Abstract

Evolution has devised countless remarkable solutions to diverse challenges. Understanding the mechanistic basis of these solutions provides insights into how biological systems can be subtly tweaked without maladaptive consequences. The knowledge gained from illuminating these mechanisms is equally important to our understanding of fundamental evolutionary mechanisms as it is to our hopes of developing truly rational plant breeding and synthetic biology. In particular, modern population genomic approaches are proving very powerful in the detection of candidate alleles for mediating consequential adaptations that can be tested functionally. Especially striking are signals gained from contexts involving genetic transfers between populations, closely related species, or indeed between kingdoms. Here we discuss two major classes of these scenarios, adaptive introgression and horizontal gene flow, illustrating discoveries made across kingdoms.

Introduction

Whether it is gradual or sudden, all organisms face change. Adaptive responses are therefore required for survival, especially in species that cannot migrate. The footprints of these responses are found throughout the genome, serving as powerful signals that tell us how populations have overcome hazards, both biotic and abiotic. The allelic changes at loci mediating adaptive changes are coming to light in rapidly increasing numbers of studies, and thanks to ongoing developments in population genomics, descriptions of these loci appear in remarkably high resolution. As a result, there is now very good evidence that diverse sources of genetic variation underlie important phenotypic changes in wild populations. Among these, introgression is emerging as a widespread fundamental evolutionary force. The term ‘introgressive hybridization’, hereafter referred to as ‘introgression’, was introduced by Anderson and Hubricht (1938). They referred to the introduction of syntenic nucleotide variation by recombination from a donor species into the genome of a recipient species, usually by means of hybridization and backcrossing. We will use the terms 'introgression' and 'gene flow' as synonyms, also in cases when the units that exchange variants are populations of the same species. Largely context-dependent, introgression is influenced by an array of ecological factors that control the degree of contact between species. These biotic or abiotic factors drive selection for or against hybrid genotypes and can lead to complex patterns of genetic admixture (Hand et al., 2015). This selection has important consequences: adaptive introgression commonly results in local adaptation to particular geographically distributed conditions and/or speciation (Dobzhansky, 1937; Mayr 1942; Coyne and Orr, 2004).

Here we discuss the three dominant sources of genetic variation and their relative contributions to adaptation. We show how genomic approaches are revolutionizing the discovery of adaptive alleles involved in natural solutions to diverse challenges. We highlight how the studies of introgression and speciation are linked, with speciation commonly occurring in the face of gene flow. This can create hallmark genomic architectures that facilitate the discovery of adaptive alleles, which can lead to ‘genomic islands of divergence’ that are resistant to gene flow. We argue that the novel merger of hybrid zone analysis and whole genome population resequencing is a powerful novel tool to detect variants mediating diverse adaptations. We provide an overview of important approaches taken to identify introgressed alleles that can be applied by researchers working on virtually any system and that have the potential to unambiguously identify strong candidates for adaptively introgressed alleles. Finally, we highlight horizontal gene transfer (HGT), a special case of adaptive introgression between species with established reproductive boundaries.

Maladaptive introgression

Before discussing introgression generally and adaptive introgression in particular, it is important to note that not all introgression is adaptive; indeed, introgression is a powerful force and can be strongly disruptive, particularly as the result of human activity. Introgression even has the potential to hinder conservation attempts (Allendorf et al., 2001; Edmands, 2007). Due to a shortage of conspecific mates (Vaz Pinto et al., 2016), introgression may drive rare species to extinction by genetic swamping (Wolf et al., 2001; Gómez et al., 2015; Todesco et al., 2016). Conservation attempts based on the translocation of species to reserves outside native ranges can result in introgression and inadvertent admixture that damages the biodiversity of the protected species (e.g. antelope; van Wyk et al., 2017). In addition, introgression between crops or domesticated animals and wild relatives can alter fitness-related traits, such as disease resistance and growth, although every case is unique. Recent examples include introgression from domesticated dogs into wolves (Anderson et al., 2009), between farmed and native salmonids (Glover et al., 2013; Ozerov et al., 2016; Karlsson et al., 2016), from wild pigs into domesticated pigs (Ai et al., 2015), from wildcats into domesticated cats (Ottoni et al., 2017), between maize and teosinte (Hufford et al., 2012; Hufford et al., 2013), from domesticated rice into wild rice (Wang et al., 2017), and between genetically modified plants and their wild relatives (den Nijs et al., 2004). Introgression between native and introduced, often invasive, species has also been reported, with examples including mussels (Saarman and Pogson, 2015), salamander (Fitzpatrick et al., 2010; Wilcox et al., 2015), trout (Hohenlohe et al., 2013; Muhlfeld et al., 2014; Kovach et al., 2015; Kovach et al., 2016) and Ulmus trees (Zalapa et al., 2009). Genomes may be resistant to the introgression of invading alleles when selection favours the native allele, as shown in, for example, trout (Kovach et al., 2015; Kovach et al., 2016) and Arabidopsis thaliana (Lee et al., 2017). Interestingly, adaptive introgression can also occur differentially from one subgenome of an allopolyploid, for example from wheat into Aegilops (Parisod et al., 2013).

Introgression as an engine of adaptive genetic variation

There are three primary sources of genetic variation: (1) pre-existing or ‘standing’ variants, which are the variants already present in a population, (2) new mutations, and (3) introgression (reviewed in Olson-Manning et al., 2012; Hedrick, 2013). Despite introgression traditionally being seen as maladaptive, there is growing literature demonstrating the widespread occurrence of adaptive introgression (reviewed in Mallet et al., 2016). Introgression is implicated to be a powerful adaptive force in a wide array of taxa, including, for example, the malaria transmitting mosquito Anopheles (Clarkson et al., 2014; Fontaine et al., 2015), Heliconius butterflies (Zhang et al., 2016), mice (Song et al., 2011), humans (Racimo et al., 2017), Arabidopsis (Arnold et al., 2016), sunflowers (Whitney et al., 2015), and monkeyflowers (Stankowski and Streisfeld, 2015). However, the relative importance of introgression, de novo mutation, and standing variation is far from resolved. Barton (2001) concluded that adaptive variation engendered by mutation is likely to exceed that brought about by introgression. This is all the more likely if the effective population sizes are large because the probability of a favourable mutation is a function of effective population size. For example, many pests and weeds have tremendous effective population sizes and there is strong evidence that cases of escape from chemical insecticides and herbicides have originated many times from independent novel mutations, for example in Drosophila (Karasov et al., 2010) and in weeds (Délye et al., 2013). However this is not always the case, for example Anopheles traits enhancing vectorial capacity, including the knockdown resistance mediated by a specific single nucleotide polymorphism, were transferred between two hybridizing species (Weill et al., 2000).

Anderson (1949) suggested that introgressed variation should have a higher initial frequency than new adaptive mutations and a lower initial frequency than standing variation. But if introgression is recurrent and results in fit progeny, early frequencies could be much higher, exceeding the adaptive potential of standing variation. Further, the impact of single introgressed variants on the genome can be sizable, typically causing multiple changes within a gene and at times affecting several gene-coding loci. Striking examples involve the transfer of entire complex adaptations via cassettes of multiple linked mutations, such as those in loci that control wing colour patterns for both mimicry and mate recognition in Heliconius butterflies (The Heliconius Genome Consortium, 2012). Such an advantage can also be found in some cases of adaptive standing variation (Bastide et al., 2016) and occasionally for de novo adaptive mutations (Karasov et al., 2010). In addition, introgressed alleles feature the benefit that they are likely to have been pre-tested by selection in a usually closely related donor species and would therefore be less likely to be deleterious than random mutations (Hedrick, 2013). Indeed, introgressed variants that initially have no strong advantage or disadvantage can accumulate in the genome as cryptic variation, which then serves as the raw material for selection when conditions change (reviewed in Paaby and Rockman, 2014).

The ability to introgress is dependent on the degree of divergence, as introgression between more divergent species is usually impeded by pre- and post-zygotic reproductive barriers. It has recently been suggested that polyploidy can occasionally rescue introgression between otherwise reproductively isolated species. For example, polyploidization re-established normal endosperm cellularization and enabled unidirectional interploidal introgression and bidirectional introgression between tetraploids of Arabidopsis arenosa and Arabidopsis lyrata (Lafon-Placette et al., 2017). A case of introgression between the autotetraploid cytotypes of these species reports the exchange of candidate alleles for mediating adaptation to highly challenging serpentine soils (Arnold et al., 2016). Interploidal introgression is generally assumed to be unidirectional, from diploid to polyploid (Stebbins, 1971), although there is evidence that it can also occur in the reverse direction (Ramsey and Schemske, 1998). There have been very few genomic studies on interploidal introgression and nearly all reports detail gene flow from diploids to tetraploids, for example in Betulus (Zohren et al., 2016). Another study in Miscanthus favoured the same unidirectionality of gene flow, although there was some evidence for occasional gene flow from polyploids to diploids (Clark et al., 2015). Due to the scarcity of genomic studies one is left with literature based on only handfuls of molecular markers, the vast majority of which support gene flow from diploids to polyploids, such as in Senecio (Kim et al., 2008; Chapman and Abbott, 2010), Epidendrum (Pinheiro et al., 2010), and Capsella (Han et al., 2015). Except for Kim et al. (2008) and Chapman and Abbott (2010) who found evidence of introgression of fitness-related genes, it is unclear whether interploidal introgression frequently leads to the transfer of adaptive alleles.

Hybridization, particularly in cases where hybridization leads to allopolyploidy, and introgression can also introduce variation at the structural level leading to genome rearrangements and novel gene regulatory pathways through the reactivation of dormant transposable elements (TEs) (reviewed in Fontdevila, 2005). For instance, introgression between cultivated and wild rice led to changes in transcription levels and DNA methylation patterns in TE-rich genomic regions (Liu et al., 2004). Introgression between bread wheat and tall wheatgrass led to various genetic and epigenetic changes, including deletions, differences in gene expression and TE reactivation (Liu et al., 2015). Similar changes have been identified in introgression lines between cauliflower (Brassica oleracea) and black mustard (Brassica nigra), creating variation that might prove useful for trait selection in breeding (Wang et al., 2016).

Introgression aids adaptive allele discovery in the genomic era

The widespread occurrence of introgression suggests promise in mining for adaptive alleles in natural populations that have overcome identifiable biotic or abiotic challenges. In parallel, there is a growing interest in how knowledge obtained from more broadly observing natural solutions devised by evolution might inform breeding efforts. Mining for introgressed alleles between natural populations that are native to contrasting environments offers the possibility to identify candidate alleles that mediate definable adaptations. Indeed, it has been long recognized that especially where fitness-related traits and genotypes show clinal variation, gene-environment associations can provide a window on the mechanisms of natural selection (Endler, 1977). Such information is virtually impossible to discover with experimental crosses, although evolve and resequence experiments can be informative (reviewed in Long et al., 2015). However, mining for adaptive alleles in natural populations with population genomics, namely ‘reverse ecology’ (Li et al., 2008), complements and in many ways surpasses candidate gene-based approaches. By providing a genome-wide view of the divergence landscape, population genomic studies can overcome key limitations of candidate-based approaches; for instance, they can detect polygenic adaptation provided markers are dense, that is to say genome resequencing as opposed to restriction site associated DNA sequencing. Such mining for co-evolved alleles involved in a trait is facilitated if frequencies of these alleles follow a clinal pattern in a hybrid zone that may coincide with an environmental gradient.

Genome-wide data from populations that are characterized by a history of extensive gene flow can provide detailed insights into the genetic basis of adaptive divergence. Usually in such cases only a few loci under selection rise above the neutral background that is homogenized by gene flow. Here we focus on these population genomic studies, because they can leverage sufficient resolution and statistical power to detect non-random patterns of introgression. We note, however, that crucial insights into the evolutionary role of adaptive introgression were made before the advent of high-throughput sequencing and population genomic datasets (reviewed in Arnold and Martin, 2009). Indeed, population genomics and comparative phylogenomics are young fields and sometimes issues with data quality or analytical rigor can raise concerns that render results ambiguous (Brower, 2013; Wen et al., 2016). Nevertheless, with well-designed sampling and increasingly sophisticated analysis, evidence of genomic heterogeneity in patterns of gene flow has been accumulating at all levels from weakly (Arnold et al., 2016) to highly differentiated (Pardo-Diaz et al., 2012) species. Understanding these population- and genome-wide patterns of introgression in species with diverse adaptations has broad potential to contribute to our understanding of the fundamental mechanisms of adaptation.

The genomic architecture of introgression

The study of introgression in natural populations is deeply connected with the study of speciation. It is increasingly recognized that speciation frequently occurs in the face of gene flow; either through continual gene flow or phases of secondary contact in species that have already established partial reproductive isolation, hereafter referred to as parapatrically isolated species (PaIS) (Nosil, 2008; Feder et al., 2012; Abbott et al., 2013; Harrison and Larson, 2014; Mallet et al., 2016; Shapiro et al., 2016; Arnold, 2016). One iconic example is the sunflower Helianthus annuus subsp. texanus, which originated as a result of introgression and shows different adaptations compared with its parents, mainly abiotic tolerance traits (Whitney et al., 2010; Whitney et al., 2015). It may be difficult to discriminate ongoing divergence from secondary contact in PaIS (Endler, 1977; Harrison, 2012; Gompert and Buerkle, 2016), especially in cases where species have recently diverged. Both scenarios of introgression offer potential insights into the mechanisms of adaptive introgression, although major findings have been made in cases of secondary contact (Table 1), as speciation-with-gene-flow studies usually aim to identify loci involved in the establishment and maintenance of reproductive isolation and not loci involved in local adaptation. Strong evidence for secondary contact in major systems that allow the study of adaptive introgression comes from clear estimates of the time of secondary contact as well as the delineation of clearly defined hybrid zones. Both the time of species divergence and secondary contact are relatively young in these systems (Table 1). Except for the hybridization between Heliconius erato and H. melpomene (Martin et al., 2013; Kozak et al., 2015) and Saccharomyces cerevisiae and S. uvarum (Kellis et al., 2003), all other study systems show divergence times for the introgressing species of usually much less than 2 mya and times of secondary contact between a few hundred and few thousand years before the present time.

Table 1.

