinear regions; evidence for unidirectional gene flow less than 200 kya, but not on the X chromosomeKulathinal et al. (2009)
Drosophila1 Drosophila simulans, 1 D. mauritiana, 1 D. sechelliaWhole-genome sequencingEvidence of gene flow between D. simulans and island endemics at up to 4.6% (autosomes) and 2.2% (X chromosome) of genomeDivergence time heterogeneityNo‘Islands of introgression, in genomes otherwise consistent with isolation’; introgression reduced on the X chromosomeGarrigan et al. (2012)
Drosophila4 Drosophila mojavensis (2 from each of 2 populations); 1 D. arizonaeWhole-genome sequencingOverall divergence similar in sympatric and allopatric comparisons, with elevated divergence in both in inverted regionsBlockwise likelihoodYesPostdivergence gene flow, ceasing at 270 kya; same timing in collinear and inverted regions, but reduced gene flow in inversions; inversions originated close to time of initial splitLohse et al. (2015)
Flycatchers10 Ficedula albiocollis (collared), 10 F. hypoleuca (pied)Whole-genome sequencing50 clusters of elevated FST (overrepresented at chromosome ends); reduced nucleotide diversity (within species), and elevated LD – but reduced Dxy; higher overall divergence on ZFST outliersNoSpecies divergence driven by directional selection at many lociEllegren et al. (2012)
Flycatchers10 Ficedula albiocollis (collared), 10 F. hypoleuca (pied)Whole-genome sequencingSubstantial fraction of shared polymorphisms (22%) with fewer fixed differences (3%)Approximate Bayesian ComputationYesRecent origin of flycatcher species, suggesting rapid evolution of postzygotic isolation; allopatric speciation followed by secondary, recent gene flow; 0.16–0.36 migrants per generation F. hypoleuca into F. albiocollis, no gene flow in the other directionNadachowska-Brzyska et al. (2013)
Flycatchers79 F. albiocollis, 79 F. hypoleuca, 20 F. speculigera, 20 F. semitorquataWhole-genome sequencingHigh heterogeneity in differentiation across the genome; concordant genomic regions of differentiation across species pairsFST outliers Dxy outliers D-statisticsNoLinked selection within species rather than variation in gene flow among species explains islands of differentiationBurri et al. (2015)
Flycatchers2 Zimmerius viridiflavus, 3 Z. chrysops 5 individuals from mosaic population 1 Z. acer 1 Z. gracilipesRAD sequencing1.1% of the alleles in the mosaic population represent introgression from Z. chrysops; introgressed regions enriched in functions involving cell projection and plasma membranesD-statisticsNoIntrogression may account for variation in plumage colorationRheindt et al. (2014)
Frogs45 Craugastor augusti 1 C. tarahumaraensis 1 C. unoRAD sequencingHigh lineage differentiation between evolutionarily significant units – numerous private alleles and high FSTD-statisticsNoTwo episodes of introgression occurred in the same direction between lineages but at different times; higher variation in introgressed lineages than in independently evolving lineagesStreicher et al. (2014)
HomininsHomo neanderthalensis, 5 H. sapiensWhole-genome sequencingEast Asians share more SNPs than sub-Saharan Africans with H. neanderthalensisD-statisticsNoGene flow from H. neanderthalensis into non-African H. sapiens occurred before split of Eurasian groupsGreen et al. (2010)
HomininsHomo neanderthalensis, 1000 H. sapiensWhole-genome sequencingRate of exponential decay of linkage disequilibrium among SNPs in H. sapiens genomic regions introgressed from H. neanderthalensis fits a model of recent gene flowHaplotypesYesLast gene flow from H. neanderthalensis into H. sapiens occurred 37–86 kyaSankararaman et al. (2012)
HomininsHomo neanderthalensis, ‘Denisovans’, 42 H. sapiensWhole-genome sequencingHigher levels of H. neanderthalensis ancestry in East Asians than in EuropeansD-statisticsNoAt least two separate episodes of hybridization between H. neanderthalensis and H. sapiens, including one after split between East Asians and EuropeansWall et al. (2013)
HomininsHomo neanderthalensis, ‘Denisovans’, H. sapiensWhole-genome sequencing1.5–2.1% of non-African H. sapiens genomes derived from H. neanderthalensisD-statisticsNo3–5 cases of interbreeding among extinct hominin populations, including an unknown species; complex history of hybridization leaving signatures in small fractions of the genomePrüfer et al. (2014)
HomininsHomo neanderthalensis, 1004 H. sapiensWhole-genome sequencingH. sapiens genomic regions of reduced H. neanderthalensis ancestry include X chromosome and genie regionsConditional random fieldNoH. neanderthalensis X chromosome contributed to hybrid sterility with H. sapiensSankararaman et al. (2014)
HomininsHomo neanderthalensis, 3 H. sapiensWhole-genome sequencing3.4–7.9% H. neanderthalensis admixture in EurasiaBlock wise likelihoodYesH. neanderthalensis – H. sapiens admixture higher than other studiesLohse & Frantz (2014)
Horses6 individuals from each of Equus quagga quagga, E. quagga boehmi, E. grevyi E. zebra hartmannae, E. kiang, E. hemionus onager, E. africanus somaliensis, E. asinus asinus, E. caballusWhole-genome sequencingFour episodes of gene flow detected: one during the early divergence of the Equus in North America, and three between contemporary Equus lineages in the Old World; D-statistics associated with chromosomal changesD-statistics CoalHMMNoGene flow occurred despite extensive variation in chromosome number, although rearrangements inhibited gene flow; speciation in the face of gene flowJónsson et al. (2014)
House mice1301 mice from two transects of hybrid zone between Mus musculus domesticus and M. m. musculusSNP arrayAsymmetric patterns of linkage disequilibrium across transectsGeographic clinesNoMovement of hybrid zone over timeWang et al. (2011)
House mice679 mice from two transects of hybrid zone between Mus musculus domesticus and M. m. musculusSNP arraySignatures of epistatic selection at many loci; disproportionate number of selected loci on the X chromosome; some selected loci overlap in two transects, although most do notGenomic clinesNoCentral and distal portions of X chromosome contribute to reproductive isolation; complex genetic basis for reproductive isolation, arising in allopatryJanoušek et al. (2012)
House miceMice from three transects of hybrid zone between Mus musculus domesticus and M. m. musculusSNP arrayIntrogression positively correlated with recombination rate and negatively correlated with differentiationGenomic clinesNoSome reproductive barriers shared across hybrid zone; genome properties (including recombination rate) influence speciationJanoušek et al. (2015)
House mice8 wild-derived inbred lines of Mus musculus castaneus, 7 M. m. domesticus, 8 M. m. musculusTranscriptome sequencingX more highly differentiated than autosomes; higher differentiation in low-recombination regions in comparison between M. m. castaneus and M. m. musculus; some high-differentiation regions overlap with previously identified loci involved in hybrid male sterilityFST outliers Dxy outliers Numbers of unique, shared, and divergent polymorphismsNoDifferences in demographic history among subspecies may account for differences in divergence patterns among subspecies pairs; X chromosome contributes disproportionately to reproductive isolationPhifer-Rixey et al. (2014)
House mice2 Mus spretus 20 M. musculus domesticusSNP arrayPhylogenetic signals of introgression from M. spretus on a few chromosomes of each M. m. domesticus individualPhylogenetic networksYesAt least three hybridization events, including one 50 years ago connected to warfarin resistanceLiu et al. (2015)
Lousewart1 Pedicularis przewalskii, 1 P. cyathophylla, 1 P. superba, 2 P. cyathophylloides, 2 P. thamnophila, 5 P. rexRAD sequencing8.7–27% of P. thamnophila genome estimated to be derived from P. rexD-statisticsNoIntrogression among nearly all taxa in the P. rex – thamnophila cladeEaton & Ree (2013)
Manakins48 Manacus candei, 52 M. vitellinus, 104 putative hybridsGenotype by sequencingGene flow highly variable across genome; rapid decay of genomic autocorrelations; FST correlated with gene flow estimated in hybrids, but genomic outliers not strongly colocalizedFST outliers PCA Genomic clinesNoBarriers to gene flow caused by many alleles with small effects; some evidence that selection drives reproductive isolation, but isolation is not simply a consequence of adaptive changeParchman et al. (2013)
MosquitoesAnopheles gambiae, A. arabiensis, A. merusRAD sequencingExtensive shared polymorphism; higher divergence in X chromosome inversionsStructure PCA Dxy outliers FST outliersNoX chromosome inversions involved in reproductive isolationO'Loughlin et al. (2014)
MosquitoesAnopheles coluzzii, A. gambiae, A. arabiensis, A. quadriannulatus, A. melas, A. merusWhole-genome sequencingStrong phylogenetic discordance between X chromosome and autosomesPhylogenetic concordanceNoPhylogenetic relationships on the X chromosome represent the species history, revealing pervasive gene flow on the autosomesFontaine et al. (2014)
Oaks4 Quercus virginiana, 4 Q. geminate, 4 Q. minima, 3 Q. sagraeana, 5 Q. oleoides, 3 Q. brandegei, 4 Q. fusiformis, 7 outgroup samplesRAD sequencingSubstantial heterogeneity in the presence of admixture between species; six of seven species show evidence of admixture with one or more congeners; significant admixture limited to samples in close proximityStructure Treemix D-statistics dadiYesOak species boundaries and ranges have been evolutionarily stableEaton et al. (2015)
PigsS. scrofa, S. verrucosus, S. cebifronsWhole-genome sequencing23% of genome admixedBlockwise likelihoodYesAsymmetric hybridization occurring long before arrival of humansFrantz et al. (2014)
Rabbits6 Oryctolagus cuniculus cuniculus 6 Oryctolagus cuniculus algirus 1 Lepws timidusTargeted capture of intron sequence; transcriptome sequencingSubstantial fraction of shared polymorphisms (31.2%), with fewer fixed differences (0.3%); higher differentiation on the X chromosome and near centromeres; islands of differentiation highly variable in size, but usually contain small numbers of genesFST outliersNoMany fixed differences probably result from positive selection; complex genetic basis to reproductive isolationCarneiro et al. (2014)
Poplars7 genotypes each from Populus alba, P. tremulaRAD sequencingDivergence estimates strongly autocorrelated along chromosomes; many autocorrelations involve low divergence blocks, including incipient sex chromosomeFST outliers Genomic autocorrelationNoPopulus species have porous genomes and exhibit high levels of introgression across incipient sex chromosomeStölting et al. (2013)
Poplars498 Populus trichocarpa, 10 P. balsamiferaSNP arrayAdmixture limited to drainages where the species' ranges overlapAdmixture PCANoHybridization contributes to clinal variation in allele frequencies observed in P. trichocarpa.Geraldes et al. (2014)
Sunflowers40 H. annuus, 25 H. petiolaris, 28 H. argophyllus, 14 H. debilisTranscriptome sequencingHigher differentiation and more fixed differences in nonhybridizing species pairs than in hybridizing species pairs, despite closer phylogenetic relationship of the former; islands of genetic differentiation small and found in areas of low recombinationFST outliers Genomic autocorrelationNoInterspecific gene flow has less impact than local genome architecture on the clustering of divergent loci; complex genetic basis to reproductive isolationRenaut et al. (2013)
Sunflowers40 H. annuus, 28 H. argophyllusTranscriptome sequencingElevated FST and D within rearrangements and adjacent regions (within 5 cM of chromosomal break points)FST outliersNoChromosomal translocations and inversions reduce interspecific gene flow following secondary contactBarb et al. (2014)
Sunflowers100 genotypes representing 20 subpopulations of Helianthus petiolaris (strong intrinsic and extrinsic isolation between dune and nondune ecotypes); 20 individuals from other species or subspeciesRAD sequencingSplit between dune and nondune ecotypes estimated between 440–10 274 ya; Ongoing asymmetric introgression; Nem (Nondune to Dune) = 2.06; Nem (Dune to Nondune) = 5.53.FST outliers Structure MCMC fit to isolation with migrationYesPossible genomic signature of reinforcement: two FST outlier regions show greater divergence in comparisons between adjacent dune edge and nondune populations than between more distant core dune and nondune populationsAndrew et al. (2013); Andrew & Rieseberg (2013)
Swordtail fish60 Xiphophorus birchmanni, 60 X. malinche; 313 hybrids from two hybrid zonesGenotype by sequencingMany unlinked pairs of loci show linkage disequilibrium in hybrid populationsLinkage disequilibrium among unlinked lociYesA large number of hybrid incompatibilities contribute to reproductive isolationSchumer et al. (2014b)
Warblers (ring species)95 individuals from 22 populations (total) from Phylloscopus trochiloides viridanus, P. t. nitidus, P. t. ludlowi, P. t. trochiloides, P. t. obscuratus, P. t. plumbeitarsusGenotype by sequencingTwo northern forms are distinct genetic clusters connected by ring of genetically intergraded formsPCA StructureNoAllopatric divergence played a major role in the formation of ring species; recent hybridization at each area of secondary contactAlcaide et al. (2014)