Important systems for the study of adaptive introgression in natural populations. Studies leveraging genome-wide high-throughput sequencing data were selected and employed methods of demographic inference and genome architecture of introgression are listed, which are described in Table 2. Especially in intensely studied systems, i.e., Drosophila, Heliconius butterflies, humans and yeast, we restricted the number of case studies to the most recent and/or in our opinion highly interesting ones

Hybrid partnersGenomic data typesMethods appliedAdaptive traits or genes influencedStart of divergence of the hybridizing taxaEstimated time of secondary contactPresence of hybrid zone(s)
Arabidopsis
Between Arabidopsis arenosa and A. lyrata (Schmickl and Koch, 2011; Arnold et al., 2016)Whole genome sequencing (Arnold et al., 2016)Fastsimcoal2, fD statistic, HybridCheck; genome scanGenes controlling specific ion homeostasis-related and drought adaptation traits (Arnold et al., 2016)~1.63 mya
(Hohmann et al., 2015)
Eastern Austrian Forealps (Schmickl and Koch, 2011)
Sunflowers
From Helianthus debilis subsp. cucumerifolius into H. annuus subsp. annuus (Whitney et al., 2015)SNP dataset (QTL mapping)Primarily abiotic tolerance traits allowing rapid growth and reproduction before summer heat and drought, but also traits affecting herbivore resistanceDivergence between the ancestors of H. debilis and H. annuus ~2 mya (Renaut et al., 2013)Maximum ~18,000 ya (Scascitelli et al., 2010)South-central USA
Senecio
From diploid Senecio squalidus into tetraploid S. vulgaris (Kim et al., 2008)[No large genome- wide dataset] Sanger sequencing, RNA in situ hybridizationGenes RAY1 and RAY2 that control ray florets in radiate flower heads, which effects attractiveness to pollinators and outcrossing rateS. squalidus is a homoploid hybrid with an origin less than ~150,000 ya (Filatov et al., 2016)Maximum ~300 ya (S. squalidus was introduced to the Oxford Botanic Garden around 300 ya)United Kingdom
Poplars
From Populus balsamifera into P. trichocarpa (Suarez- Gonzalez et al., 2016)Whole genome sequencing, RNAseqGenome scanNumerous genes, particularly in a telomeric region (PRR5 and COMT1); related to response to far red light, RNA processing and ATPase activity~76,000 ya (Levsen et al., 2012)Northwestern North America
Mussels
From Mytilus edulis into M. galloprovincialis (Fraïsse et al., 2014; Fraïsse et al., 2016)AFLP and few other markers, Sanger sequencing (Fraïsse et al., 2014); target enrichment (Fraïsse et al., 2016)Genome scanActin gene mac-1 (Fraïsse et al., 2014; but not Fraïsse et al., 2016)~2.5 mya (Roux et al., 2014)~0.7 mya (Roux et al., 2014)In Europe from the Mediterranean Sea to the North Sea
Drosophila
From Drosophila simulans into D. sechellia (Garrigan et al., 2012; Brand et al., 2013)Garrigan et al., 2012: whole genome sequencing; Brand et al., 2013: Sanger sequencing of the candidate region identified by Garrigan et al., 2012 Genome scan (Garrigan et al., 2012)A 15 kb region (Garrigan et al., 2012) – the target of selection appears to be a regulatory sequence; evidence for a trans-specific selective sweep (Brand et al., 2013), i.e., the 15 kb region harbours a polymorphism that is adaptive in both species~200,000–300,000 ya (Garrigan et al., 2012)2,600–18,600 yaSeychelles
From D. yakuba into D. santomea (Llopart et al., 2014; Beck et al., 2015)Llopart et al., 2014: Sanger sequencing and high-throughput sequencing of the mitochondrial genome; Beck et al., 2015: Sanger sequencingPotential cytonuclear co-introgression of genes involved in the oxidative phosphorylation pathway; strong signal of introgression in the three nuclear genes composing subunit V of the cytochrome c oxidase complex (Beck et al., 2015)~400,000 ya (Llopart et al., 2002)~14,000 ya (complete replacement of the mitochondrial genome) and ~2,500 years ago (Llopart et al., 2014)São Tomé (invasion by D. yakuba)
Anopheles
Between Anopheles arabiensis and the ancestor of A. gambiae and A. coluzzii
(Fontaine et al., 2015)
Whole genome sequencingD and DFOIL statistics, window-based gene tree distribution across chromosomesInsecticide-resistance alleles might have introgressed into A. arabiensis with the 2La inversion introgressed from the ancestor of A. gambiae and A. coluzziiDivergence between A. gambiae and A. coluzzii from its ancestor ~2 mya~0.5 myaCentral Africa
From A. gambiae into A. coluzzii (Norris et al., 2015)SNP genotyping, whole genome sequencingTraits enhancing vectorial capacity: A. coluzzii inherited the entire A. gambiae- associated 2L divergence island including the knockdown resistance SNP L1014F in the kdr gene (L1014F has been present in A. gambiae populations for at least 20 years)Periods of isolation interrupted by hybridization; recent introgression reported in 2002 and 2006 (Lee et al., 2013b )Selinkenyi, Mali
Heliconius butterflies
Other studies that focused on speciation with ongoing gene flow and adaptive introgression: e.g. Pardo-Diaz et al., 2012; The Heliconius Genome Consortium, 2012; Nadeau et al., 2013; Kronforst et al., 2013; Martin et al., 2013)
Between subspecies of Heliconius erato and H. melpomene (Nadeau et al., 2014)RAD-seqStructure; genome scanLoci that control wing color patterns (for both mimicry and mate recognition); the aim of this study was to identify novel loci underlying phenotypic variation by means of association mapping in hybrid zones~11.8 mya (Kozak et al., 2015)Several time points post speciation (Martin et al., 2013)Parallel zones in the Andes
Between H. beschkei, H. numata and H. numata subsp. nanna (Zhang et al., 2016)Whole genome sequencingD and fD statisticsGenomic region upstream of the gene optix (known to control red wing patterning) and 39 additional genomic regionsDivergence between H. beschkei and H. numata ~1.93 myaEast-central South America
Darwin’s finches
Species from different Galápagos islands
(Lamichhaney et al., 2015)
Whole genome sequencingPSMC, NeighborNet, D statistic; genome scanALX1 gene encodes a transcription factor controlling craniofacial development, in particular beak shape~0.5 myaSeveral time points post speciationGalápagos islands
Mice
From Mus spretus into M. musculus subsp. domesticus (Song et al., 2011)[No large genome- wide dataset] Sanger sequencing>10 Mb region including the molecular target of anticoagulants vkorc1; M. spretus-specific alleles can cause resistance to rodenticides~1.5–3 myaAfrica and Europe
Between M. musculus subsp. domesticus and M. musculus subsp. musculus
(Staubach et al., 2012)
SNP genotypingGenome scanNo general pattern of types of genes subject to introgression300,000–500,000 ya (Guénet and Bonhomme, 2003)Maximum ~3,000 ya (Cucchi et al., 2005)Eastern and southern Germany, western Czech Republic
Humans
Other studies identified genomic regions with evidence of adaptive introgression between archaic and modern humans (from Neanderthals into modern humans: e.g., Vernot and Akey, 2014; Sankararaman et al., 2014 / from Denisovans into Tibetans: e.g., Jeong et al., 2014; Huerta-Sánchez et al., 2014)
From Neanderthals and Denisovans into modern humans (Racimo et al., 2017)Whole genome sequencing; reevaluation of published dataD and fD statistics, RD, U and Q95 statistics, HaplostripsLipid metabolism, pigmentation and innate immunityDivergence between the Neanderthal/ Denisovan lineage and modern humans 170,000–700,000 ya (Meyer et al., 2012)Between Neanderthals and modern humans 47,000– 65,000 ya (Sankararaman et al., 2012)Eurasia
Yeast
Between Saccharomyces cerevisiae and S. uvarum (Dunn et al., 2013)Whole genome sequencing, colony-PCR and qRT-PCRMEP2 fusion gene encodes for a high-affinity ammonium permease, which may be adaptive under nitrogen-poor environments with ammonium as only nitrogen source~20 mya (Kellis et al., 2003)de novo creation of hybrid followed by continuous ammonium limitationExperimental populations
Albugo candida
Between the ancestors of three races; an adaptive radiation in progress (McMullan et al., 2015)Whole genome sequencing, RNAseqHybridCheckEffector alleles that may result in a fitness advantage on most potential hosts, enabling host jumpsMultiple time points; oldest event ~200,000 ya, most recent event ~2,200 yaInfection with virulent A. candida suppresses host immunity enabling co-colonization by otherwise non-virulent races
Hybrid partnersGenomic data typesMethods appliedAdaptive traits or genes influencedStart of divergence of the hybridizing taxaEstimated time of secondary contactPresence of hybrid zone(s)
Arabidopsis
Between Arabidopsis arenosa and A. lyrata (Schmickl and Koch, 2011; Arnold et al., 2016)Whole genome sequencing (Arnold et al., 2016)Fastsimcoal2, fD statistic, HybridCheck; genome scanGenes controlling specific ion homeostasis-related and drought adaptation traits (Arnold et al., 2016)~1.63 mya
(Hohmann et al., 2015)
Eastern Austrian Forealps (Schmickl and Koch, 2011)
Sunflowers
From Helianthus debilis subsp. cucumerifolius into H. annuus subsp. annuus (Whitney et al., 2015)SNP dataset (QTL mapping)Primarily abiotic tolerance traits allowing rapid growth and reproduction before summer heat and drought, but also traits affecting herbivore resistanceDivergence between the ancestors of H. debilis and H. annuus ~2 mya (Renaut et al., 2013)Maximum ~18,000 ya (Scascitelli et al., 2010)South-central USA
Senecio
From diploid Senecio squalidus into tetraploid S. vulgaris (Kim et al., 2008)[No large genome- wide dataset] Sanger sequencing, RNA in situ hybridizationGenes RAY1 and RAY2 that control ray florets in radiate flower heads, which effects attractiveness to pollinators and outcrossing rateS. squalidus is a homoploid hybrid with an origin less than ~150,000 ya (Filatov et al., 2016)Maximum ~300 ya (S. squalidus was introduced to the Oxford Botanic Garden around 300 ya)United Kingdom
Poplars
From Populus balsamifera into P. trichocarpa (Suarez- Gonzalez et al., 2016)Whole genome sequencing, RNAseqGenome scanNumerous genes, particularly in a telomeric region (PRR5 and COMT1); related to response to far red light, RNA processing and ATPase activity~76,000 ya (Levsen et al., 2012)Northwestern North America
Mussels
From Mytilus edulis into M. galloprovincialis (Fraïsse et al., 2014; Fraïsse et al., 2016)AFLP and few other markers, Sanger sequencing (Fraïsse et al., 2014); target enrichment (Fraïsse et al., 2016)Genome scanActin gene mac-1 (Fraïsse et al., 2014; but not Fraïsse et al., 2016)~2.5 mya (Roux et al., 2014)~0.7 mya (Roux et al., 2014)In Europe from the Mediterranean Sea to the North Sea
Drosophila
From Drosophila simulans into D. sechellia (Garrigan et al., 2012; Brand et al., 2013)Garrigan et al., 2012: whole genome sequencing; Brand et al., 2013: Sanger sequencing of the candidate region identified by Garrigan et al., 2012 Genome scan (Garrigan et al., 2012)A 15 kb region (Garrigan et al., 2012) – the target of selection appears to be a regulatory sequence; evidence for a trans-specific selective sweep (Brand et al., 2013), i.e., the 15 kb region harbours a polymorphism that is adaptive in both species~200,000–300,000 ya (Garrigan et al., 2012)2,600–18,600 yaSeychelles
From D. yakuba into D. santomea (Llopart et al., 2014; Beck et al., 2015)Llopart et al., 2014: Sanger sequencing and high-throughput sequencing of the mitochondrial genome; Beck et al., 2015: Sanger sequencingPotential cytonuclear co-introgression of genes involved in the oxidative phosphorylation pathway; strong signal of introgression in the three nuclear genes composing subunit V of the cytochrome c oxidase complex (Beck et al., 2015)~400,000 ya (Llopart et al., 2002)~14,000 ya (complete replacement of the mitochondrial genome) and ~2,500 years ago (Llopart et al., 2014)São Tomé (invasion by D. yakuba)
Anopheles
Between Anopheles arabiensis and the ancestor of A. gambiae and A. coluzzii
(Fontaine et al., 2015)
Whole genome sequencingD and DFOIL statistics, window-based gene tree distribution across chromosomesInsecticide-resistance alleles might have introgressed into A. arabiensis with the 2La inversion introgressed from the ancestor of A. gambiae and A. coluzziiDivergence between A. gambiae and A. coluzzii from its ancestor ~2 mya~0.5 myaCentral Africa
From A. gambiae into A. coluzzii (Norris et al., 2015)SNP genotyping, whole genome sequencingTraits enhancing vectorial capacity: A. coluzzii inherited the entire A. gambiae- associated 2L divergence island including the knockdown resistance SNP L1014F in the kdr gene (L1014F has been present in A. gambiae populations for at least 20 years)Periods of isolation interrupted by hybridization; recent introgression reported in 2002 and 2006 (Lee et al., 2013b )Selinkenyi, Mali
Heliconius butterflies
Other studies that focused on speciation with ongoing gene flow and adaptive introgression: e.g. Pardo-Diaz et al., 2012; The Heliconius Genome Consortium, 2012; Nadeau et al., 2013; Kronforst et al., 2013; Martin et al., 2013)
Between subspecies of Heliconius erato and H. melpomene (Nadeau et al., 2014)RAD-seqStructure; genome scanLoci that control wing color patterns (for both mimicry and mate recognition); the aim of this study was to identify novel loci underlying phenotypic variation by means of association mapping in hybrid zones~11.8 mya (Kozak et al., 2015)Several time points post speciation (Martin et al., 2013)Parallel zones in the Andes
Between H. beschkei, H. numata and H. numata subsp. nanna (Zhang et al., 2016)Whole genome sequencingD and fD statisticsGenomic region upstream of the gene optix (known to control red wing patterning) and 39 additional genomic regionsDivergence between H. beschkei and H. numata ~1.93 myaEast-central South America
Darwin’s finches
Species from different Galápagos islands
(Lamichhaney et al., 2015)
Whole genome sequencingPSMC, NeighborNet, D statistic; genome scanALX1 gene encodes a transcription factor controlling craniofacial development, in particular beak shape~0.5 myaSeveral time points post speciationGalápagos islands
Mice
From Mus spretus into M. musculus subsp. domesticus (Song et al., 2011)[No large genome- wide dataset] Sanger sequencing>10 Mb region including the molecular target of anticoagulants vkorc1; M. spretus-specific alleles can cause resistance to rodenticides~1.5–3 myaAfrica and Europe
Between M. musculus subsp. domesticus and M. musculus subsp. musculus
(Staubach et al., 2012)
SNP genotypingGenome scanNo general pattern of types of genes subject to introgression300,000–500,000 ya (Guénet and Bonhomme, 2003)Maximum ~3,000 ya (Cucchi et al., 2005)Eastern and southern Germany, western Czech Republic
Humans
Other studies identified genomic regions with evidence of adaptive introgression between archaic and modern humans (from Neanderthals into modern humans: e.g., Vernot and Akey, 2014; Sankararaman et al., 2014 / from Denisovans into Tibetans: e.g., Jeong et al., 2014; Huerta-Sánchez et al., 2014)
From Neanderthals and Denisovans into modern humans (Racimo et al., 2017)Whole genome sequencing; reevaluation of published dataD and fD statistics, RD, U and Q95 statistics, HaplostripsLipid metabolism, pigmentation and innate immunityDivergence between the Neanderthal/ Denisovan lineage and modern humans 170,000–700,000 ya (Meyer et al., 2012)Between Neanderthals and modern humans 47,000– 65,000 ya (Sankararaman et al., 2012)Eurasia
Yeast
Between Saccharomyces cerevisiae and S. uvarum (Dunn et al., 2013)Whole genome sequencing, colony-PCR and qRT-PCRMEP2 fusion gene encodes for a high-affinity ammonium permease, which may be adaptive under nitrogen-poor environments with ammonium as only nitrogen source~20 mya (Kellis et al., 2003)de novo creation of hybrid followed by continuous ammonium limitationExperimental populations
Albugo candida
Between the ancestors of three races; an adaptive radiation in progress (McMullan et al., 2015)Whole genome sequencing, RNAseqHybridCheckEffector alleles that may result in a fitness advantage on most potential hosts, enabling host jumpsMultiple time points; oldest event ~200,000 ya, most recent event ~2,200 yaInfection with virulent A. candida suppresses host immunity enabling co-colonization by otherwise non-virulent races
Table 1.