Overall, available methods provide useful tools for reconstructing hybridization. The diversity of focal genomic patterns (Table 1) suggests that collectively, methods can access gene flow occurring over a range of timescales. For example, phylogenetic patterns among species can reveal older hybridization events, whereas the distribution of shared haplotype lengths is shaped by recent gene flow. Although no methods incorporate all aspects of genomic variation, many approaches adopt powerful likelihood or Bayesian frameworks that avoid drastic summaries of the data. For example, several strategies are available that analyse the complete (unsummarized), site frequency spectrum across multiple populations.

The ancestries of individuals provide special insights into hybridization history. Importantly, diverse strategies now exist for probabilistically assigning individual ancestry, assuming that appropriate reference populations can be surveyed (Gompert & Buerkle 2013). Since the Bayesian clustering approach Structure revolutionized the inference of ancestry proportions (Pritchard et al. 2000), methods have been developed that use the same likelihood (genotype data conditional on ancestry) but different algorithms to achieve substantial increases in computational speed (Tang et al. 2005; Alexander et al. 2009; Raj et al. 2014). With other methods, including HapMix (Price et al. 2009) and RASPberry (Wegmann et al. 2011), the locus-specific histories of individuals can be reconstructed, enabling changes in ancestry to be detected over short chromosomal distances.

Computational and statistical advances have significantly expanded the hybridization scenarios that may be considered. With approximate Bayesian computation, genomic data can be fit to any model that can be rapidly simulated, allowing (in principle) arbitrarily complex histories of gene flow to be compared. Existing methods also enable testing of specific hypotheses about reproductive barriers. For example, the genomic clines framework (Gompert & Buerkle 2009, 2011) statistically evaluates evidence for selection against gene flow at specific genomic locations through comparison to the genomewide hybrid index.

Despite their potential, available methods for characterizing hybridization suffer from several challenges when applied to genomic studies of speciation. First, options for measuring gene flow often exclude scenarios of interest. Among the subset of methods that estimate gene flow, most assume it happens continuously at an invariant rate. Few approaches reliably reconstruct the timing of hybridization. For example, popular methods based on the isolation with migration framework may produce wide confidence intervals on the timing of gene flow and falsely infer the presence of gene flow under some conditions (Sousa et al. 2011; Strasburg & Rieseberg 2011; Cruickshank & Hahn 2014; Hey et al. 2015). These restrictions do not match the reality of the speciation process, in which the opportunity for hybridization varies over time due to accumulating reproductive isolation, fluctuating geographic range sizes and other factors. Indeed, temporal changes in gene flow are key ingredients for distinguishing major models of speciation. For example, sympatric and para-patric speciation features gene flow during the earliest stages of divergence, whereas speciation by reinforcement is thought to be triggered by secondary contact and gene flow after a period of divergence in allopatry.

A second challenge with existing methods is the limited treatment of natural selection. Approaches that estimate the rate of gene flow usually assume that population history and neutral mutation account for genomic patterns. Therefore, selection has the potential to bias geneflow estimates. This problem could be severe for lineages with a history of reproductive isolation, especially if many loci contribute to isolation. A more fundamental methodological issue in the context of speciation is the assumption that one rate of gene flow characterizes the entire genome (Sousa et al. 2013). Because selection acting on hybrids targets specific loci, rates of introgression could vary significantly across the genome (Barton & Hewitt 1985; Harrison 1990; Payseur 2010); indeed, documenting this variation is a major goal of genomic studies of speciation. Alternatively, genome scan approaches search for locus-specific distortions in summary statistics, but usually leave unestimated rates of gene flow and associated selection coefficients. These shortcomings are not surprising as most methods were designed to analyse population structure within species rather than gene flow among diverging lineages. Although strategies specifically developed to analyse the genomic consequences of mixing in hybrid zones (e.g. geographic and genomic clines) enable the detection of selection and differential introgression in hybrids, they downplay the effects of demographic history and do not estimate locus-specific rates of gene flow.

The rate of meiotic recombination is another key parameter to consider in genomic studies of hybridization (Sousa & Hey 2013). Differential gene flow across a chromosome requires recombination. Recombination rate determines the chromosomal scale over which selection targeting reproductive isolation mutations reduces gene flow at neighbouring loci (Bazykin 1969; Barton 1979; Bengtsson 1985; Barton & Bengtsson 1986; Gavrilets 1997), the focal signature of genomic scan methods. Furthermore, several models of speciation propose that loci involved in reproductive isolation will preferentially accumulate in rearranged (Noor et al. 2001; Rieseberg 2001; Navarro & Barton 2003) or collinear (Butlin 2005) chromosomal regions with little recombination, allowing divergence to continue in the face of hybridization. Unfortunately, genomic methods for studying speciation ignore the reality that recombination rates vary across chromosomes on broad and fine scales (McVean et al. 2004; Kong et al. 2010; Comeron et al. 2012; Liu et al. 2014b). In addition, some methods assume that recombination is absent over short distances (i.e. within loci). As hybridization studies transition to using whole-genome sequences, another practical challenge emerges. How should the genome be partitioned into loci before geneflow analyses are conducted? For example, the typical genomic scan strategy of comparing windows of the same physical size (e.g. 10 kb) implicitly assumes that the rate of recombination is invariant. Most methods do not yet account for the expectation that switches in ancestry associated with hybridization will occur over finer scales in genomic regions with higher recombination rates.

Several approaches assume free recombination among polymorphisms, even when they are closely linked. A more general challenge arises from the treatment of unlinked loci as independent, a feature of all available methods. Loci on different chromosomes can behave nonindependently in scenarios of interest. For example, recent hybridization events involving only a few individuals can generate introgression at many loci (Goodman et al. 1999).

The role of geography in speciation has long been and continues to be a source of controversy (Mayr 1963; Coyne & Orr 2004; Seehausen et al. 2014). Knowledge of the geographic locations of past hybridization events is vital for understanding the connection between gene flow and speciation. With few exceptions (Yang et al. 2014), genomic methods for analysing hybridization do not consider geographic information nor do they infer the geographic context of gene flow.

Outcomes of hybridization may vary depending on whether the hybridizing populations diverged in the presence of gene flow or came into contact secondarily after a period of allopatry. But reliable methods for distinguishing between these scenarios are not readily available. Theoretical expectations for patterns of genome divergence following secondary contact or divergence with gene flow appear to be similar under equilibrium conditions (e.g. Barton 1979; Barton & Hewitt 1985; Charlesworth et al. 1997; Feder & Nosil 2010). However, given the extent of temporal variation in geographic ranges due to major climate oscillations, nonequilibrium conditions are likely to be common and could impact inferences about hybridization, as well as its consequences. For example, hybrid zones in the Arctic and some temperate regions are thought to have formed repeatedly following interglacial range expansions of hybridizing lineages, in some instances leading to introgression or hybrid speciation (Hewitt 2011).

In summary, researchers may choose from a variety of analytical approaches to draw inferences about hybridization from genomic data. Nevertheless, existing strategies are missing some components that are important in the context of speciation, suggesting that investment in further method development is warranted.

To identify emerging empirical patterns in the genomics of hybridization and speciation, we surveyed the literature. We focused on studies that: (i) articulated a specific interest in characterizing hybridization between diverging lineages (rather than simply scanning genomes for evidence of natural selection); (ii) measured variation across the genome at thousands to millions of loci; and (iii) examined species, subspecies or populations with independent evidence for reproductive isolation (rather than focusing on ecotypic differentiation). Both studies that sampled geographically separate populations and those that examined currently hybridizing populations were considered.

Inspection of the studies (listed in Table 2) reveals several themes. Genomic data confirm what is now conventional wisdom: hybridization between diverging lineages is common (Arnold 1997; Rieseberg 1997; Mallet 2005). Many instances of hybridization were previously unreported and unanticipated (e.g. genetic exchange between humans and Neanderthals), an indication of the potentially high power of genomic data to detect gene flow, even when it involves extinct populations. The history of hybridization among diverging lineages is sometimes complex, including multiple periods of gene flow as well as asymmetries in its direction. In some cases, genomic data raise the possibility that hybridization facilitated speciation (e.g. butterflies, cichlids). Nevertheless, the inference that most of the hybridizing lineages in Table 2 – including a ring species (Alcaide et al. 2014) – evolved in isolation for some period of time suggests that geographic separation contributes to reproductive isolation in many organismal lineages (Mayr 1963; Coyne & Orr 2004).

A consistent observation across surveyed taxa is rampant heterogeneity in patterns of differentiation and admixture within the genome, sometimes over short chromosomal distances. As discussed above, this patchwork reflects a combination of evolutionary factors, including incomplete lineage sorting, effects of selection before and after hybridization, and recombination rate variation, as well as gene flow and selection targeting hybrids. That heterogeneity occurs over a fine genomic scale suggests that some aspects of hybridization can only be captured by examining the entire genome. In those studies that sample hybrid zones, this heterogeneity raises the prospect that boundaries between nascent species are semipermeable, with a subset of loci maintaining reproductive barriers (Harrison 1986, 1990; Wu 2001). In those studies that sample geographically separate populations, this conclusion is more difficult to reach: the null model of a low rate of gene flow that is constant across the genome could still produce strong heterogeneity if only a few regions introgress. From a phylogenetic perspective, the genomes of recently diverged species are complex mosaics of alternating histories, highlighting the difficulty of reconstructing species relationships, even with genomic data. Indeed, extensive hybridization challenges basic assumptions of phylogenetic methods, suggesting that some species histories are better represented as phylogenetic networks (Yu et al. 2014).

Genomic regions that display high differentiation (in allopatric populations) or narrow clines (in hybrid zones) could contain genetic changes that confer reproductive isolation. Under this assumption, the studies in Table 2 collectively raise a few genomic themes that could characterize reproductive barriers between species. First, the number of loci that appear to maintain differentiation in the face of gene flow is usually high, suggesting a complex genetic basis to reproductive isolation. An alternative explanation for this pattern is that many loci are falsely labelled as ‘genomic outliers’ because significance thresholds are based on null models that ignore or mischaracterize demographic history.

A second observation is that although outlying loci are scattered throughout the genome, there is evidence for a bias towards regions of low recombination. In several species, this pattern takes the form of higher differentiation near centromeres, which are known or suspected to recombine less. In others, the relationship between outlier location and recombination rate is more obvious. Cline width and recombination rate are positively correlated across the genome in a hybrid zone between two subspecies of house mice (Janoušek et al. 2015). Selection to remove long genomic blocks from one species in hybrids was postulated to explain a negative correlation between absolute divergence and recombination rate in monkey flowers (Brandvain et al. 2014). Rearranged chromosomal regions exhibit higher differentiation than collinear regions in several species pairs. Perhaps these results indicate a role for suppressed recombination in the evolution of reproductive isolation. Alternatively, they could simply reflect increased power to detect isolation loci using linked markers in low-recombination regions. In some cases, the pattern appears to be caused by selection at linked sites within lineages rather than barriers to gene flow (Cruickshank & Hahn 2014; Burri et al. 2015). Regardless, these results reveal the importance of considering variation in local recombination rate when interpreting genomic patterns of differentiation and admixture.

A third trend among surveyed species is the over-representation of highly differentiated loci on the sex chromosomes relative to the autosomes, seen in both X-Y and Z-W systems. This pattern agrees with results from controlled crosses in the laboratory, which often associate isolation phenotypes – especially hybrid sterility and hybrid inviability – with the sex chromosomes (Coyne & Orr 2004). Whether this genomic tendency indicates that the sex chromosomes harbour a higher density of loci involved in reproductive isolation (Masly & Presgraves 2007), loci with stronger phenotypic effects on isolation, or both, remains to be seen. Interpretations of this pattern must also recognize the complications inherent in setting neutral expectations for the sex chromosomes, which can experience different rates of genetic drift and migration than the autosomes.

The collection of empirical studies (Table 2) reveals considerable interest in using genomic data to understand the role of natural selection in speciation. The degree to which selection can increase divergence in the face of gene flow remains controversial and difficult to determine (Kirkpatrick & Barton 2006; Via 2012; Andrew & Rieseberg 2013; Flaxman et al. 2013, 2014; Yeaman 2013; Cruickshank & Hahn 2014; Feder et al. 2014; Seehausen et al. 2014). In contrast, the idea that reproductive isolation is a by-product of local adaptation in allopatry is widely accepted, but genetic tests of this hypothesis remain rare. Those genomic studies that jointly consider currently allopatric populations and currently hybridizing populations provide a special opportunity to evaluate this influential prediction. If the genetic changes that block gene flow are initially driven to fixation by positive selection, genomic regions with restricted introgression in hybrid zones could display higher differentiation between nascent species. Genomic cline width and differentiation are negatively correlated in lycaenid butterflies and in manakins, suggesting that adaptive evolution contributes to reproductive isolation (Gompert et al. 2012; Parchman et al. 2013). Nevertheless, the weakness of these correlations leaves open the possibility that genetic drift plays an important role in the establishment of reproductive barriers.