Important systems for the study of adaptive introgression in natural populations. Studies leveraging genome-wide high-throughput sequencing data were selected and employed methods of demographic inference and genome architecture of introgression are listed, which are described in Table 2. Especially in intensely studied systems, i.e., Drosophila, Heliconius butterflies, humans and yeast, we restricted the number of case studies to the most recent and/or in our opinion highly interesting ones

Hybrid partnersGenomic data typesMethods appliedAdaptive traits or genes influencedStart of divergence of the hybridizing taxaEstimated time of secondary contactPresence of hybrid zone(s)
Arabidopsis
Between Arabidopsis arenosa and A. lyrata (Schmickl and Koch, 2011; Arnold et al., 2016)Whole genome sequencing (Arnold et al., 2016)Fastsimcoal2, fD statistic, HybridCheck; genome scanGenes controlling specific ion homeostasis-related and drought adaptation traits (Arnold et al., 2016)~1.63 mya
(Hohmann et al., 2015)
Eastern Austrian Forealps (Schmickl and Koch, 2011)
Sunflowers
From Helianthus debilis subsp. cucumerifolius into H. annuus subsp. annuus (Whitney et al., 2015)SNP dataset (QTL mapping)Primarily abiotic tolerance traits allowing rapid growth and reproduction before summer heat and drought, but also traits affecting herbivore resistanceDivergence between the ancestors of H. debilis and H. annuus ~2 mya (Renaut et al., 2013)Maximum ~18,000 ya (Scascitelli et al., 2010)South-central USA
Senecio
From diploid Senecio squalidus into tetraploid S. vulgaris (Kim et al., 2008)[No large genome- wide dataset] Sanger sequencing, RNA in situ hybridizationGenes RAY1 and RAY2 that control ray florets in radiate flower heads, which effects attractiveness to pollinators and outcrossing rateS. squalidus is a homoploid hybrid with an origin less than ~150,000 ya (Filatov et al., 2016)Maximum ~300 ya (S. squalidus was introduced to the Oxford Botanic Garden around 300 ya)United Kingdom
Poplars
From Populus balsamifera into P. trichocarpa (Suarez- Gonzalez et al., 2016)Whole genome sequencing, RNAseqGenome scanNumerous genes, particularly in a telomeric region (PRR5 and COMT1); related to response to far red light, RNA processing and ATPase activity~76,000 ya (Levsen et al., 2012)Northwestern North America
Mussels
From Mytilus edulis into M. galloprovincialis (Fraïsse et al., 2014; Fraïsse et al., 2016)AFLP and few other markers, Sanger sequencing (Fraïsse et al., 2014); target enrichment (Fraïsse et al., 2016)Genome scanActin gene mac-1 (Fraïsse et al., 2014; but not Fraïsse et al., 2016)~2.5 mya (Roux et al., 2014)~0.7 mya (Roux et al., 2014)In Europe from the Mediterranean Sea to the North Sea
Drosophila
From Drosophila simulans into D. sechellia (Garrigan et al., 2012; Brand et al., 2013)Garrigan et al., 2012: whole genome sequencing; Brand et al., 2013: Sanger sequencing of the candidate region identified by Garrigan et al., 2012 Genome scan (Garrigan et al., 2012)A 15 kb region (Garrigan et al., 2012) – the target of selection appears to be a regulatory sequence; evidence for a trans-specific selective sweep (Brand et al., 2013), i.e., the 15 kb region harbours a polymorphism that is adaptive in both species~200,000–300,000 ya (Garrigan et al., 2012)2,600–18,600 yaSeychelles
From D. yakuba into D. santomea (Llopart et al., 2014; Beck et al., 2015)Llopart et al., 2014: Sanger sequencing and high-throughput sequencing of the mitochondrial genome; Beck et al., 2015: Sanger sequencingPotential cytonuclear co-introgression of genes involved in the oxidative phosphorylation pathway; strong signal of introgression in the three nuclear genes composing subunit V of the cytochrome c oxidase complex (Beck et al., 2015)~400,000 ya (Llopart et al., 2002)~14,000 ya (complete replacement of the mitochondrial genome) and ~2,500 years ago (Llopart et al., 2014)São Tomé (invasion by D. yakuba)
Anopheles
Between Anopheles arabiensis and the ancestor of A. gambiae and A. coluzzii
(Fontaine et al., 2015)
Whole genome sequencingD and DFOIL statistics, window-based gene tree distribution across chromosomesInsecticide-resistance alleles might have introgressed into A. arabiensis with the 2La inversion introgressed from the ancestor of A. gambiae and A. coluzziiDivergence between A. gambiae and A. coluzzii from its ancestor ~2 mya~0.5 myaCentral Africa
From A. gambiae into A. coluzzii (Norris et al., 2015)SNP genotyping, whole genome sequencingTraits enhancing vectorial capacity: A. coluzzii inherited the entire A. gambiae- associated 2L divergence island including the knockdown resistance SNP L1014F in the kdr gene (L1014F has been present in A. gambiae populations for at least 20 years)Periods of isolation interrupted by hybridization; recent introgression reported in 2002 and 2006 (Lee et al., 2013b )Selinkenyi, Mali
Heliconius butterflies
Other studies that focused on speciation with ongoing gene flow and adaptive introgression: e.g. Pardo-Diaz et al., 2012; The Heliconius Genome Consortium, 2012; Nadeau et al., 2013; Kronforst et al., 2013; Martin et al., 2013)
Between subspecies of Heliconius erato and H. melpomene (Nadeau et al., 2014)RAD-seqStructure; genome scanLoci that control wing color patterns (for both mimicry and mate recognition); the aim of this study was to identify novel loci underlying phenotypic variation by means of association mapping in hybrid zones~11.8 mya (Kozak et al., 2015)Several time points post speciation (Martin et al., 2013)Parallel zones in the Andes
Between H. beschkei, H. numata and H. numata subsp. nanna (Zhang et al., 2016)Whole genome sequencingD and fD statisticsGenomic region upstream of the gene optix (known to control red wing patterning) and 39 additional genomic regionsDivergence between H. beschkei and H. numata ~1.93 myaEast-central South America
Darwin’s finches
Species from different Galápagos islands
(Lamichhaney et al., 2015)
Whole genome sequencingPSMC, NeighborNet, D statistic; genome scanALX1 gene encodes a transcription factor controlling craniofacial development, in particular beak shape~0.5 myaSeveral time points post speciationGalápagos islands
Mice
From Mus spretus into M. musculus subsp. domesticus (Song et al., 2011)[No large genome- wide dataset] Sanger sequencing>10 Mb region including the molecular target of anticoagulants vkorc1; M. spretus-specific alleles can cause resistance to rodenticides~1.5–3 myaAfrica and Europe
Between M. musculus subsp. domesticus and M. musculus subsp. musculus
(Staubach et al., 2012)
SNP genotypingGenome scanNo general pattern of types of genes subject to introgression300,000–500,000 ya (Guénet and Bonhomme, 2003)Maximum ~3,000 ya (Cucchi et al., 2005)Eastern and southern Germany, western Czech Republic
Humans
Other studies identified genomic regions with evidence of adaptive introgression between archaic and modern humans (from Neanderthals into modern humans: e.g., Vernot and Akey, 2014; Sankararaman et al., 2014 / from Denisovans into Tibetans: e.g., Jeong et al., 2014; Huerta-Sánchez et al., 2014)
From Neanderthals and Denisovans into modern humans (Racimo et al., 2017)Whole genome sequencing; reevaluation of published dataD and fD statistics, RD, U and Q95 statistics, HaplostripsLipid metabolism, pigmentation and innate immunityDivergence between the Neanderthal/ Denisovan lineage and modern humans 170,000–700,000 ya (Meyer et al., 2012)Between Neanderthals and modern humans 47,000– 65,000 ya (Sankararaman et al., 2012)Eurasia
Yeast
Between Saccharomyces cerevisiae and S. uvarum (Dunn et al., 2013)Whole genome sequencing, colony-PCR and qRT-PCRMEP2 fusion gene encodes for a high-affinity ammonium permease, which may be adaptive under nitrogen-poor environments with ammonium as only nitrogen source~20 mya (Kellis et al., 2003)de novo creation of hybrid followed by continuous ammonium limitationExperimental populations
Albugo candida
Between the ancestors of three races; an adaptive radiation in progress (McMullan et al., 2015)Whole genome sequencing, RNAseqHybridCheckEffector alleles that may result in a fitness advantage on most potential hosts, enabling host jumpsMultiple time points; oldest event ~200,000 ya, most recent event ~2,200 yaInfection with virulent A. candida suppresses host immunity enabling co-colonization by otherwise non-virulent races
Hybrid partnersGenomic data typesMethods appliedAdaptive traits or genes influencedStart of divergence of the hybridizing taxaEstimated time of secondary contactPresence of hybrid zone(s)
Arabidopsis
Between Arabidopsis arenosa and A. lyrata (Schmickl and Koch, 2011; Arnold et al., 2016)Whole genome sequencing (Arnold et al., 2016)Fastsimcoal2, fD statistic, HybridCheck; genome scanGenes controlling specific ion homeostasis-related and drought adaptation traits (Arnold et al., 2016)~1.63 mya
(Hohmann et al., 2015)
Eastern Austrian Forealps (Schmickl and Koch, 2011)
Sunflowers
From Helianthus debilis subsp. cucumerifolius into H. annuus subsp. annuus (Whitney et al., 2015)SNP dataset (QTL mapping)Primarily abiotic tolerance traits allowing rapid growth and reproduction before summer heat and drought, but also traits affecting herbivore resistanceDivergence between the ancestors of H. debilis and H. annuus ~2 mya (Renaut et al., 2013)Maximum ~18,000 ya (Scascitelli et al., 2010)South-central USA
Senecio
From diploid Senecio squalidus into tetraploid S. vulgaris (Kim et al., 2008)[No large genome- wide dataset] Sanger sequencing, RNA in situ hybridizationGenes RAY1 and RAY2 that control ray florets in radiate flower heads, which effects attractiveness to pollinators and outcrossing rateS. squalidus is a homoploid hybrid with an origin less than ~150,000 ya (Filatov et al., 2016)Maximum ~300 ya (S. squalidus was introduced to the Oxford Botanic Garden around 300 ya)United Kingdom
Poplars
From Populus balsamifera into P. trichocarpa (Suarez- Gonzalez et al., 2016)Whole genome sequencing, RNAseqGenome scanNumerous genes, particularly in a telomeric region (PRR5 and COMT1); related to response to far red light, RNA processing and ATPase activity~76,000 ya (Levsen et al., 2012)Northwestern North America
Mussels
From Mytilus edulis into M. galloprovincialis (Fraïsse et al., 2014; Fraïsse et al., 2016)AFLP and few other markers, Sanger sequencing (Fraïsse et al., 2014); target enrichment (Fraïsse et al., 2016)Genome scanActin gene mac-1 (Fraïsse et al., 2014; but not Fraïsse et al., 2016)~2.5 mya (Roux et al., 2014)~0.7 mya (Roux et al., 2014)In Europe from the Mediterranean Sea to the North Sea
Drosophila
From Drosophila simulans into D. sechellia (Garrigan et al., 2012; Brand et al., 2013)Garrigan et al., 2012: whole genome sequencing; Brand et al., 2013: Sanger sequencing of the candidate region identified by Garrigan et al., 2012 Genome scan (Garrigan et al., 2012)A 15 kb region (Garrigan et al., 2012) – the target of selection appears to be a regulatory sequence; evidence for a trans-specific selective sweep (Brand et al., 2013), i.e., the 15 kb region harbours a polymorphism that is adaptive in both species~200,000–300,000 ya (Garrigan et al., 2012)2,600–18,600 yaSeychelles
From D. yakuba into D. santomea (Llopart et al., 2014; Beck et al., 2015)Llopart et al., 2014: Sanger sequencing and high-throughput sequencing of the mitochondrial genome; Beck et al., 2015: Sanger sequencingPotential cytonuclear co-introgression of genes involved in the oxidative phosphorylation pathway; strong signal of introgression in the three nuclear genes composing subunit V of the cytochrome c oxidase complex (Beck et al., 2015)~400,000 ya (Llopart et al., 2002)~14,000 ya (complete replacement of the mitochondrial genome) and ~2,500 years ago (Llopart et al., 2014)São Tomé (invasion by D. yakuba)
Anopheles
Between Anopheles arabiensis and the ancestor of A. gambiae and A. coluzzii
(Fontaine et al., 2015)
Whole genome sequencingD and DFOIL statistics, window-based gene tree distribution across chromosomesInsecticide-resistance alleles might have introgressed into A. arabiensis with the 2La inversion introgressed from the ancestor of A. gambiae and A. coluzziiDivergence between A. gambiae and A. coluzzii from its ancestor ~2 mya~0.5 myaCentral Africa
From A. gambiae into A. coluzzii (Norris et al., 2015)SNP genotyping, whole genome sequencingTraits enhancing vectorial capacity: A. coluzzii inherited the entire A. gambiae- associated 2L divergence island including the knockdown resistance SNP L1014F in the kdr gene (L1014F has been present in A. gambiae populations for at least 20 years)Periods of isolation interrupted by hybridization; recent introgression reported in 2002 and 2006 (Lee et al., 2013b )Selinkenyi, Mali
Heliconius butterflies
Other studies that focused on speciation with ongoing gene flow and adaptive introgression: e.g. Pardo-Diaz et al., 2012; The Heliconius Genome Consortium, 2012; Nadeau et al., 2013; Kronforst et al., 2013; Martin et al., 2013)
Between subspecies of Heliconius erato and H. melpomene (Nadeau et al., 2014)RAD-seqStructure; genome scanLoci that control wing color patterns (for both mimicry and mate recognition); the aim of this study was to identify novel loci underlying phenotypic variation by means of association mapping in hybrid zones~11.8 mya (Kozak et al., 2015)Several time points post speciation (Martin et al., 2013)Parallel zones in the Andes
Between H. beschkei, H. numata and H. numata subsp. nanna (Zhang et al., 2016)Whole genome sequencingD and fD statisticsGenomic region upstream of the gene optix (known to control red wing patterning) and 39 additional genomic regionsDivergence between H. beschkei and H. numata ~1.93 myaEast-central South America
Darwin’s finches
Species from different Galápagos islands
(Lamichhaney et al., 2015)
Whole genome sequencingPSMC, NeighborNet, D statistic; genome scanALX1 gene encodes a transcription factor controlling craniofacial development, in particular beak shape~0.5 myaSeveral time points post speciationGalápagos islands
Mice
From Mus spretus into M. musculus subsp. domesticus (Song et al., 2011)[No large genome- wide dataset] Sanger sequencing>10 Mb region including the molecular target of anticoagulants vkorc1; M. spretus-specific alleles can cause resistance to rodenticides~1.5–3 myaAfrica and Europe
Between M. musculus subsp. domesticus and M. musculus subsp. musculus
(Staubach et al., 2012)
SNP genotypingGenome scanNo general pattern of types of genes subject to introgression300,000–500,000 ya (Guénet and Bonhomme, 2003)Maximum ~3,000 ya (Cucchi et al., 2005)Eastern and southern Germany, western Czech Republic
Humans
Other studies identified genomic regions with evidence of adaptive introgression between archaic and modern humans (from Neanderthals into modern humans: e.g., Vernot and Akey, 2014; Sankararaman et al., 2014 / from Denisovans into Tibetans: e.g., Jeong et al., 2014; Huerta-Sánchez et al., 2014)
From Neanderthals and Denisovans into modern humans (Racimo et al., 2017)Whole genome sequencing; reevaluation of published dataD and fD statistics, RD, U and Q95 statistics, HaplostripsLipid metabolism, pigmentation and innate immunityDivergence between the Neanderthal/ Denisovan lineage and modern humans 170,000–700,000 ya (Meyer et al., 2012)Between Neanderthals and modern humans 47,000– 65,000 ya (Sankararaman et al., 2012)Eurasia
Yeast
Between Saccharomyces cerevisiae and S. uvarum (Dunn et al., 2013)Whole genome sequencing, colony-PCR and qRT-PCRMEP2 fusion gene encodes for a high-affinity ammonium permease, which may be adaptive under nitrogen-poor environments with ammonium as only nitrogen source~20 mya (Kellis et al., 2003)de novo creation of hybrid followed by continuous ammonium limitationExperimental populations
Albugo candida
Between the ancestors of three races; an adaptive radiation in progress (McMullan et al., 2015)Whole genome sequencing, RNAseqHybridCheckEffector alleles that may result in a fitness advantage on most potential hosts, enabling host jumpsMultiple time points; oldest event ~200,000 ya, most recent event ~2,200 yaInfection with virulent A. candida suppresses host immunity enabling co-colonization by otherwise non-virulent races