Despite the trends described above, the genomic landscape of divergence appears to differ substantially among species pairs. Estimated hybridization rates vary widely. Whereas some species show evidence for extensive gene flow with a few genomic regions maintaining divergence (e.g. rabbits, mosquitoes), others show widespread differentiation with limited pockets of introgression (e.g. Drosophila simulans vs. island endemics). The genomic scale of heterogeneity also varies. Although these comparisons are complicated by contrasting study designs, some observed disparities probably reflect biological differences among species pairs in the timing and mode of speciation.

This empirical survey also reveals challenges that face genomic studies of hybridization. From the perspective of speciation, a primary motivation for analysing genomic data is its potential for characterizing the magnitude and timing of gene flow among diverging lineages. Only a handful of studies (Table 2) reported estimates of these important quantities. The most widely used measure of locus-specific differentiation was FST, a summary statistic that compares variation between populations to variation within populations (Wright 1951). Although FST enables comparisons of loci that differ in within-population diversity, it suffers from interpretive challenges when used to measure gene flow across the genome. Reduced gene flow, which maintains between-population divergence, is confounded with selection at linked sites, which decreases within-population diversity (Charlesworth 1998; Noor & Bennett 2009; Nachman & Payseur 2012; Cruickshank & Hahn 2014). Furthermore, most studies identified unusually differentiated loci as those with values that simply fell in the tail of the genomic distribution, a strategy complicated by the fact that completely neutral genomic distributions also have tails. Other studies used neutral simulations to construct distributions of summary statistics expected in the absence of selection. This approach ignores the likely possibility that selection against gene flow (and other forms of selection) occurs at many genomic locations in hybrids.

Furthermore, Table 2 reveals biases in sampling strategy. The majority of studies examined populations that are currently allopatric, with no a priori evidence of hybridization. Compared to hybrid zones, allopatric populations are usually easier to find and investigators may be more likely to sample them for purposes other than studying gene flow. Nevertheless, connecting genomic patterns with hybridization is much simpler in populations that are currently hybridizing in visible ways (e.g. hybrid zones). Other biases are taxonomic. Animal species – especially vertebrates – were the most popular subjects. It is too early to identify differences between plants and animals from genomic studies of hybridization. But the presence of heteromorphic sex chromosomes seems to alter the genetic architecture of reproductive barriers and to increase the speed with which they accumulate, and most plants lack heteromorphic sex chromosomes (Rieseberg 2001; Coyne & Orr 2004; Levin 2012; Phillips & Edmands 2012; Lima 2014). Likewise, it has long been known that the frequency of hybridization tends to be higher in species with external fertilization, such as plants and fishes, than in taxa with internal fertilization (Stebbins 1959). These and other related observations should motivate genomic studies of hybridization in organismal groups that differ in the genetics of sex determination and/or other key features of reproductive biology.

Our overview of methods and empirical studies suggests fruitful directions for research on the genomics of hybridization and speciation. First, model-based estimation of the magnitude and timing of hybridization should be prioritized. Reconstructing the history of gene flow between diverging lineages is a necessary step towards understanding the causes and consequences of speciation, and it may provide key insights. For example, the alternative scenarios of speciation with gene flow and allopatric speciation followed by hybridization during secondary contact could be distinguished by quantifying the dynamics of gene flow. Accurate estimates of the magnitude and timing of hybridization are also needed to interpret the results from scans that seek to identify genomic regions with restricted gene flow, which are growing in popularity. Because the expected genomic distribution of differentiation depends on the level and timing of hybridization, the determination that certain loci experienced reduced gene flow ultimately requires a demographic model including these parameters. At present, some authors seem to assume that evidence of hybridization elsewhere in the genome is sufficient to demonstrate that a differentiated genomic region is resistant to gene flow. But a wide spectrum of differentiation among loci is expected, even in the absence of reproductive barriers.

Second, natural selection should be directly incorporated into methods for characterizing hybridization between diverging lineages. Even partial reproductive isolation can generate strong selection against hybrids that shapes variation throughout the genome (Barton 1983; Barton & Bengtsson 1986). Accounting for the effects of selection is therefore a necessary step towards accurately reconstructing the history of hybridization. Furthermore, model-based characterization of selection could provide insights into genetic mechanisms of speciation. By estimating the contributions of individual loci to hybrid fitness, the genetic architecture of reproductive barriers in nature, as well as the identity of genomic regions and candidate genes that underlie adaptive introgression or heterosis, could be revealed. It may be possible to compare the fit of observed genomic patterns to models featuring different architectures. For example, selection against heterozygotes at single loci is predicted to impede gene flow at linked variants more strongly than epistatic selection against two-locus, hetero-specific genotypes (‘hybrid incompatibilities’) (Gavrilets 2004; Gompert & Buerkle 2011; Lindtke & Buerkle 2015). The genomic clines approach (Gompert & Buerkle 2009, 2011) provides a good foundation for incorporating selection into genomic studies of hybridization in the context of hybrid zones. For currently allopatric populations that hybridized in the past, model-based assignment of loci to one of a few classes with different rates of gene flow (due to selection) (Sousa et al. 2013) may offer a promising start towards the construction of more realistic methods.

Building geographic information into genomic methods for analysing hybridization is another worthwhile goal. This strategy would recognize that the prior probabilities associated with past hybridization are functions of current geographic ranges. It would also enable tests of speciation models that depend on geography. For example, reinforcement is expected to generate ‘inverse clines’ in which greater divergence for the trait/loci under reinforcing selection is expected between populations at the centre than at the ends of the cline (Caisse & Antonovics 1978; Andrew & Rieseberg 2013). Finally, incorporating geographic information could help identify the selective forces responsible for initial divergence and the formation of reproductive isolation, an important goal (Shaw & Mullen 2011). In this regard, clinal models of adaptation with gene flow (Endler 1977; Novembre & Di Rienzo 2009) could provide useful guidelines.

Even without these methodological advances, authors of empirical studies should be explicit about the models they are considering and the quantitative patterns they expect to find. In general, it is important to bear in mind that hybrid genomes are outcomes of a complex mixture of processes. As a result, intuition about expected patterns will sometimes mislead. Studies that embrace the complications inherent in the speciation process – rather than automatically attributing observed genomic patterns to differential gene flow generated by hybridization – should be prioritized.

Our survey suggests other guidelines for genomic studies of hybridization and speciation. Reproductive isolation should be directly measured whenever possible. Genomic comparisons alone are unlikely to identify genes responsible for reproductive barriers. Understanding how genetic variants restrict gene flow will ultimately require characterizing their functional effects on specific isolation phenotypes. More broadly, the interpretation of genomic patterns will be enriched by comparison to knowledge of reproductive barriers. For example, the over-representation of sex chromosome loci in genetic dissections of isolation traits provides a potential explanation for higher differentiation of sex-linked loci in genomic studies of hybridization. Likewise, a recent critique of the evidence for homoploid hybrid speciation showed that while admixture was well-documented in most putative hybrid species, only a handful of studies connected hybridization to reproductive barrier formation – a key criterion for hybrid speciation (Schumer et al. 2014a).

Genomic studies of hybridization should stimulate the development of new organismal models for speciation. Most information about mechanisms of speciation still comes from a small subset of species, but Table 2 suggests that the dynamics and outcomes of hybridization could differ significantly across taxa. Examining those systems featuring multiple species pairs that collectively sample a range of divergence times will enable the effects of hybridization to be evaluated at different stages of speciation. Furthermore, species that currently hybridize offer the most direct insights into the genomic consequences of hybridization, indicating that these species should be prioritized in speciation studies.

Finally, the limits to inferring hybridization from genomic data should be examined and embraced. The space of potential hybridization scenarios is expansive and which subset of historical information is recorded in genomes is highly stochastic. Determining what we can realistically hope to learn about speciation from genomic data – rather than treating it as a panacea – is an important next step.

We thank Nick Barton, Jeff Good and Richard Abbott for organizing this special issue of Molecular Ecology and for inviting us to contribute. We thank Nick Barton and three anonymous reviewers for insightful comments that improved the manuscript. BAP's work on hybridization is supported by NSF grant DEB 1353737. LHR's work on hybridization is supported by an NSERC Discovery grant.