Introgression between divergent populations or closely related species does not generally homogenize divergence levels across the entire genome (Turner et al., 2005; Coleman et al., 2006; Harr, 2006; Michel et al., 2010; Hohenlohe et al., 2012; Via, 2012; Renaut et al., 2013; Malinsky et al., 2015; Marques et al., 2016; reviewed in Wolf and Ellegren, 2017). Genomic islands of divergence (GIsD), also known as genomic islands of speciation, manifest as regions of elevated divergence in a ‘sea’ of neutral, non-differentiated background that has been homogenized by a history of gene flow. These GIsD are typically attributed to loci under divergent selection contributing to adaptation to the local environment or the establishment of reproductive isolation relatively independent of the external environment (Wu and Ting, 2004; Orr et al., 2004; Rieseberg and Blackman, 2010; Nosil and Schluter, 2011). One example is the adaptation to different foraging behaviours in Darwin’s finches, which might have largely been driven by the ALX1 gene that encodes a transcription factor affecting craniofacial development. Variation in this gene is strongly associated with beak shape diversity across Darwin’s finches and the medium ground finch (Lamichhaney et al., 2015). GIsD have been reported in many cases of recent species divergence with ongoing gene flow (Martin et al., 2013; Jónsson et al., 2014; Supple et al., 2015; Rougeux et al., 2016; Royer et al., 2016; Morales et al., 2017; Kumar et al., 2017). The region of divergence typically extends away from the selected locus due to physical linkage, allowing neutral polymorphisms to hitchhike along with a selected polymorphism, namely divergence hitchhiking, and become part of the GIsD. In many cases, this can obscure signals of selection and lead to ambiguity over the exact genetic loci under selection. It is therefore advisable to incorporate linked selection as a null model for the identification of genomic regions exhibiting pronounced differentiation (Burri et al., 2015).

Low recombination rates in some genomic regions can also create GIsD. Regions of suppressed recombination are frequently pericentromeric and often involve structural changes such as inversions that can physically block recombination. They are thought to sometimes play a role in adaptive genomic divergence (Rieseberg, 2001; Noor et al., 2001; Feder and Nosil, 2009), for example in the case of the 2L inversion divergence island, which was introgressed from Anopheles gambiae into A. coluzzii and harbours a suite of insecticide-resistance alleles (Lee et al., 2013a; Norris et al., 2015). However, it is questionable if most regions of very low recombination contain loci under divergent selection or whether they were simply established due to the stochastic effects of genome structure and genetic drift (Turner and Hahn, 2010). If the amount of gene flow is very low, for example due to geographic isolation, then large blocks of pronounced genomic differentiation may arise from genetic drift. This could be further accentuated by variable mutation rates and low recombination rates (Noor and Bennett, 2009; White et al., 2010; Cruickshank and Hahn, 2014).

Traits impacted by adaptive introgression

The traits impacted by introgression are as diverse as the organisms that have been studied (Table 1). In plants, for example, a broad array of abiotic tolerance traits have been affected, for example ion homeostasis and drought adaptations (Arabidopsis arenosa, Arnold et al., 2016; sunflowers, Whitney et al., 2015), but also traits that control biotic interactions between plants and herbivores (sunflowers, Whitney et al., 2015) or pollinators (Senecio, RAY1 and RAY2 genes, Kim et al., 2008), and others (poplars, PRR5 and COMT1 genes, Suarez-Gonzalez et al., 2016). In animals, clear examples that include traits influenced by adaptive introgression include insecticide resistance in Anopheles, the single nucleotide polymorphism L1014F in the gene kdr (Fontaine et al., 2015; Norris et al., 2015) and rodenticide resistance in mice, the vkorc1 gene (Song et al., 2011). Other introgressed alleles of potentially adaptive value stem from the actin gene mac-1 in mussels (Fraïsse et al., 2014), several loci that control wing colour patterns for both mimicry and mate recognition, amongst them the optix locus in Heliconius butterflies (Nadeau et al., 2014; Zhang et al., 2016), the beak shape-associated locus ALX1 in Darwin’s finches (Lamichhaney et al., 2015), variants that control lipid metabolism, pigmentation and innate immunity in humans (Racimo et al., 2017), but also loci of so far uncharacterized adaptive value in Drosophila (Garrigan et al., 2012; Brand et al., 2013) and mice (Staubach et al., 2012). An interesting case of introgression between Drosophila species was reported to be trans-specific, meaning that the introgressed 15 kb region harbours a mutation that appears adaptive in both species despite their different genomic backgrounds and ecological requirements (Brand et al., 2013). In addition, cytonuclear co-introgression has been shown for some Drosophila species (Beck et al., 2015). In the fungal genus Saccharomyces experimental hybridisation followed by ammonium limitation resulted in the origin of an interspecific MEP2 fusion gene, which encodes for a high affinity ammonium permease that may be adaptive under nitrogen-poor environments with ammonium as the only nitrogen source (Dunn et al., 2013).

Adaptive allele mining in hybrid zones

In secondary contact scenarios, PaIS come into contact following a period of reproductive isolation. This process is often induced by geographic and environmental factors, such as habitat shifts due to climate changes, for example oscillations during Pleistocene and Anthropocene, or catastrophic human-mediated events such as fires, flooding, road construction or introduction of alien species (Crispo et al., 2011). Such secondary contact between divergent but interfertile populations frequently results in hybrid zones where the otherwise geographically distinct distribution ranges of the two species overlap, permitting the production of offspring of mixed ancestry (Barton and Hewitt, 1985; Barton and Hewitt, 1989; Harrison, 1990). In the fungus-like pathogen Albugo candida a spectacular example of adaptive introgression after secondary contact has recently been reported, in which infection of a host with virulent A. candida suppresses host immunity enabling co-colonization by otherwise non-virulent races (McMullan et al., 2015). This creates hybrid zones allowing sexual reproduction and gene flow between isolated races; effector alleles are transferred that exhibit a fitness advantage on many potential hosts, facilitating host jumps. Then once a new hybrid race has established, it rapidly reproduces asexually without continued exchange with other races.

Hybrid zones represent natural laboratories (Hewitt, 1988) for the study of introgression dynamics as well as their evolutionary significance. They offer venues to observe allele flow between partly isolated populations and allow hybrid fitness to be estimated in nature. As described above, alleles under divergent selection are relatively inhibited from introgressing between populations. This barrier to the exchange of certain alleles generates an allele frequency gradient, or cline, across the hybrid zone, which partly blocks the homogenization of the hybridizing populations. Such a cline might be detected as a phenotypic transition from one parental species to the other, which should coincide with the underlying allelic cline. Hybrid zones can be understood as natural mapping experiments to determine the alleles responsible for complex traits that differentiate parental populations (Buerkle and Lexer, 2008; Crawford and Nielsen, 2013). Over generations, recombination breaks down linkage, allowing alleles at discrete loci to be associated with phenotypes and environmental variables (Crawford and Nielsen, 2013). This is especially clear in large hybrid populations with varying levels of admixture that are established along environmental gradients, such as in spruce (Hamilton et al., 2013; Hamilton et al., 2015). Substantial levels of clinal variation and allele-environment associations with climatic variables such as temperature and precipitation were found in hybrid zones of Sitka, white, and Engelmann spruce (Hamilton et al., 2015), which suggests that species integrity is maintained through exogenous selection in parental habitats and that hybridisation might facilitate fine-scale adaptation of the species along environmental gradients.

However, genotype-phenotype-environment associations along clines have rarely been addressed using dense genome-wide markers. Highlights include work in mice (Turner and Harr, 2014; Pallares et al., 2014; Pallares et al., 2016), for which genes involved in craniofacial shape variation were found that act in a polygenic manner (Pallares et al., 2016). The Heliconius and Helianthus systems are other examples of association mapping across hybrid zones, although Helianthus showed spurious associations in early generation hybrids due to linkage disequilibrium (Rieseberg and Buerkle, 2002). Heliconius in contrast is a relatively old hybrid zone and close to linkage equilibrium, with phenotypic variation largely controlled by major effect loci, for which smaller sample sizes are sufficient (Nadeau et al., 2014). With emerging probabilistic frameworks to infer allele frequencies in low coverage sequencing data it should be possible in the future to also address minor effect loci, for which larger sample sizes are required, even in polyploids (Blischak et al., 2017).

Selective landscapes vary widely in different types of hybrid zones. If selection acts on genotypes of mixed ancestry independent of spatial variation in selection pressures, the hybrid zone can be considered a ‘tension zone’ (Barton and Hewitt, 1985; Barton and Hewitt, 1989). In a tension zone selection is usually endogenous, for example the hybrid zone between two Senecio species on Mount Etna (Brennan et al., 2009), with occasional exceptions such as exogenous frequency-dependent selection by predatory birds in Heliconius butterflies (Mallet et al., 1990). Alternatively, Endler (1977) postulated spatially varying exogenous selection along a geographic-environmental gradient. According to another model, the bounded hybrid superiority model (Moore, 1977), selection in intermediate habitats can favour individuals of mixed ancestry, which seems to be the case for the hybrid zone between Sitka and white spruce (Hamilton et al., 2013). In mosaic hybrid zones, parental populations are distributed in a patchy landscape. Examples of such hybrid zones are frequently found in parasites that rely on patchily distributed hosts, for example in Albugo candida (McMullan et al., 2015), but also in non-parasitic species such as mussels (Fraïsse et al., 2014), which may reflect a common scenario of complex environmental mosaics in nature (Harrison, 1986). In a tension zone, with endogenous selection on the hybrids, the hybrid zone can move freely if there are asymmetries in selection, dispersal, or population density and will finally arrest at a geographic barrier or in an area of low population density (Hewitt, 1975; Barton, 1979; Barton and Hewitt, 1985). Hybrid zones might move in response to climate change or co-varying environmental factors (Taylor et al., 2015).

Population genomic approaches to infer introgression in natural populations

A convergence of recent work on new methods of refined demographic inference and methods to describe the genomic architecture of introgression have great potential for the discovery of introgressed alleles (Table 2). Demographic inference methods evaluate population structure. They provide a good estimate of admixture between populations. They estimate fine-scale demographic histories, such as population history, including fluctuations in effective population size over evolutionary time, partly based on rare polymorphisms, or they infer parameter values, such as population split times and migration rates, and test hypotheses for comparison with alternative demographic scenarios. The genomic architecture of introgression is observed using metrics that determine the level of allele sharing and linkage disequilibrium, along with visualization methods, including topology weighting of trees across the genome (Table 2). Genome scans assay for regions exhibiting unusual levels of divergence among populations or species that represent candidate selected alleles (Nielsen et al., 2005; Yant et al., 2013; Lotterhos and Whitlock, 2015; Jensen et al., 2016). There are two main approaches to identify candidate loci: taking outliers of differentiation metrics or performing explicit hypothesis testing to determine whether the value is significantly greater than expected by chance or under neutrality. Genome scans can be confounded by other modes of selection and by effects of demography and are therefore best interpreted as hypothesis generators, providing candidate loci and processes to be tested in downstream functional studies, where possible. It may also be more confidently concluded that introgression is adaptive if a combination of methods, including descriptors of the genomic architecture of introgression and appropriately chosen selection metrics, converge on particular loci. In some cases, compelling a priori evidence that the trait is under similar directional selection in both species is also available. The combination of using methods assessing the genomic architecture of introgression and population genomics metrics was successfully applied when mining for adaptively introgressed candidate alleles for mediating adaptation to serpentine soils in Arabidopsis (Arnold et al., 2016). In Heliconius butterflies there was a priori evidence for wing colour patterns involved in mimicry and mate recognition to be under directional selection in the hybridising species, and it was then shown that the introgressed alleles are responsible for this trait (Pardo-Diaz et al., 2012).

Table 2.

Useful population genomic and phylogenomic approaches to detect and interpret introgression events

Analysis TypeSoftwareDescriptionReferences
Demographic inferencePopulation structure
adegenetEvaluation of population clustering using principal component and discriminant analysisJombart, 2008; Jombart et al., 2010
STRUCTURE;
fastSTRUCTURE;
fineSTRUCTURE;
ADMIXTURE;
SpaceMix
Estimation of (fine-scale) population structure, taking admixture into accountPritchard et al., 2000; Hubisz et al., 2009;
Raj et al., 2014;
Lawson et al., 2012;
Alexander et al., 2009;
Bradburd et al., 2016
Demographic history
IM; IMa2Fitting an isolation with migration model to haplotype data from two populations (IM) or up to ten populations (IMa2)Hey and Nielsen, 2007; Hey J. 2010a ; Hey J. 2010b
dadi (diffusion approximations for demographic inference)Inference of the demographic history of multiple populations from SNP frequency dataGutenkunst et al., 2009
PSMC (pairwise sequentially Markovian coalescent model); MSMC (multiple sequentially Markovian coalescent model)Inference of fluctuations in effective population size over time from a single genome sequence (PSMC) or multiple genome sequences (MSMC)Li and Durbin, 2011; Schiffels and Durbin, 2014
simcoal2; fastsimcoal2Inference of parameter values, such as population split times and migration rates, and testing of hypotheses to compare to alternative, neutral demographic scenariosLaval and Excoffier, 2004; Excoffier et al., 2013
RarecoalInference of population history and fine-scale ancestry from rare variantsSchiffels et al., 2016
Network and tree methods
NeighbourNet (SplitsTree4)Visualization of reticulate relationships in the form of splits networksHuson and Bryant, 2006
TreeMixInference of patterns of population splits and admixturePickrell and Pritchard, 2012
Phylo-NetCoalescent-based species tree and evolutionary network method; testing the number of introgression eventsThan et al., 2008
Genomic architecture of introgressionIntrogression detection
D statistic; f statisticDetection of SNP window-based or genome-wide evidence of shared alleles (ABBA-BABA test) in a four-taxon caseComparison of two genomes: Green et al., 2010; Durand et al., 2011 / comparison of two populations: Kronforst et al., 2013; Smith and Kronforst, 2013
DFOIL statisticD statistic for a symmetric five-taxon phylogeny; determination of the directionality of introgressionPease and Hahn, 2015
fD statisticRefinement of the f statistic (Green et al., 2010) by being less sensitive to differences in diversity along the genomeMartin et al., 2015
RD, U, Q95 statisticsIdentification of genomic windows that are likely to have undergone adaptive introgressionRacimo et al., 2017
Methods with additional visualization aspect
HybridCheckDetection of introgressed genomic blocks and visualization of the heterogeneous, mosaic-like genome structure; dating of introgressed blocksWard and van Oosterhout, 2016
HaplostripsPlot of haplotype structure at candidate regions for adaptive introgressionMarnetto et al., in prep. (cited in Racimo et al., 2017)
TwisstTopology weighting of SNP window-based trees across the genomeMartin and Van Belleghem, 2017
Analysis TypeSoftwareDescriptionReferences
Demographic inferencePopulation structure
adegenetEvaluation of population clustering using principal component and discriminant analysisJombart, 2008; Jombart et al., 2010
STRUCTURE;
fastSTRUCTURE;
fineSTRUCTURE;
ADMIXTURE;
SpaceMix
Estimation of (fine-scale) population structure, taking admixture into accountPritchard et al., 2000; Hubisz et al., 2009;
Raj et al., 2014;
Lawson et al., 2012;
Alexander et al., 2009;
Bradburd et al., 2016
Demographic history
IM; IMa2Fitting an isolation with migration model to haplotype data from two populations (IM) or up to ten populations (IMa2)Hey and Nielsen, 2007; Hey J. 2010a ; Hey J. 2010b
dadi (diffusion approximations for demographic inference)Inference of the demographic history of multiple populations from SNP frequency dataGutenkunst et al., 2009
PSMC (pairwise sequentially Markovian coalescent model); MSMC (multiple sequentially Markovian coalescent model)Inference of fluctuations in effective population size over time from a single genome sequence (PSMC) or multiple genome sequences (MSMC)Li and Durbin, 2011; Schiffels and Durbin, 2014
simcoal2; fastsimcoal2Inference of parameter values, such as population split times and migration rates, and testing of hypotheses to compare to alternative, neutral demographic scenariosLaval and Excoffier, 2004; Excoffier et al., 2013
RarecoalInference of population history and fine-scale ancestry from rare variantsSchiffels et al., 2016
Network and tree methods
NeighbourNet (SplitsTree4)Visualization of reticulate relationships in the form of splits networksHuson and Bryant, 2006
TreeMixInference of patterns of population splits and admixturePickrell and Pritchard, 2012
Phylo-NetCoalescent-based species tree and evolutionary network method; testing the number of introgression eventsThan et al., 2008
Genomic architecture of introgressionIntrogression detection
D statistic; f statisticDetection of SNP window-based or genome-wide evidence of shared alleles (ABBA-BABA test) in a four-taxon caseComparison of two genomes: Green et al., 2010; Durand et al., 2011 / comparison of two populations: Kronforst et al., 2013; Smith and Kronforst, 2013
DFOIL statisticD statistic for a symmetric five-taxon phylogeny; determination of the directionality of introgressionPease and Hahn, 2015
fD statisticRefinement of the f statistic (Green et al., 2010) by being less sensitive to differences in diversity along the genomeMartin et al., 2015
RD, U, Q95 statisticsIdentification of genomic windows that are likely to have undergone adaptive introgressionRacimo et al., 2017
Methods with additional visualization aspect
HybridCheckDetection of introgressed genomic blocks and visualization of the heterogeneous, mosaic-like genome structure; dating of introgressed blocksWard and van Oosterhout, 2016
HaplostripsPlot of haplotype structure at candidate regions for adaptive introgressionMarnetto et al., in prep. (cited in Racimo et al., 2017)
TwisstTopology weighting of SNP window-based trees across the genomeMartin and Van Belleghem, 2017
Table 2.

Useful population genomic and phylogenomic approaches to detect and interpret introgression events

Analysis TypeSoftwareDescriptionReferences
Demographic inferencePopulation structure
adegenetEvaluation of population clustering using principal component and discriminant analysisJombart, 2008; Jombart et al., 2010
STRUCTURE;
fastSTRUCTURE;
fineSTRUCTURE;
ADMIXTURE;
SpaceMix
Estimation of (fine-scale) population structure, taking admixture into accountPritchard et al., 2000; Hubisz et al., 2009;
Raj et al., 2014;
Lawson et al., 2012;
Alexander et al., 2009;
Bradburd et al., 2016
Demographic history
IM; IMa2Fitting an isolation with migration model to haplotype data from two populations (IM) or up to ten populations (IMa2)Hey and Nielsen, 2007; Hey J. 2010a ; Hey J. 2010b
dadi (diffusion approximations for demographic inference)Inference of the demographic history of multiple populations from SNP frequency dataGutenkunst et al., 2009
PSMC (pairwise sequentially Markovian coalescent model); MSMC (multiple sequentially Markovian coalescent model)Inference of fluctuations in effective population size over time from a single genome sequence (PSMC) or multiple genome sequences (MSMC)Li and Durbin, 2011; Schiffels and Durbin, 2014
simcoal2; fastsimcoal2Inference of parameter values, such as population split times and migration rates, and testing of hypotheses to compare to alternative, neutral demographic scenariosLaval and Excoffier, 2004; Excoffier et al., 2013
RarecoalInference of population history and fine-scale ancestry from rare variantsSchiffels et al., 2016
Network and tree methods
NeighbourNet (SplitsTree4)Visualization of reticulate relationships in the form of splits networksHuson and Bryant, 2006
TreeMixInference of patterns of population splits and admixturePickrell and Pritchard, 2012
Phylo-NetCoalescent-based species tree and evolutionary network method; testing the number of introgression eventsThan et al., 2008
Genomic architecture of introgressionIntrogression detection
D statistic; f statisticDetection of SNP window-based or genome-wide evidence of shared alleles (ABBA-BABA test) in a four-taxon caseComparison of two genomes: Green et al., 2010; Durand et al., 2011 / comparison of two populations: Kronforst et al., 2013; Smith and Kronforst, 2013
DFOIL statisticD statistic for a symmetric five-taxon phylogeny; determination of the directionality of introgressionPease and Hahn, 2015
fD statisticRefinement of the f statistic (Green et al., 2010) by being less sensitive to differences in diversity along the genomeMartin et al., 2015
RD, U, Q95 statisticsIdentification of genomic windows that are likely to have undergone adaptive introgressionRacimo et al., 2017
Methods with additional visualization aspect
HybridCheckDetection of introgressed genomic blocks and visualization of the heterogeneous, mosaic-like genome structure; dating of introgressed blocksWard and van Oosterhout, 2016
HaplostripsPlot of haplotype structure at candidate regions for adaptive introgressionMarnetto et al., in prep. (cited in Racimo et al., 2017)
TwisstTopology weighting of SNP window-based trees across the genomeMartin and Van Belleghem, 2017
Analysis TypeSoftwareDescriptionReferences
Demographic inferencePopulation structure
adegenetEvaluation of population clustering using principal component and discriminant analysisJombart, 2008; Jombart et al., 2010
STRUCTURE;
fastSTRUCTURE;
fineSTRUCTURE;
ADMIXTURE;
SpaceMix
Estimation of (fine-scale) population structure, taking admixture into accountPritchard et al., 2000; Hubisz et al., 2009;
Raj et al., 2014;
Lawson et al., 2012;
Alexander et al., 2009;
Bradburd et al., 2016
Demographic history
IM; IMa2Fitting an isolation with migration model to haplotype data from two populations (IM) or up to ten populations (IMa2)Hey and Nielsen, 2007; Hey J. 2010a ; Hey J. 2010b
dadi (diffusion approximations for demographic inference)Inference of the demographic history of multiple populations from SNP frequency dataGutenkunst et al., 2009
PSMC (pairwise sequentially Markovian coalescent model); MSMC (multiple sequentially Markovian coalescent model)Inference of fluctuations in effective population size over time from a single genome sequence (PSMC) or multiple genome sequences (MSMC)Li and Durbin, 2011; Schiffels and Durbin, 2014
simcoal2; fastsimcoal2Inference of parameter values, such as population split times and migration rates, and testing of hypotheses to compare to alternative, neutral demographic scenariosLaval and Excoffier, 2004; Excoffier et al., 2013
RarecoalInference of population history and fine-scale ancestry from rare variantsSchiffels et al., 2016
Network and tree methods
NeighbourNet (SplitsTree4)Visualization of reticulate relationships in the form of splits networksHuson and Bryant, 2006
TreeMixInference of patterns of population splits and admixturePickrell and Pritchard, 2012
Phylo-NetCoalescent-based species tree and evolutionary network method; testing the number of introgression eventsThan et al., 2008
Genomic architecture of introgressionIntrogression detection
D statistic; f statisticDetection of SNP window-based or genome-wide evidence of shared alleles (ABBA-BABA test) in a four-taxon caseComparison of two genomes: Green et al., 2010; Durand et al., 2011 / comparison of two populations: Kronforst et al., 2013; Smith and Kronforst, 2013
DFOIL statisticD statistic for a symmetric five-taxon phylogeny; determination of the directionality of introgressionPease and Hahn, 2015
fD statisticRefinement of the f statistic (Green et al., 2010) by being less sensitive to differences in diversity along the genomeMartin et al., 2015
RD, U, Q95 statisticsIdentification of genomic windows that are likely to have undergone adaptive introgressionRacimo et al., 2017
Methods with additional visualization aspect
HybridCheckDetection of introgressed genomic blocks and visualization of the heterogeneous, mosaic-like genome structure; dating of introgressed blocksWard and van Oosterhout, 2016
HaplostripsPlot of haplotype structure at candidate regions for adaptive introgressionMarnetto et al., in prep. (cited in Racimo et al., 2017)
TwisstTopology weighting of SNP window-based trees across the genomeMartin and Van Belleghem, 2017

Horizontal gene transfer events as a source of adaptive novelty

Alleles conferring an adaptive advantage may also be exchanged between reproductively isolated species via HGT. In prokaryotes HGT between distantly related species is well established as a source of novelty and a key driver of adaptation, most notably in the acquisition of antibiotic resistance and genes conferring pathogenicity (Gillings, 2017). Similarly, HGT has been shown to be prevalent among single-celled eukaryotes (Keeling and Palmer, 2008; Andersson, 2009). This penchant for HGT in prokaryotes has an ongoing effect on multicellular eukaryotes; HGT between the mitochondria of distantly related plant species is rampant (Won and Renner, 2003; Bergthorsson et al., 2003; Mower et al., 2010; Rice et al., 2013;). HGT in the plastid genome seems relatively infrequent, although there are cases (Rice et al., 2006; Park et al., 2007). This may be due to the fact that plant mitochondria have a mechanism for the active uptake of DNA and frequently fuse (Richardson and Palmer, 2007; Rice et al., 2013), while plastids lack this tendency. HGT into the nuclear genomes of multicellular eukaryotes is less frequent. For any foreign DNA to be heritable it must be integrated into the germline, which is separated from the somatic cells and often protected from the environment by elaborate structures. Despite this, HGT does occur in the nuclear genomes of multicellular eukaryotes; fungi (Ambrose et al., 2014), arthropods (Wybouw et al., 2016), nematodes (Danchin et al., 2010), mosquitos (Klasson et al., 2009), fish (Sun et al., 2015), sea anemones (Starcevic et al., 2008) and a broad range of plants (Bock, 2010) have all been shown to contain nuclear genes of HGT origin from diverse sources.