BAP developed the outline of the paper, conducted most of the literature survey, and completed the bulk of the writing. LHR critiqued the outline, contributed to the writing, and helped with the literature survey, especially with respect to plant hybridization.

  • Alcaide M, Scordato ESC, Price TD, Irwin DE. Genomic divergence in a ring species complex. Nature. 2014;511:83–85. [PubMed] [Google Scholar]
  • Alexander DH, Novembre J, Lange K. Fast model-based estimation of ancestry in unrelated individuals. Genome Research. 2009;19:1655–1664. [PMC free article] [PubMed] [Google Scholar]
  • Altmann A, Weber P, Bader D, et al. A beginners guide to SNP calling from high-throughput DNA-sequencing data. Human Genetics. 2012;131:1541–1554. [PubMed] [Google Scholar]
  • Andrew RL, Rieseberg LH. Divergence is focused on few genomic regions early in speciation: incipient speciation of sunflower ecotypes. Evolution. 2013;67:2468–2482. [PubMed] [Google Scholar]
  • Andrew RL, Kane NC, Baute GJ, Grassa CJ, Rieseberg LH. Recent nonhybrid origin of sunflower ecotypes in a novel habitat. Molecular Ecology. 2013;22:799–813. [PubMed] [Google Scholar]
  • Ane C, Larget B, Baum DA, Smith SD, Rokas A. Bayesian estimation of concordance among gene trees. Molecular Biology and Evolution. 2007;24:412–426. [PubMed] [Google Scholar]
  • Arnold ML. Natural Hybridization and Evolution. Oxford University Press; Oxford: 1997. [Google Scholar]
  • Arnold B, Corbett-Detig RB, Hartl D, Bomblies K. RAD-seq underestimates diversity and introduces genealogical biases due to nonrandom haplotype sampling. Molecular Ecology. 2013;22:3179–3190. [PubMed] [Google Scholar]
  • Barb JG, Bowers JE, Renaut S, et al. Chromosomal evolution and patterns of introgression in Helianthus. Genetics. 2014;197:969–979. [PMC free article] [PubMed] [Google Scholar]
  • Barton NH. Gene flow past a cline. Heredity. 1979;43:333–339. [Google Scholar]
  • Barton NH. Multi-locus clines. Evolution. 1983;37:454–471. [PubMed] [Google Scholar]
  • Barton N, Bengtsson BO. The barrier to genetic exchange between hybridising populations. Heredity. 1986;57:357–376. [PubMed] [Google Scholar]
  • Barton NH, Hewitt GM. Analysis of hybrid zones. Annual Review of Ecology and Systematics. 1985;16:113–148. [Google Scholar]
  • Bazykin AD. Hypothetical mechanism of speciation. Evolution. 1969;23:685–687. [PubMed] [Google Scholar]
  • Beaumont MA, Balding DJ. Identifying adaptive genetic divergence among populations from genome scans. Molecular Ecology. 2004;13:969–980. [PubMed] [Google Scholar]
  • Beaumont MA, Zhang W, Balding DJ. Approximate Bayesian computation in population genetics. Genetics. 2002;162:2025–2035. [PMC free article] [PubMed] [Google Scholar]
  • Becquet C, Przeworski M. A new approach to estimate parameters of speciation models with application to apes. Genome Research. 2007;17:1505–1519. [PMC free article] [PubMed] [Google Scholar]
  • Bengtsson BO. The flow of genes through a genetic barrier. In: Greenwood PJ, Slatkin M, editors. Evolution: Essays in Honor of John Maynard Smith. Cambridge University Press; Cambridge: 1985. pp. 31–42. [Google Scholar]
  • Brandvain Y, Kenney AM, Flagel L, Coop G, Sweigart AL. Speciation and introgression between Mimulus nasutus and Mimulus guttatus. PLoS Genetics. 2014;10:e1004410. [PMC free article] [PubMed] [Google Scholar]
  • Buerkle C, Gompert Z. Population genomics based on low coverage sequencing: how low should we go? Molecular Ecology. 2013;22:3028–3035. [PubMed] [Google Scholar]
  • Burri R, Nater A, Kawakami T, et al. Linked selection and recombination rate variation drive the evolution of the genomic landscape of differentiation across the speciation continuum of Ficedula flycatchers. Genome Research. 2015;25:1656–1665. [PMC free article] [PubMed] [Google Scholar]
  • Butlin RK. Recombination and speciation. Molecular Ecology. 2005;14:2621–2635. [PubMed] [Google Scholar]
  • Caisse M, Antonovics J. Evolution in closely adjacent plant populations. IX. Evolution of reproductive isolation in clinal populations. Heredity. 1978;40:371–384. [Google Scholar]
  • Carneiro M, Albert FW, Afonso S, et al. The genomic architecture of population divergence between subspecies of the European rabbit. PLoS Genetics. 2014;10:e1003519. [PMC free article] [PubMed] [Google Scholar]
  • Charlesworth B. Measures of divergence between populations and the effect of forces that reduce variability. Molecular Biology and Evolution. 1998;15:538–543. [PubMed] [Google Scholar]
  • Charlesworth B, Nordborg M, Charlesworth D. The effects of local selection, balanced polymorphism and background selection on equilibrium patterns of genetic diversity in subdivided populations. Genetical Research. 1997;70:155–174. [PubMed] [Google Scholar]
  • Clark AG. Neutral behavior of shared polymorphism. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:7730–7734. [PMC free article] [PubMed] [Google Scholar]
  • Comeron JM, Ratnappan R, Bailin S. The many landscapes of recombination in Drosophila melanogaster. PLoS Genetics. 2012;8:e1002905. [PMC free article] [PubMed] [Google Scholar]
  • Coyne JA, Orr HA. Speciation. Sinauer Associates, Sunderland; Massachusetts: 2004. [Google Scholar]
  • Cruickshank TE, Hahn MW. Reanalysis suggests that genomic islands of speciation are due to reduced diversity, not reduced gene flow. Molecular Ecology. 2014;23:3133–3157. [PubMed] [Google Scholar]
  • Dobzhansky T. Genetics and the Origin of Species. Columbia University Press; New York, New York: 1937. [Google Scholar]
  • Durand EY, Patterson N, Reich D, Slatkin M. Testing for ancient admixture between closely related populations. Molecular Biology and Evolution. 2011;28:2239–2252. [PMC free article] [PubMed] [Google Scholar]
  • Eaton DAR, Ree RH. Inferring phylogeny and introgression using RADseq data: an example from flowering plants (Pedicularis: Orobanchaceae) Systematic Biology. 2013;62:689–706. [PMC free article] [PubMed] [Google Scholar]
  • Eaton DAR, Hipp AL, González-Rodríguez A, Cavender-Bares J. Historical introgression among the American live oaks and the comparative nature of tests for introgression. Evolution. 2015;69:2587–2601. [PubMed] [Google Scholar]
  • Ellegren H, Smeds L, Burri R, et al. The genomic landscape of species divergence in Ficedula flycatchers. Nature. 2012;491:756–760. [PubMed] [Google Scholar]
  • Endler JA. Geographic Variation, Speciation, and Clines. Princeton University Press; Princeton, New Jersey: 1977. [PubMed] [Google Scholar]
  • Falush D, Stephens M, Pritchard JK. Inference of population structure using multilocus genotype data: linked loci and correlated allele frequencies. Genetics. 2003;164:1567–1587. [PMC free article] [PubMed] [Google Scholar]
  • Feder JL, Nosil P. The efficacy of divergence hitchhiking in generating genomic islands during ecological speciation. Evolution. 2010;64:1729–1747. [PubMed] [Google Scholar]
  • Feder JL, Nosil P, Wacholder AC, et al. Genome-wide congealing and rapid transitions across the speciation continuum during speciation with gene flow. The Journal of Heredity. 2014;105(Suppl):810–820. [PubMed] [Google Scholar]
  • Fitzpatrick BM. Alternative forms for genomic clines. Ecology and Evolution. 2013;3:1951–1966. [PMC free article] [PubMed] [Google Scholar]
  • Flaxman SM, Feder JL, Nosil P. Genetic hitchhiking and the dynamic buildup of genomic divergence during speciation with gene flow. Evolution. 2013;67:2577–2591. [PubMed] [Google Scholar]
  • Flaxman SM, Wacholder AC, Feder JL, Nosil P. Theoretical models of the influence of genomic architecture on the dynamics of speciation. Molecular Ecology. 2014;23:4074–4088. [PubMed] [Google Scholar]
  • Fontaine MC, Pease JB, Steele A, et al. Extensive introgression in a malaria vector species complex revealed by phylogenomics. Science. 2014;347:1258524. [PMC free article] [PubMed] [Google Scholar]