Many possible mechanisms for genetic transfers by HGT have been suggested. Lifestyle traits that allow intimate contact between unrelated species may increase the likelihood of HGT. For example, close contact between a host and its parasite, which can involve the exchange of macromolecules including mRNAs (Kim et al., 2014), presents an opportunity for genetic exchange (Yoshida et al., 2010; Xi et al., 2012; Zhang et al., 2013; Zhang et al., 2014; Davis and Xi, 2015; Yang et al., 2016) as does a reproductive cycle in which components of the germline are more exposed to the environment or even free living, such as those of bryophytes, lycophytes, and ferns (Li et al., 2014). It has also been demonstrated that rare cases of grafting between unrelated species can result in the transformation of cells at the graft site (Bergthorsson, 2003; Stegemann, 2009; Stegemann et al. 2012). There are also many vectors that may transport genetic material between species, for example bacteria, viruses or mobile genetic elements. It has been established that pathogenic bacteria can transform eukaryotic host cells through the injection of proteins and/or genetic material (Lacroix and Citovsky, 2016). HGT of selfish genetic elements, such as transposons, is rampant and could mediate the movement of host DNA. Indeed, a study of group I introns in angiosperm mitochondria found 32 separate instances of HGT into plants (Cho et al., 1998). Another study showed that 65% of the plant genomes analyzed contained at least one instance of HGT of a long terminal repeat retrotransposon (El Baidouri et al., 2014). This likely has an important evolutionary function for selfish genetic elements, allowing them to escape resistant host genomes that effectively silence them. Finally, despite the complexity involved in shielding gametes from the environment in seed plants, it has been suggested that exposure to foreign pollen could result in small windows of opportunity for illegitimate pollination and HGT into the germline (Keeling and Palmer, 2008; Bock, 2010; Christin et al., 2012).

A key outstanding question, however, is what, if any, is the adaptive impact of these HGT events? Arguably genes that do not confer an adaptive benefit are expected to decay by neutral drift, eventually to be relegated to pseudogene status and then lost altogether (Keeling and Palmer, 2008; Soucy, 2015). Many factors could render a gene acquired by HGT from a divergent organism useless, or even deleterious. A novel genetic background could abrogate interactions essential for gene function, while incompatible and divergent codon biases, transcriptional elements or intron/exon splice sites could also inactivate foreign genes. Indeed, the fate of many genes acquired by HGT in eukaryotes is decay (Bergthorsson, 2003; Mower et al., 2010; Rice, 2013,Mahelka et al., 2017). However, a large body of evidence suggests that this is not always the case (Emiliani et al., 2009; Danchin et al., 2010; Yue et al., 2012; Christin et al., 2012; Acuña et al., 2012; Zhang et al., 2013; Yang et al., 2013; Li et al., 2014; Ambrose et al., 2014; Prentice et al., 2015; Sun et al., 2015; Yin et al., 2016). The transcriptional and developmental regulation of horizontally transferred loci, alongside evidence for purifying or positive selection, suggests that many such events are adaptive.

Phylogenetic analysis is the gold standard for HGT inference (Keeling and Palmer, 2008; Brock, 2009; Soucy, 2015). It relies on detecting incongruence between individual locus trees and species phylogenies. Often these cases are obvious. However, alternative evolutionary processes as well as sampling and analytic errors can produce incongruent gene trees and these must be considered in each case. Several potential confounding causes include gene duplication and subsequent loss, inadequate taxonomic sampling, historical allopolyploidization and long branch attraction (Keeling and Palmer, 2008; Brock, 2009; Soucy 2015). Further the possibility exists of intracellular gene transfer from the mitochondrial and plastid genomes; thus, genes of alpha-proteobacterial or cyanobacterial origin should be excluded from analysis (Huang and Gogarten, 2008; Yang et al., 2013). However, diverse evidence classes can corroborate candidate HGT events, such as codon usage differences, intron structure or GC content.

HGT and the emergence of land plants

Strong evidence implicates ancient HGT as a key source of adaptive novelty when the pioneering ancestor of green plants adapted to terrestrial environments (Huang and Gogarten, 2008; Emiliani et al., 2009; Yue et al., 2012; Yue et al., 2013; Yang et al., 2013). This lineage experienced an array of novel abiotic and biotic challenges upon colonisation of land, including desiccation, UV irradiation, and microbial attack. Analysis of the moss Physcomitrella patens identified 39 gene families that had been acquired by HGT from prokaryotes, fungi or viruses after the split between plants and green algae, 35 of which were shared with seed plants. These loci are involved in a broad range of plant-specific processes including biosynthesis, defence, stress tolerance, vascular development, and seed germination (Yue et al., 2012; Yue et al., 2013). In contrast to seed plants, the gametophytes and zygotes of mosses are more exposed to the environment, presenting an opportunity for the integration of foreign DNA. Another example is the phenylpropanoid pathway that produces compounds such as lignin and flavonoids, critical components in plant structure and defence against microbes and UV (Emiliani et al., 2009). The common ancestor of land plants acquired the enzyme phenylpropanoid that performs the first critical step in this pathway via HGT from what was ultimately a bacterial source (Emiliani et al., 2009). Another example is represented by the L-Ala-D/L-Glu epimerases (AEEs), which are ubiquitous in land plants and were initially acquired by HGT from prokaryotes (Yang et al., 2013). The fixation of AEEs in land plants was driven by positive selection and was specific to land plants; they have not been found in any other eukaryotes including the extant progenitors of land plants, red and green algae (Yang et al., 2013). Acquisition of genes by HGT prior to the split between red alga and green plants is also an important source of genes involved in the functionality of plastids (Huang and Gogarten, 2008).

Box 1: Established and promising model systems for adaptive introgression

It has long been the subject of debate whether introgression is more common in plants than other kingdoms (Mallet, 2005; Mallet, 2007). Introgression may be less common in animals compared with plants due to stronger assortative mating and lower F1 hybrid fitness. Interestingly, there are more animal than plant studies that find evidence for adaptive introgression (see Table 1): for example in animals, mussels (Fraïsse et al., 2014), Drosophila (Garrigan et al., 2012; Brand et al., 2013; Llopart et al., 2014; Beck et al., 2015), Anopheles (Fontaine et al., 2015; Norris et al., 2015), Heliconius butterflies (Nadeau et al., 2014; Zhang et al., 2016), freshwater fish such as salmonids (Glover et al., 2013; Ozerov et al., 2016; Karlsson et al., 2016), birds such as Darwin’s finches (Lamichhaney et al., 2015), mice (Song et al., 2011; Staubach et al., 2012), and humans (Racimo et al., 2017); and in plants, Arabidopsis (Arnold et al., 2016), sunflowers (Whitney et al., 2015), Senecio (Kim et al., 2008), and poplars (Suarez-Gonzalez et al., 2016). This might be due to stronger economic/medical interest in animals, such as freshwater fish and malaria-transmitting mosquitos, resulting in more intense adaptive allele mining. However, this could also be due to higher rates of adaptive evolution in organisms with large effective population sizes (Gossmann et al., 2010), such as Drosophila, Anopheles, Heliconius butterflies, some freshwater fish, as well as mice. Introgression may provide fertile ground for adaptive radiations (Seehausen, 2004), either by enriching genetic variation in an initial hybridization event between two species that may then fuel radiation or by introducing adaptations that allow species of radiating lineages to occupy new niches and further diversify. Introgression has been shown to partly drive the adaptive radiations of some plants, such as Mimulus (Stankowski and Streisfeld, 2015) and Solanum (Pease et al., 2016), and some animals, such as Darwin’s finches (Lamichhaney et al., 2015) and cichlids (Meier et al., 2017). Numerous cases of adaptive introgression are also reported for fungi, such as yeast (Dunn et al., 2013; Almeida et al., 2017), as well as Albugo (McMullan et al., 2015).

HGT mediating pathogen resistance and environmental adaptation

HGT is also a source of novelty that is exploited in pathogen resistance. For example, a gene important for virus resistance in domestic tomato is derived from the fusion of two genes, both of which were acquired by HGT ultimately from bacteria, although one of the genes appears to have passed through a fungus (Yang et al., 2016). The obligatory parasite Phelipanche aegyptiaca acquired an albumin 1 gene, known to function as a storage protein and insect toxin, by HGT from legumes (Zhang et al., 2013). Structural predictions, the conservation of key functional residues, and evidence of purifying selection all suggest that the gene serves an adaptive role. This HGT event was likely the result of historical host-parasite interactions. Another example is the fungus Epichloë, an intracellular plant symbiont, which repeatedly was conferred insect resistance to its hosts by an insect toxin acquired by the fungus via HGT from bacteria (Ambrose et al., 2014).

The reverse is also true; HGT in plants, fungi, and insects has allowed enhanced pathogenic exploitation of plants. In the fungus responsible for apple canker, Valsa mali, multiple genes acquired by HGT from bacteria and fungi were implicated in pathogenicity with putative roles in the avoidance of host immune responses and the degradation of host tissues (Yin et al., 2016). There is evidence that the berry borer beetle Hypothenemus hampei adapted to its role as a pathogen of coffee beans by the acquisition of the HhMAN1 locus (Acuña et al., 2012). The enzyme encoded is capable of breaking down galactomannan, the main storage polysaccharide in coffee beans and likely allows the beetles to exploit coffee beans as a food source. HGT has allowed plant-parasitic root knot nematodes to acquire a repertoire of plant cell wall degradation enzymes that facilitate parasitism from bacteria (Danchin et al., 2010). Finally, a study of HGT in multiple parasitic plant lineages showed that genes acquired from the host via HGT are not only evolving under purifying or positive selection but are most likely to be expressed in the haustorium, the interface between host and parasite, implying they may have an adaptive role in host-parasite interactions (Yang et al., 2016).

HGT has also been implicated in mediating repeated adaptation to stringent environmental conditions. C4 photosynthesis is a more efficient photosynthetic pathway in hot arid conditions that has evolved multiple times from C3 progenitors. In the Alloteropsis grasses, two key C4 photosynthesis proteins, phosphoenolpyruvate carboxylase and phosphoenolpyruvate carboxykinase, have been acquired by HGT at least four times from distantly related plants in close ecological contact with the grasses (Christin et al., 2012). These pre-adapted C4 genes likely replaced their suboptimal homologues, rapidly optimizing the C4 pathway in the recipient. A metabolic enzyme with a key role in glucose metabolism that was acquired by plant-to-plant HGT has been shown to be associated with fine-scale biotic and abiotic environmental differences in the grass Festuca ovina (Prentice et al., 2015).

A particularly striking example of adaptation mediated by HGT involves the photoreceptor neochrome, a chimeric photoreceptor originating from a gene fusion that is thought to play a pivotal role in the enhanced phototropic response of ferns, a key adaptive trait that allowed their diversification following the advent of angiosperms (Kawai et al., 2003; Schneider et al., 2004; Kanegae et al., 2006; Schuettpelz and Pryer, 2009). This protein, previously thought to have arisen independently in ferns and algae, was initially acquired by the fern lineage through HGT from hornworts followed by subsequent inter-fern HGT (Li et al., 2014). As discussed above, both ferns and hornworts have lifestyle traits that likely make them more susceptible to HGT. Finally, there is evidence of HGT in commercially important species: domesticated sweet potato contains four transcribed Agrobacterium genes (Kyndt et al., 2015) and the silkworm Bombyx mori contains 10 expressed genes from bacterial sources, with putative functions in disease resistance and metabolism (Zhu et al., 2011).

Thus, HGT has had a profound impact on the genesis of many important adaptive plant traits such as the acquisition of endosymbionts, the origin of C4 photosynthesis, and the emergence of terrestrial plants. Compared with prokaryotes, HGT in eukaryotes is considered rare but the impact of such HGT events on the evolutionary trajectories of their recipients can be large. In this era of high-throughput sequencing technologies and especially whole genome sequencing, cases of HGT in multicellular eukaryotes are increasingly likely to be identified.

Near-term perspectives on adaptive allele mining using adaptive introgression

When phenotype-driven, allele mining for introgressed loci is a powerful tool for the identification of strong candidate alleles underlying particular adaptations. Indeed, we are at a watershed moment in the history of these approaches. There stands behind us a rich history, with long established model systems poised to be married to modern population genomics. It is now possible to detect loci under divergent selection, to test if the selected alleles have been introgressed and to associate these candidate alleles with phenotypes in ultra-high genomic resolution. In cases where these studies focus on clear phenotypes, it is obvious that such mixtures of approaches will engender rapid developments in understanding the mechanisms of adaptation (Box 1). Further, they stand to reveal the historic and geographic context of adaptive introgression mediating complex traits in the rich complexity presented by nature.