  • Frantz LAF, Madsen O, Megens HJ, Groenen MAM, Lohse K. Testing models of speciation from genome sequences: divergence and asymmetric admixture in Island South-East Asian Sus species during the Plio-Pleistocene climatic fluctuations. Molecular Ecology. 2014;23:5566–5574. [PMC free article] [PubMed] [Google Scholar]
  • Garrigan D, Kingan SB, Geneva AJ, et al. Genome sequencing reveals complex speciation in the Drosophila simulans clade. Genome Research. 2012;22:1499–1511. [PMC free article] [PubMed] [Google Scholar]
  • Gavrilets S. Hybrid zones with Dobzhansky-type epistatic selection. Evolution. 1997;51:1027–1035. [PubMed] [Google Scholar]
  • Gavrilets S. Fitness Landscapes and the Origin of Species. Princeton University Press; Princeton, New Jersey: 2004. [Google Scholar]
  • Geraldes A, Farzaneh N, Grassa CJ, et al. Landscape genomics of Populus trichocarpa: the role of hybridization, limited gene flow, and natural selection in shaping patterns of population structure. Evolution. 2014;68:3260–3280. [PubMed] [Google Scholar]
  • Gompert Z, Buerkle CA. A powerful regression-based method for admixture mapping of isolation across the genome of hybrids. Molecular Ecology. 2009;18:1207–1224. [PubMed] [Google Scholar]
  • Gompert Z, Buerkle CA. Bayesian estimation of genomic clines. Molecular Ecology. 2011;20:2111–2127. [PubMed] [Google Scholar]
  • Gompert Z, Buerkle CA. Analyses of genetic ancestry enable key insights for molecular ecology. Molecular Ecology. 2013;22:5278–5294. [PubMed] [Google Scholar]
  • Gompert Z, Lucas LK, Nice CC, et al. Genomic regions with a history of divergent selection affect fitness of hybrids between two butterfly species. Evolution. 2012;66:2167–2181. [PubMed] [Google Scholar]
  • Good JM, Vanderpool D, Keeble S, Bi K. Negligible nuclear introgression despite complete mitochondrial capture between two species of chipmunks. Evolution. 2015;69:1961–1972. [PubMed] [Google Scholar]
  • Goodman SJ, Barton NH, Swanson G, Abernathy K, Pemberton JM. Introgression through rare hybridization: a genetic study of a hybrid zone between red and sika deer (genus Cervus) in Argyll, Scotland. Genetics. 1999;152:355–371. [PMC free article] [PubMed] [Google Scholar]
  • Gravel S. Population genetics models of local ancestry. Genetics. 2012;191:607–619. [PMC free article] [PubMed] [Google Scholar]
  • Green RE, Krause J, Briggs AW, et al. A draft sequence of the Neandertal genome. Science. 2010;328:710–722. [PMC free article] [PubMed] [Google Scholar]
  • Gutenkunst RN, Hernandez RD, Williamson SH, Bustamante CD. Inferring the joint demographic history of multiple populations from multidimensional SNP frequency data. PLoS Genetics. 2009;5:e1000695. [PMC free article] [PubMed] [Google Scholar]
  • Harris K, Nielsen R. Inferring demographic history from a spectrum of shared haplotype lengths. PLoS Genetics. 2013;9:e1003521. [PMC free article] [PubMed] [Google Scholar]
  • Harrison RG. Pattern and process in a narrow hybrid zone. Heredity. 1986;56:337–349. [Google Scholar]
  • Harrison RG. Hybrid zones: windows on evolutionary processes. Oxford Surveys in Evolutionary Biology. 1990;7:69–128. [Google Scholar]
  • Heliconius Genome Consortium. Butterfly genome reveals promiscuous exchange of mimicry adaptations among species. Nature. 2012;487:94–98. [PMC free article] [PubMed] [Google Scholar]
  • Hewitt GM. Quaternary phylogeography: the roots of hybrid zones. Genetica. 2011;139:617–638. [PubMed] [Google Scholar]
  • Hey J. The divergence of chimpanzee species and subspecies as revealed in multipopulation isolation-with-migration analyses. Molecular Biology and Evolution. 2010;27:921–933. [PMC free article] [PubMed] [Google Scholar]
  • Hey J, Nielsen R. Multilocus methods for estimating population sizes, migration rates and divergence time, with applications to the divergence of Drosophila pseudoobscura and D-persimilis. Genetics. 2004;167:747–760. [PMC free article] [PubMed] [Google Scholar]
  • Hey J, Nielsen R. Integration within the Felsenstein equation for improved Markov chain Monte Carlo methods in population genetics. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:2785–2790. [PMC free article] [PubMed] [Google Scholar]
  • Hey J, Chung Y, Sethuraman A. On the occurrence of false positives in tests of migration under an isolation-with-migration model. Molecular Ecology. 2015;24:5078–5083. [PMC free article] [PubMed] [Google Scholar]
  • Hohenlohe PA, Day MD, Amish SJ, et al. Genomic patterns of introgression in rainbow and westslope cutthroat trout illuminated by overlapping paired-end RAD sequencing. Molecular Ecology. 2013;22:3002–3013. [PMC free article] [PubMed] [Google Scholar]
  • Hubisz MJ, Falush D, Stephens M, Pritchard JK. Inferring weak population structure with the assistance of sample group information. Molecular Ecology Resources. 2009;9:1322–1332. [PMC free article] [PubMed] [Google Scholar]
  • Huelsenbeck JP, Andolfatto P. Inference of population structure under a Dirichlet process model. Genetics. 2007;175:1787–1802. [PMC free article] [PubMed] [Google Scholar]
  • Janoušek V, Wang L, Luzynski K, et al. Genome-wide architecture of reproductive isolation in a naturally occurring hybrid zone between Mus musculus musculus and M. m. domesticus. Molecular Ecology. 2012;21:3032–3047. [PMC free article] [PubMed] [Google Scholar]
  • Janoušek V, Munclinger P, Wang L, Teeter KC, Tucker PK. Functional organization of the genome may shape the species boundary in the house mouse. Molecular Biology and Evolution. 2015;32:1208–1220. [PMC free article] [PubMed] [Google Scholar]
  • Jónsson H, Schubert M, Seguin-Orlando A, et al. Speciation with gene flow in equids despite extensive chromosomal plasticity. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:18655–18660. [PMC free article] [PubMed] [Google Scholar]
  • Keller I, Wagner CE, Greuter L, et al. Population genomic signatures of divergent adaptation, gene flow and hybrid speciation in the rapid radiation of Lake Victoria cichlid fishes. Molecular Ecology. 2013;22:2848–2863. [PubMed] [Google Scholar]
  • Kirkpatrick M, Barton N. Chromosome inversions, local adaptation and speciation. Genetics. 2006;173:419–434. [PMC free article] [PubMed] [Google Scholar]
  • Kong A, Thorleifsson G, Gudbjartsson DF, et al. Fine-scale recombination rate differences between sexes, populations and individuals. Nature. 2010;467:1099–1103. [PubMed] [Google Scholar]
  • Kulathinal RJ, Stevison LS, Noor MAF. The genomics of speciation in Drosophila: diversity, divergence, and introgression estimated using low-coverage genome sequencing. PLoS Genetics. 2009;5:e1000550. [PMC free article] [PubMed] [Google Scholar]
  • Lawson DJ, Hellenthal G, Myers S, Falush D. Inference of population structure using dense haplotype data. PLoS Genetics. 2012;8:e1002453. [PMC free article] [PubMed] [Google Scholar]
  • Levin DA. The long wait for hybrid sterility in flowering plants. New Phytologist. 2012;196:666–670. [PubMed] [Google Scholar]
  • Lima TG. Higher levels of sex chromosome heteromorphism are associated with markedly stronger reproductive isolation. Nature Communications. 2014;5:4743. [PubMed] [Google Scholar]
  • Lindtke D, Buerkle CA. The genetic architecture of hybrid incompatibilities and their effect on barriers to introgression in secondary contact. Evolution. 2015;69:1987–2004. [PubMed] [Google Scholar]
  • Liu KJ, Dai J, Truong K, et al. An HMM-based comparative genomic framework for detecting introgression in eukaryotes. In: Lengauer T, editor. PLoS Computational Biology. Vol. 10. 2014a. p. e1003649. [PMC free article] [PubMed] [Google Scholar]
  • Liu EY, Morgan AP, Chesler EJ, et al. High-resolution sex-specific linkage maps of the mouse reveal polarized distribution of crossovers in male germline. Genetics. 2014b;197:91–106. [PMC free article] [PubMed] [Google Scholar]
  • Liu KJ, Steinberg E, Yozzo A, et al. Interspecific introgressive origin of genomic diversity in the house mouse. Proceedings of the National Academy of Sciences of the United States of America. 2015;112:196–201. [PMC free article] [PubMed] [Google Scholar]
  • Lohse K, Frantz LA. Neandertal admixture in Eurasis confirmed by maximum-likelihood analysis of three genomes. Genetics. 2014;196:1241–1251. [PMC free article] [PubMed] [Google Scholar]
  • Lohse K, Harrison RJ, Barton NH. A general method for calculating likelihoods under the coalescent process. Genetics. 2011;189:977–987. [PMC free article] [PubMed] [Google Scholar]
  • Lohse K, Clarke M, Ritchie MG, Etges WJ. Genome-wide tests for introgression between cactophilic Drosophila implicate a role of inversions during speciation. Evolution. 2015;69:1178–1190. [PMC free article] [PubMed] [Google Scholar]
  • Lohse K, Chmelik M, Martin SH, Barton NH. Efficient strategies for calculating blockwise likelihoods under the coalescent. Genetics. 2016;202:775–786. [PMC free article] [PubMed] [Google Scholar]
  • Mailund T, Halager AE, Westergaard M, et al. A new isolation with migration model along complete genomes infers very different divergence processes among closely related great ape species. PLoS Genetics. 2012;8:e1003125. [PMC free article] [PubMed] [Google Scholar]
  • Mallet J. Hybridization as an invasion of the genome. Trends in Ecology and Evolution. 2005;20:229–237. [PubMed] [Google Scholar]
  • Mardis ER. Next-generation sequencing platforms. Annual Review of Analytical Chemistry. 2013;6:287–303. [PubMed] [Google Scholar]
  • Martin SH, Dasmahapatra KK, Nadeau NJ, et al. Genome-wide evidence for speciation with gene flow in Heliconius butterflies. Genome Research. 2013;23:1817–1828. [PMC free article] [PubMed] [Google Scholar]
  • Martin CH, Cutler JS, Friel JP, et al. Complex histories of repeated gene flow in Cameroon crater lake cichlids cast doubt on one of the clearest examples of sympatric speciation. Evolution. 2015a;69:1406–1422. [PubMed] [Google Scholar]
  • Martin SH, Davey JW, Jiggins CD. Evaluating the use of ABBA-BABA statistics to locate introgressed loci. Molecular Biology and Evolution. 2015b;32:244–257. [PMC free article] [PubMed] [Google Scholar]
  • Masly JP, Presgraves DC. High-resolution genome-wide dissection of the two rules of speciation in Drosophila. PLoS Biology. 2007;5:e243. [PMC free article] [PubMed] [Google Scholar]
  • Mayr E. Animal Species and Evolution. Belknap Press; Cambridge, Massachusetts: 1963. [Google Scholar]
  • McVean GA, Myers SR, Hunt S, et al. The fine-scale structure of recombination rate variation in the human genome. Science. 2004;304:581–584. [PubMed] [Google Scholar]
  • Meng C, Kubatko LS. Detecting hybrid speciation in the presence of incomplete lineage sorting using gene tree incongruence: a model. Theoretical Population Biology. 2009;75:35–45. [PubMed] [Google Scholar]
  • Nachman MW, Payseur BA. Recombination rate variation and speciation: theoretical predictions and empirical results from rabbits and mice. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences. 2012;367:409–421. [PMC free article] [PubMed] [Google Scholar]
  • Nadachowska-Brzyska K, Burri R, Olason PI, et al. Demographic divergence history of pied flycatcher and collared flycatcher inferred from whole-genome re-sequencing data. PLoS Genetics. 2013;9:e1003942. [PMC free article] [PubMed] [Google Scholar]
  • Nadeau NJ, Martin SH, Kozak KM, et al. Genome-wide patterns of divergence and gene flow across a butterfly radiation. Molecular Ecology. 2013;22:814–826. [PubMed] [Google Scholar]
  • Narum SR, Buerkle CA, Davey JW, Miller MR, Hohenlohe PA. Genotyping-by-sequencing in ecological and conservation genomics. Molecular Ecology. 2013;22:2841–2847. [PMC free article] [PubMed] [Google Scholar]
  • Navarro A, Barton NH. Accumulating postzygotic isolation genes in parapatry: a new twist on chromosomal speciation. Evolution. 2003;57:447–459. [PubMed] [Google Scholar]
  • Nielsen R, Paul JS, Albrechtsen A, Song YS. Genotype and SNP calling from next-generation sequencing data. Nature Reviews Genetics. 2011;12:443–451. [PMC free article] [PubMed] [Google Scholar]
  • Noor MAF, Bennett SM. Islands of speciation or mirages in the desert? Examining the role of restricted recombination in maintaining species. Heredity. 2009;103:439–444. [PMC free article] [PubMed] [Google Scholar]
  • Noor MAF, Grams KL, Bertucci LA, Reiland J. Chromosomal inversions and the reproductive isolation of species. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:12084–12088. [PMC free article] [PubMed] [Google Scholar]
  • Nordborg M. Structured coalescent processes on different time scales. Genetics. 1997;146:1501–1514. [PMC free article] [PubMed] [Google Scholar]
  • Novembre J, Di Rienzo A. Spatial patterns of variation due to natural selection in humans. Nature Reviews Genetics. 2009;10:745–755. [PMC free article] [PubMed] [Google Scholar]
  • O'Loughlin SM, Magesa S, Mbogo C, et al. Genomic analyses of three malaria vectors reveals extensive shared polymorphism but contrasting population histories. Molecular Biology and Evolution. 2014;31:889–902. [PMC free article] [PubMed] [Google Scholar]
  • Olson ND, Lund SP, Colman RE, et al. Best practices for evaluating single nucleotide variant calling methods for microbial genomics. Frontiers in Genetics. 2015;6:235. [PMC free article] [PubMed] [Google Scholar]
  • Parchman TL, Gompert Z, Braun MJ, et al. The genomic consequences of adaptive divergence and reproductive isolation between species of manakins. Molecular Ecology. 2013;22:3304–3317. [PubMed] [Google Scholar]
  • Patterson NJ, Price AL, Reich D. Population structure and eigenanalysis. Plos Genetics. 2006;2:e190. [PMC free article] [PubMed] [Google Scholar]
  • Patterson N, Moorjani P, Luo Y, et al. Ancient admixture in human history. Genetics. 2012;192:1065–1093. [PMC free article] [PubMed] [Google Scholar]
  • Payseur BA. Using differential introgression in hybrid zones to identify genomic regions involved in speciation. Molecular Ecology Resources. 2010;10:806–820. [PubMed] [Google Scholar]
  • Pease JB, Hahn MW. Detection and polarization of introgression in a five-taxon phylogeny. Systematic Biology. 2015;64:651–662. [PubMed] [Google Scholar]
  • Phifer-Rixey M, Bomhoff M, Nachman MW. Genome-wide patterns of differentiation among house mouse subspecies. Genetics. 2014;198:283–297. [PMC free article] [PubMed] [Google Scholar]
  • Phillips BC, Edmands S. Does the speciation clock tick more slowly in the absence of heteromorphic sex chromosomes? BioEssays. 2012;34:166–169. [PubMed] [Google Scholar]
  • Pickrell JK, Pritchard JK. Inference of population splits and mixtures from genome-wide allele frequency data. PLoS Genetics. 2012;8:e1002967. [PMC free article] [PubMed] [Google Scholar]
  • Pool JE, Nielsen R. Inference of historical changes in migration rate from the lengths of migrant tracts. Genetics. 2009;181:711–719. [PMC free article] [PubMed] [Google Scholar]
  • Porter AH, Wenger R, Geiger H, Scholl A, Shapiro AM. The Pontia daplidice-edusa hybrid zone in northwestern Italy. Evolution. 1997;51:1561–1573. [PubMed] [Google Scholar]
  • Price AL, Patterson NJ, Plenge RM, et al. Principal components analysis corrects for stratification in genome-wide association studies. Nature Genetics. 2006;38:904–909. [PubMed] [Google Scholar]
  • Price AL, Tandon A, Patterson N, et al. Sensitive detection of chromosomal segments of distinct ancestry in admixed populations. PLoS Genetics. 2009;5:e1000519. [PMC free article] [PubMed] [Google Scholar]
  • Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000;155:945–959. [PMC free article] [PubMed] [Google Scholar]
  • Prüfer K, Racimo F, Patterson N, et al. The complete genome sequence of a Neanderthal from the Altai Mountains. Nature. 2014;505:43–49. [PMC free article] [PubMed] [Google Scholar]