Abbreviations

     
  • GIsD

    genomic islands of divergence

  •  
  • HGT

    horizontal gene transfer

  •  
  • PaIS

    parapatrically isolated species

  •  
  • TEs

    transposable elements.

Acknowledgements

LY acknowledges funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement 679056) and BBSRC Institute Strategic Programme Grant (BB/J004588/1). RS acknowledges funding from the Czech Science Foundation (GAČR), grant number 16-15134Y. The authors thank Patrick Monnahan, Jordan Koch, and Pirita Paajanen for discussion.

References

Abbott
R
,
Albach
D
,
Ansell
S
et al. 
2013
.
Hybridization and speciation
.
Journal of Evolutionary Biology
26
,
229
246
.

Acuña
R
,
Padilla
BE
,
Flórez-Ramos
CP
et al. 
2012
.
Adaptive horizontal transfer of a bacterial gene to an invasive insect pest of coffee
.
Proceedings of the National Academy of Sciences, USA
109
,
4197
4202
.

Ai
H
,
Fang
X
,
Yang
B
et al. 
2015
.
Adaptation and possible ancient interspecies introgression in pigs identified by whole-genome sequencing
.
Nature Genetics
47
,
217
225
.

Alexander
DH
,
Novembre
J
,
Lange
K
.
2009
.
Fast model-based estimation of ancestry in unrelated individuals
.
Genome Research
19
,
1655
1664
.

Allendorf
FW
,
Leary
RF
,
Spruell
P
,
Wenburg
JK
.
2001
.
The problems with hybrids: setting conservation guidelines
.
Trends in Ecology and Evolution
16
,
613
622
.

Almeida
P
,
Barbosa
R
,
Bensasson
D
,
Gonçalves
P
,
Sampaio
JP
.
2017
.
Adaptive divergence in wine yeasts and their wild relatives suggests a prominent role for introgressions and rapid evolution at noncoding sites
.
Molecular Ecology
26
,
2167
2182
.

Ambrose
KV
,
Koppenhöfer
AM
,
Belanger
FC
.
2014
.
Horizontal gene transfer of a bacterial insect toxin gene into the Epichloë fungal symbionts of grasses
.
Scientific Reports
4
,
5562
.

Anderson
E
.
1949
.
Introgressive hybridization
.
New York
:
John Wiley and Sons
.

Anderson
E
,
Hubricht
L
.
1938
.
Hybridization in Tradescantia. III. The evidence for introgressive hybridization
.
American Journal of Botany
25
,
396
402
.

Anderson
TM
,
vonHoldt
BM
,
Candille
SI
et al. 
2009
.
Molecular and evolutionary history of melanism in North American gray wolves
.
Science
323
,
1339
1343
.

Andersson
JO
.
2009
.
Gene transfer and diversification of microbial eukaryotes
.
Annual Review of Microbiology
63
,
177
193
.

Arnold
ML
.
2016
.
Divergence with genetic exchange
.
Oxford
:
Oxford University Press
.

Arnold
ML
,
Martin
NH
.
2009
.
Adaptation by introgression
.
Journal of Biology
8
,
82
.

Arnold
BJ
,
Lahner
B
,
DaCosta
JM
,
Weisman
CM
,
Hollister
JD
,
Salt
DE
,
Bomblies
K
,
Yant
L
.
2016
.
Borrowed alleles and convergence in serpentine adaptation
.
Proceedings of the National Academy of Sciences, USA
113
,
8320
8325
.

Barton
NH
.
1979
.
The dynamics of hybrid zones
.
Heredity
43
,
341
359
.

Barton
NH
.
2001
.
The role of hybridization in evolution
.
Molecular Ecology
10
,
551
568
.

Barton
NH
,
Hewitt
GM
.
1985
.
Analysis of hybrid zones
.
Annual Review of Ecology, Evolution, and Systematics
16
,
113
148
.

Barton
NH
,
Hewitt
GM
.
1989
.
Adaptation, speciation and hybrid zones
.
Nature
341
,
497
503
.

Bastide
H
,
Lange
JD
,
Lack
JB
,
Yassin
A
,
Pool
JE
.
2016
.
A variable genetic architecture of melanic evolution in Drosophila melanogaster
.
Genetics
204
,
1307
1319
.

Beck
EA
,
Thompson
AC
,
Sharbrough
J
,
Brud
E
,
Llopart
A
.
2015
.
Gene flow between Drosophila yakuba and Drosophila santomea in subunit V of cytochrome c oxidase: a potential case of cytonuclear cointrogression
.
Evolution
69
,
1973
1986
.

Bergthorsson
U
,
Adams
KL
,
Thomason
B
,
Palmer
JD
.
2003
.
Widespread horizontal transfer of mitochondrial genes in flowering plants
.
Nature
424
,
197
201
.

Blischak
PD
,
Kubatko
LS
,
Wolfe
AD
.
2017
.
SNP genotyping and parameter estimation in polyploids using low-coverage sequencing data
. bioRxiv
doi: 10.1101/120261
.

Bock
R
.
2010
.
The give-and-take of DNA: horizontal gene transfer in plants
.
Trends in Plant Science
15
,
11
22
.

Bradburd
GS
,
Ralph
PL
,
Coop
GM
.
2016
.
A spatial framework for understanding population structure and admixture
.
PLOS Genetics
12
,
e1005703
.

Brand
CL
,
Kingan
SB
,
Wu
L
,
Garrigan
D
.
2013
.
A selective sweep across species boundaries in Drosophila
.
Molecular Biology and Evolution
30
,
2177
2186
.

Brennan
AC
,
Bridle
JR
,
Wang
AL
,
Hiscock
SJ
,
Abbott
RJ
.
2009
.
Adaptation and selection in the Senecio (Asteraceae) hybrid zone on Mount Etna, Sicily
.
New Phytologist
183
,
702
717
.

Brower
AVZ
.
2013
.
Introgression of wing pattern alleles and speciation via homoploid hybridization in Heliconius butterflies: a review of evidence from the genome
.
Proceedings of the Royal Society B
280
,
20122302
.

Buerkle
CA
,
Lexer
C
.
2008
.
Admixture as the basis for genetic mapping
.
Trends in Ecology & Evolution
23
,
686
694
.

Burri
R
,
Nater
A
,
Kawakami
T
et al. 
2015
.
Linked selection and recombination rate variation drive the evolution of the genomic landscape of differentiation across the speciation continuum of Ficedula flycatchers
.
Genome Research
25
,
1656
1665
.

Chapman
MA
,
Abbott
RJ
.
2010
.
Introgression of fitness genes across a ploidy barrier
.
New Phytologist
186
,
63
71
.

Cho
Y
,
Qiu
YL
,
Kuhlman
P
,
Palmer
JD
.
1998
.
Explosive invasion of plant mitochondria by a group I intron
.
Proceedings of the National Academy of Sciences, USA
95
,
14244
14249
.

Christin
PA
,
Edwards
EJ
,
Besnard
G
,
Boxall
SF
,
Gregory
R
,
Kellogg
EA
,
Hartwell
J
,
Osborne
CP
.
2012
.
Adaptive evolution of C(4) photosynthesis through recurrent lateral gene transfer
.
Current Biology
22
,
445
449
.

Clark
LV
,
Stewart
JR
,
Nishiwaki
A
et al. 
2015
.
Genetic structure of Miscanthus sinensis and Miscanthus sacchariflorus in Japan indicates a gradient of bidirectional but asymmetric introgression
.
Journal of Experimental Botany
66
,
4213
4225
.

Clarkson
CS
,
Weetman
D
,
Essandoh
J
et al. 
2014
.
Adaptive introgression between Anopheles sibling species eliminates a major genomic island but not reproductive isolation
.
Nature Communications
5
,
4248
.

Coleman
ML
,
Sullivan
MB
,
Martiny
AC
,
Steglich
C
,
Barry
K
,
Delong
EF
,
Chisholm
SW
.
2006
.
Genomic islands and the ecology and evolution of Prochlorococcus
.
Science
311
,
1768
1770
.

Coyne
JA
,
Orr
HA
.
2004
.
Speciation
.
Sunderland, Massachusetts
:
Sinauer Associates
.

Crawford
JE
,
Nielsen
R
.
2013
.
Detecting adaptive trait loci in nonmodel systems: divergence or admixture mapping?
Molecular Ecology
22
,
6131
6148
.

Crispo
E
,
Moore
J-S
,
Lee-Yaw
JA
,
Gray
SM
,
Haller
BC
.
2011
.
Broken barriers: human-induced changes to gene flow and introgression in animals
.
BioEssays
33
,
508
518
.

Cruickshank
TE
,
Hahn
MW
.
2014
.
Reanalysis suggests that genomic islands of speciation are due to reduced diversity, not reduced gene flow
.
Molecular Ecology
23
,
3133
3157
.

Cucchi
T
,
Vigne
J-D
,
Auffray
J-C
.
2005
.
First occurence of the house mouse (Mus musculus domesticus Schwarz & Schwarz, 1943) in the Western Mediterranean: a zooarchaeological revision of subfossil occurrences
.
Biological Journal of the Linnean Society
84
,
429
445
.

Danchin
EGJ
,
Rosso
M-N
,
Vieira
P
,
de Almeida-Engler
J
,
Coutinho
PM
,
Henrissat
B
,
Abad
P
.
2010
.
Multiple lateral gene transfers and duplications have promoted plant parasitism ability in nematodes
.
Proceedings of the National Academy of Sciences, USA
107
,
17651
17656
.

Davis
CC
,
Xi
Z
.
2015
.
Horizontal gene transfer in parasitic plants
.
Current Opinion in Plant Biology
26
,
14
19
.

Délye
C
,
Jasieniuk
M
,
Le Corre
V
.
2013
.
Deciphering the evolution of herbicide resistance in weeds
.
Trends in Genetics
29
,
649
658
.

den Nijs
HCM
,
Bartsch
D
,
Sweet
J
(eds.).
2004
.
Introgression from genetically modified plants into wild relatives
.
Cambridge, Massachusetts
:
CABI Publishing
.

Dobzhansky
T
.
1937
.
Genetics and the origin of species
.
New York
:
Columbia University Press
.

Dunn
B
,
Paulish
T
,
Stanbery
A
et al. 
2013
.
Recurrent rearrangement during adaptive evolution in an interspecific yeast hybrid suggests a model for rapid introgression
.
PLOS Genetics
9
,
e1003366
.

Durand
EY
,
Patterson
N
,
Reich
D
,
Slatkin
M
.
2011
.
Testing for ancient admixture between closely related populations
.
Molecular Biology and Evolution
28
,
2239
2252
.

Edmands
S
.
2007
.
Between a rock and a hard place: evaluating the relative risks of inbreeding and outbreeding for conservation and management
.
Molecular Ecology
16
,
463
475
.

El Baidouri
M
,
Carpentier
MC
,
Cooke
R
et al. 
2014
.
Widespread and frequent horizontal transfers of transposable elements in plants
.
Genome Research
24
,
831
838
.

Emiliani
G
,
Fondi
M
,
Fani
R
,
Gribaldo
S
.
2009
.
A horizontal gene transfer at the origin of phenylpropanoid metabolism: a key adaptation of plants to land
.
Biology Direct
4
,
7
.

Endler
JA
.
1977
.
Geographic variation, speciation and clines
.
Monographs in Population Biology
, No. 10,
Princeton, New Jersey
:
Princeton University Press
.

Excoffier
L
,
Dupanloup
I
,
Huerta-Sánchez
E
,
Sousa
VC
,
Foll
M
.
2013
.
Robust demographic inference from genomic and SNP data
.
PLOS Genetics
9
,
e1003905
.

Feder
JL
,
Egan
SP
,
Nosil
P
.
2012
.
The genomics of speciation-with-gene-flow
.
Trends in Genetics
28
,
342
350
.

Feder
JL
,
Nosil
P
.
2009
.
Chromosomal inversions and species differences: when are genes affecting adaptive divergence and reproductive isolation expected to reside within inversions?
Evolution
63
,
3061
3075
.

Filatov
DA
,
Osborne
OG
,
Papadopulos
AS
.
2016
.
Demographic history of speciation in a Senecio altitudinal hybrid zone on Mt. Etna
.
Molecular Ecology
25
,
2467
2481
.

Fitzpatrick
BM
,
Johnson
JR
,
Kump
DK
,
Smith
JJ
,
Voss
SR
,
Shaffer
HB
.
2010
.
Rapid spread of invasive genes into a threatened native species
.
Proceedings of the National Academy of Sciences, USA
107
,
3606
3610
.

Fontaine
MC
,
Pease
JB
,
Steele
A
et al. 
2015
.
Extensive introgression in a malaria vector species complex revealed by phylogenomics
.
Science
347
,
1258524
.

Fontdevila
A
.
2005
.
Hybrid genome evolution by transposition
.
Cytogenetic and Genome Research
110
,
49
55
.

Fraïsse
C
,
Belkhir
K
,
Welch
JJ
,
Bierne
N
.
2016
.
Local interspecies introgression is the main cause of extreme levels of intraspecific differentiation in mussels
.
Molecular Ecology
25
,
269
286
.

Fraïsse
C
,
Roux
C
,
Welch
JJ
,
Bierne
N
.
2014
.
Gene-flow in a mosaic hybrid zone: is local introgression adaptive?
Genetics
197
,
939
951
.

Garrigan
D
,
Kingan
SB
,
Geneva
AJ
,
Andolfatto
P
,
Clark
AG
,
Thornton
KR
,
Presgraves
DC
.
2012
.
Genome sequencing reveals complex speciation in the Drosophila simulans clade
.
Genome Research
22
,
1499
1511
.

Gillings
MR
.
2017
.
Lateral gene transfer, bacterial genome evolution, and the Anthropocene
.
Annals of the New York Academy of Sciences
1389
,
20
36
.

Glover
KA
,
Pertoldi
C
,
Besnier
F
,
Wennevik
V
,
Kent
M
,
Skaala
Ø
.
2013
.
Atlantic salmon populations invaded by farmed escapees: quantifying genetic introgression with a Bayesian approach and SNPs
.
BMC Genetics
14
,
74
.

Gómez
JM
,
González-Megías
A
,
Lorite
J
,
Abdelaziz
M
,
Perfectti
F
.
2015
.
The silent extinction: climate change and the potential hybridization-mediated extinction of endemic high-mountain plants
.
Biodiversity and Conservation
24
,
1843
1857
.