  • Raj A, Stephens M, Pritchard JK. fastSTRUCTURE: variational inference of population structure in large SNP data sets. Genetics. 2014;197:573–589. [PMC free article] [PubMed] [Google Scholar]
  • Renaut S, Grassa CJ, Yeaman S, et al. Genomic islands of divergence are not affected by geography of speciation in sunflowers. Nature Communications. 2013;4:1827. [PubMed] [Google Scholar]
  • Rheindt FE, Fujita MK, Wilton PR, Edwards SV. Introgression and phenotypic assimilation in Zimmerius Flycatchers (Tyrannidae): population genetic and phylogenetic inferences from genome-wide SNPs. Systematic Biology. 2014;63:134–152. [PubMed] [Google Scholar]
  • Rieseberg LH. Hybrid origins of plant species. Annual Review of Ecology and Systematics. 1997;28:359–389. [Google Scholar]
  • Rieseberg LH. Chromosomal rearrangements and speciation. Trends in Ecology and Evolution. 2001;16:351–358. [PubMed] [Google Scholar]
  • Rieseberg LH, Whitton J, Gardner K. Hybrid zones and the genetic architecture of a barrier to gene flow between two sunflower species. Genetics. 1999;152:713–727. [PMC free article] [PubMed] [Google Scholar]
  • Robinson JD, Bunnefeld L, Hearn J, Stone GN, Hickerson MJ. ABC inference of multi-population divergence with admixture from unphased population genomic data. Molecular Ecology. 2014;23:4458–4471. [PMC free article] [PubMed] [Google Scholar]
  • Sankararaman S, Patterson N, Li H, Pääabo S, Reich D. The date of interbreeding between Neandertals and modern humans. PLoS Genetics. 2012;8:e1002947. [PMC free article] [PubMed] [Google Scholar]
  • Sankararaman S, Mallick S, Dannemann M, et al. The genomic landscape of Neanderthal ancestry in present-day humans. Nature. 2014;507:354–357. [PMC free article] [PubMed] [Google Scholar]
  • Schlötterer C, Tobler R, Kofler R, Nolte V. Sequencing pools of individuals — mining genome-wide polymorphism data without big funding. Nature Reviews Genetics. 2014;15:749–763. [PubMed] [Google Scholar]
  • Schumer M, Rosenthal GG, Andolfatto P. How common is homoploid hybrid speciation? Evolution. 2014a;68:1553–1560. [PubMed] [Google Scholar]
  • Schumer M, Cui R, Powell DL, Dresner R, Rosenthal GG, Andolfatto P. High-resolution mapping reveals hundreds of genetic incompatibilities in hybridizing fish species. eLife. 2014b;3:e02535. [PMC free article] [PubMed] [Google Scholar]
  • Seehausen O, Butlin RK, Keller I, et al. Genomics and the origin of species. Nature Reviews Genetics. 2014;15:176–192. [PubMed] [Google Scholar]
  • Sethuraman A, Hey J. IMa2p - parallel MCMC and inference of ancient demography under the Isolation with migration (IM) model. Molecular Ecology Resources. 2015;16:206–215. [PMC free article] [PubMed] [Google Scholar]
  • Shaw KL, Mullen SP. Genes versus phenotypes in the study of speciation. Genetica. 2011;139:649–661. [PubMed] [Google Scholar]
  • Smukowski CS, Noor MAF. Recombination rate variation in closely related species. Heredity. 2011;107:496–508. [PMC free article] [PubMed] [Google Scholar]
  • Sousa V, Hey J. Understanding the origin of species with genome-scale data: modelling gene flow. Nature Reviews Genetics. 2013;14:404–414. [PMC free article] [PubMed] [Google Scholar]
  • Sousa VC, Grelaud A, Hey J. On the nonidentifiability of migration time estimates in isolation with migration models. Molecular Ecology. 2011;20:3956–3962. [PMC free article] [PubMed] [Google Scholar]
  • Sousa VC, Carneiro M, Ferrand N, Hey J. Identifying loci under selection against gene flow in isolation-with-migration models. Genetics. 2013;194:211–233. [PMC free article] [PubMed] [Google Scholar]
  • Stebbins GL. The role of hybridization in evolution. Proceedings of the American Philosophical Society. 1959;103:231–251. [Google Scholar]
  • Steinrücken M, Kamm JA, Song YS. Inference of complex population histories using whole genome sequences from multiple populations. Bioarxiv. 2016 doi: 10.1101/026591. [CrossRef] [Google Scholar]
  • Stölting KN, Nipper R, Lindtke D, et al. Genomic scan for single nucleotide polymorphisms reveals patterns of divergence and gene flow between ecologically divergent species. Molecular Ecology. 2013;22:842–855. [PubMed] [Google Scholar]
  • Strasburg JL, Rieseberg LH. Interpreting the estimated timing of migration events between hybridizing species. Molecular Ecology. 2011;20:2353–2366. [PubMed] [Google Scholar]
  • Streicher JW, Devitt TJ, Goldberg CS, et al. Diversification and asymmetrical gene flow across time and space: lineage sorting and hybridization in polytypic barking frogs. Molecular Ecology. 2014;23:3273–3291. [PubMed] [Google Scholar]
  • Szymura JM, Barton NH. Genetic analysis of a hybrid zone between the fire-bellied toads Bombina bombina and B. variegata, near Cracow in Southern Poland. Evolution. 1986;40:1141–1159. [PubMed] [Google Scholar]
  • Tang H, Peng J, Wang P, Risch NJ. Estimation of individual admixture: analytical and study design considerations. Genetic Epidemiology. 2005;28:289–301. [PubMed] [Google Scholar]
  • Via S. Divergence hitchhiking and the spread of genomic isolation during ecological speciation-with-gene-flow. Philosophical Transactions of the Royal Society of London Series B, Biological Sciences. 2012;367:451–460. [PMC free article] [PubMed] [Google Scholar]
  • Wakeley J, Hey J. Estimating ancestral population parameters. Genetics. 1997;145:847–855. [PMC free article] [PubMed] [Google Scholar]
  • Wall JD, Yang MA, Jay F, et al. Higher levels of Neanderthal ancestry in East Asians than in Europeans. Genetics. 2013;194:199–209. [PMC free article] [PubMed] [Google Scholar]
  • Wang Y, Hey J. Estimating divergence parameters with small samples from a large number of loci. Genetics. 2010;184:363–379. [PMC free article] [PubMed] [Google Scholar]
  • Wang L, Luzynski K, Pool JE, et al. Measures of linkage disequilibrium among neighbouring SNPs indicate asymmetries across the house mouse hybrid zone. Molecular Ecology. 2011;20:2985–3000. [PubMed] [Google Scholar]
  • Wegmann D, Kessner DE, Veeramah KR, et al. Recombination rates in admixed individuals identified by ancestry-based inference. Nature Genetics. 2011;43:847–853. [PMC free article] [PubMed] [Google Scholar]
  • Wright S. The genetical structure of populations. Annals of Eugenics. 1951;15:323–354. [PubMed] [Google Scholar]
  • Wright S. Evolution and the Genetics of Populations, vol 2, The Theory of Gene Frequencies. University of Chicago Press; Chicago, Illinois: 1969. [Google Scholar]
  • Wu CI. The genic view of the process of speciation. Journal of Evolutionary Biology. 2001;14:851–865. [Google Scholar]
  • Yang Z. A likelihood ratio test of speciation with gene flow using genomic sequence data. Genome Biology and Evolution. 2010;2:200–211. [PMC free article] [PubMed] [Google Scholar]
  • Yang WY, Platt A, Chiang CWK, et al. Spatial localization of recent ancestors for admixed individuals. G3. 2014;4:2505–2518. [PMC free article] [PubMed] [Google Scholar]
  • Yeaman S. Genomic rearrangements and the evolution of clusters of locally adaptive loci. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:E1743–E1751. [PMC free article] [PubMed] [Google Scholar]
  • Yu Y, Degnan JH, Nakhleh L. The probability of a gene tree topology within a phylogenetic network with applications to hybridization detection. PLoS Genetics. 2012;8:e1002660. [PMC free article] [PubMed] [Google Scholar]
  • Yu Y, Barnett RM, Nakhleh L. Parsimonious inference of hybridization in the presence of incomplete lineage sorting. Systematic Biology. 2013;62:738–751. [PMC free article] [PubMed] [Google Scholar]
  • Yu Y, Dong J, Liu KJ, Nakhleh L. Maximum likelihood inference of reticulate evolutionary histories. Proceedings of the National Academy of Sciences of the United States of America. 2014;111:16448–16453. [PMC free article] [PubMed] [Google Scholar]