Gompert
Z
,
Buerkle
CA
.
2016
.
What, if anything, are hybrids: enduring truths and challenges associated with population structure and gene flow
.
Evolutionary Applications
9
,
909
923
.

Gossmann
TI
,
Song
BH
,
Windsor
AJ
,
Mitchell-Olds
T
,
Dixon
CJ
,
Kapralov
MV
,
Filatov
DA
,
Eyre-Walker
A
.
2010
.
Genome wide analyses reveal little evidence for adaptive evolution in many plant species
.
Molecular Biology and Evolution
27
,
1822
1832
.

Green
RE
,
Krause
J
,
Briggs
AW
et al. 
2010
.
A draft sequence of the Neandertal genome
.
Science
328
,
710
722
.

Guénet
JL
,
Bonhomme
F
.
2003
.
Wild mice: an ever-increasing contribution to a popular mammalian model
.
Trends in Genetics
19
,
24
31
.

Gutenkunst
RN
,
Hernandez
RD
,
Williamson
SH
,
Bustamante
CD
.
2009
.
Inferring the joint demographic history of multiple populations from multidimensional SNP frequency data
.
PLOS Genetics
5
,
e1000695
.

Hamilton
JA
,
De la Torre
AR
,
Aitken
SN
.
2015
.
Fine-scale environmental variation contributes to introgression in a three-species spruce hybrid complex
.
Tree Genetics and Genomes
11
,
817
.

Hamilton
JA
,
Lexer
C
,
Aitken
SN
.
2013
.
Differential introgression reveals candidate genes for selection across a spruce (Picea sitchensis × P. glauca) hybrid zone
.
New Phytologist
197
,
927
938
.

Han
TS
,
Wu
Q
,
Hou
XH
,
Li
ZW
,
Zou
YP
,
Ge
S
,
Guo
YL
.
2015
.
Frequent introgressions from diploid species contribute to the adaptation of the tetraploid Shepherd’s purse (Capsella bursa-pastoris)
.
Molecular Plant
8
,
427
438
.

Hand
BK
,
Lowe
WH
,
Kovach
RP
,
Muhlfeld
CC
,
Luikart
G
.
2015
.
Landscape community genomics: understanding eco-evolutionary processes in complex environments
.
Trends in Ecology & Evolution
30
,
161
168
.

Harr
B
.
2006
.
Genomic islands of differentiation between house mouse subspecies
.
Genome Research
16
,
730
737
.

Harrison
RG
.
1986
.
Pattern and process in a narrow hybrid zone
.
Heredity
56
,
337
349
.

Harrison
RG
.
1990
.
Hybrid zones: windows on evolutionary process
. In:
Futuyma
D
,
Antonovics
J
, eds.,
Oxford surveys in evolutionary biology
,
Oxford
:
Oxford University Press
,
69
128
.

Harrison
RG
.
2012
.
The language of speciation
.
Evolution
66
,
3643
3657
.

Harrison
RG
,
Larson
EL
.
2014
.
Hybridization, introgression, and the nature of species boundaries
.
Journal of Heredity
105
,
795
809
.

Hedrick
PW
.
2013
.
Adaptive introgression in animals: examples and comparison to new mutation and standing variation as sources of adaptive variation
.
Molecular Ecology
22
,
4606
4618
.

Hewitt
GM
.
1975
.
A sex-chromosome hybrid zone in the grasshopper Podisma pedestris (Orthoptera: Acrididae)
.
Heredity
35
,
375
387
.

Hewitt
GM
.
1988
.
Hybrid zones-natural laboratories for evolutionary studies
.
Trends in ecology & evolution
3
,
158
167
.

Hey
J
.
2010a
.
Isolation with migration models for more than two populations
.
Molecular Biology and Evolution
27
,
905
920
.

Hey
J
.
2010b
.
The divergence of chimpanzee species and subspecies as revealed in multipopulation isolation-with-migration analyses
.
Molecular Biology and Evolution
27
,
921
933
.

Hey
J
,
Nielsen
R
.
2007
.
Integration within the Felsenstein equation for improved Markov chain Monte Carlo methods in population genetics
.
Proceedings of the National Academy of Sciences, USA
104
,
2785
2790
.

Hohenlohe
PA
,
Bassham
S
,
Currey
M
,
Cresko
WA
.
2012
.
Extensive linkage disequilibrium and parallel adaptive divergence across threespine stickleback genomes
.
Philosophical Transactions of the Royal Society B
367
,
395
408
.

Hohenlohe
PA
,
Day
MD
,
Amish
SJ
et al. 
2013
.
Genomic patterns of introgression in rainbow and westslope cutthroat trout illuminated by overlapping paired-end RAD sequencing
.
Molecular Ecology
22
,
3002
3013
.

Hohmann
N
,
Wolf
EM
,
Lysak
MA
,
Koch
MA
.
2015
.
A time-calibrated road map of brassicaceae species radiation and evolutionary history
.
The Plant Cell
27
,
2770
2784
.

Huang
J
,
Gogarten
JP
.
2008
.
Concerted gene recruitment in early plant evolution
.
Genome Biology
9
,
R109
.

Hubisz
MJ
,
Falush
D
,
Stephens
M
,
Pritchard
JK
.
2009
.
Inferring weak population structure with the assistance of sample group information
.
Molecular Ecology Resources
9
,
1322
1332
.

Huerta-Sánchez
E
,
Jin
X
,
Asan
et al. 
2014
.
Altitude adaptation in Tibetans caused by introgression of Denisovan-like DNA
.
Nature
512
,
194
197
.

Hufford
MB
,
Lubinksy
P
,
Pyhäjärvi
T
,
Devengenzo
MT
,
Ellstrand
NC
,
Ross-Ibarra
J
.
2013
.
The genomic signature of crop-wild introgression in maize
.
PLOS Genetics
9
,
e1003477
.

Hufford
MB
,
Xu
X
,
van Heerwaarden
J
et al. 
2012
.
Comparative population genomics of maize domestication and improvement
.
Nature Genetics
44
,
808
811
.

Huson
DH
,
Bryant
D
.
2006
.
Application of phylogenetic networks in evolutionary studies
.
Molecular Biology and Evolution
23
,
254
267
.

Jensen
JD
,
Foll
M
,
Bernatchez
L
.
2016
.
Special issue: detecting selection in natural populations
.
Molecular Ecology
25
,
1
430
.

Jeong
C
,
Alkorta-Aranburu
G
,
Basnyat
B
,
Neupane
M
,
Witonsky
DB
,
Pritchard
JK
,
Beall
CM
,
Di Rienzo
A
.
2014
.
Admixture facilitates genetic adaptations to high altitude in Tibet
.
Nature Communications
5
,
3281
.

Jombart
T
.
2008
.
adegenet: a R package for the multivariate analysis of genetic markers
.
Bioinformatics
24
,
1403
1405
.

Jombart
T
,
Devillard
S
,
Balloux
F
.
2010
.
Discriminant analysis of principal components: a new method for the analysis of genetically structured populations
.
BMC Genetics
11
,
94
.

Jónsson
H
,
Schubert
M
,
Seguin-Orlando
A
et al. 
2014
.
Speciation with gene flow in equids depite extensive chromosomal plasticity
.
Proceedings of the National Academy of Sciences, USA
111
,
18655
18660
.

Kanegae
T
,
Hayashida
E
,
Kuramoto
C
,
Wada
M
.
2006
.
A single chromoprotein with triple chromophores acts as both a phytochrome and a phototropin
.
Proceedings of the National Academy of Sciences, USA
103
,
17997
18001
.

Karasov
T
,
Messer
PW
,
Petrov
DA
.
2010
.
Evidence that adaptation in Drosophila is not limited by mutation at single sites
.
PLOS Genetics
6
,
e1000924
.

Karlsson
S
,
Diserud
OH
,
Fiske
P
,
Hindar
K
.
2016
.
Widespread genetic introgression of escaped farmed Atlantic salmon in wild salmon populations
.
ICES Journal of Marine Science
73
,
2488
2498
.

Kawai
H
,
Kanegae
T
,
Christensen
S
,
Kiyosue
T
,
Sato
Y
,
Imaizumi
T
,
Kadota
A
,
Wada
M
.
2003
.
Responses of ferns to red light are mediated by an unconventional photoreceptor
.
Nature
421
,
287
290
.

Keeling
PJ
,
Palmer
JD
.
2008
.
Horizontal gene transfer in eukaryotic evolution
.
Nature reviews. Genetics
9
,
605
618
.

Kellis
M
,
Patterson
N
,
Endrizzi
M
,
Birren
B
,
Lander
ES
.
2003
.
Sequencing and comparison of yeast species to identify genes and regulatory elements
.
Nature
423
,
241
254
.

Kim
M
,
Cui
M-L
,
Cubas
P
,
Gillies
A
,
Lee
K
,
Chapman
MA
,
Abbott
RJ
,
Coen
E
.
2008
.
Regulatory genes control a key morphological and ecological trait transferred between species
.
Science
322
,
1116
1119
.

Kim
G
,
LeBlanc
ML
,
Wafula
EK
,
dePamphilis
CW
,
Westwood
JH
.
2014
.
Genomic-scale exchange of mRNA between a parasitic plant and its hosts
.
Science
345
,
808
811
.

Klasson
L
,
Kambris
Z
,
Cook
PE
,
Walker
T
,
Sinkins
SP
.
2009
.
Horizontal gene transfer between Wolbachia and the mosquito Aedes aegypti
.
BMC Genomics
10
,
33
.

Kovach
RP
,
Hand
BK
,
Hohenlohe
PA
et al. 
2016
.
Vive la résistance: genome-wide selection against introduced alleles in invasive hybrid zones
.
Proceedings of the Royal Society B
283
,
20161380
.

Kovach
RP
,
Muhlfeld
CC
,
Boyer
MC
,
Lowe
WH
,
Allendorf
FW
,
Luikart
G
.
2015
.
Dispersal and selection mediate hybridization between a native and invasive species
.
Proceedings of the Royal Society B
282
,
20142454
.

Kozak
KM
,
Wahlberg
N
,
Neild
AF
,
Dasmahapatra
KK
,
Mallet
J
,
Jiggins
CD
.
2015
.
Multilocus species trees show the recent adaptive radiation of the mimetic heliconius butterflies
.
Systematic Biology
64
,
505
524
.

Kronforst
MR
,
Hansen
ME
,
Crawford
NG
,
Gallant
JR
,
Zhang
W
,
Kulathinal
RJ
,
Kapan
DD
,
Mullen
SP
.
2013
.
Hybridization reveals the evolving genomic architecture of speciation
.
Cell Reports
5
,
666
677
.

Kumar
V
,
Lammers
F
,
Bidon
T
,
Pfenninger
M
,
Kolter
L
,
Nilsson
MA
,
Janke
A
.
2017
.
The evolutionary history of bears is characterized by gene flow across species
.
Scientific Reports
7
,
46487
.

Kyndt
T
,
Quispe
D
,
Zhai
H
,
Jarret
R
,
Ghislain
M
,
Liu
Q
,
Gheysen
G
,
Kreuze
JF
.
2015
.
The genome of cultivated sweet potato contains Agrobacterium T-DNAs with expressed genes: An example of a naturally transgenic food crop
.
Proceedings of the National Academy of Sciences, USA
112
,
5844
5849
.

Lacroix
B
,
Citovsky
V
.
2016
.
Transfer of DNA from bacteria to eukaryotes
.
mBio
7
,
1
9
.

Lafon-Placette
C
,
Johannessen
IM
,
Hornslien
KS
et al. 
2017
.
Endosperm-based hybridization barriers explain the pattern of gene flow between Arabidopsis lyrata and Arabidopsis arenosa in Central Europe
.
Proceedings of the National Academy of Sciences, USA
114
,
E1027
E1035
.

Lamichhaney
S
,
Berglund
J
,
Almén
MS
et al. 
2015
.
Evolution of Darwin’s finches and their beaks revealed by genome sequencing
.
Nature
518
,
371
375
.

Laval
G
,
Excoffier
L
.
2004
.
SIMCOAL 2.0: a program to simulate genomic diversity over large recombining regions in a subdivided population with a complex history
.
Bioinformatics
20
,
2485
2487
.

Lawson
DJ
,
Hellenthal
G
,
Myers
S
,
Falush
D
.
2012
.
Inference of population structure using dense haplotype data
.
PLOS Genetics
8
,
e1002453
.

Lee
Y
,
Collier
TC
,
Sanford
MR
,
Marsden
CD
,
Fofana
A
,
Cornel
AJ
,
Lanzaro
GC
.
2013
.
Chromosome inversions, genomic differentiation and speciation in the African malaria mosquito Anopheles gambiae
.
PLOS One
8
,
e57887
.

Lee
Y
,
Marsden
CD
,
Norris
LC
,
Collier
TC
,
Main
BJ
,
Fofana
A
,
Cornel
AJ
,
Lanzaro
GC
.
2013b
.
Spatiotemporal dynamics of gene flow and hybrid fitness between the M and S forms of the malaria mosquito, Anopheles gambiae
.
Proceedings of the National Academy of Sciences, USA
110
,
19854
19859
.

Lee
CR
,
Svardal
H
,
Farlow
A
,
Exposito-Alonso
M
,
Ding
W
,
Novikova
P
,
Alonso-Blanco
C
,
Weigel
D
,
Nordborg
M
.
2017
.
On the post-glacial spread of human commensal Arabidopsis thaliana
.
Nature Communications
8
,
14458
.

Levsen
ND
,
Tiffin
P
,
Olson
MS
.
2012
.
Pleistocene speciation in the genus Populus (salicaceae)
.
Systematic Biology
61
,
401
412
.

Li
YF
,
Costello
JC
,
Holloway
AK
,
Hahn
MW
.
2008
.
“Reverse ecology” and the power of population genomics
.
Evolution
62
,
2984
2994
.

Li
H
,
Durbin
R
.
2011
.
Inference of human population history from whole genome sequence of a single individual
.
Nature
475
,
493
496
.

Li
F-W
,
Villarreal
JC
,
Kelly
S
et al. 
2014
.
Horizontal transfer of an adaptive chimeric photoreceptor from bryophytes to ferns
.
Proceedings of the National Academy of Sciences, USA
111
,
6672
6677
.

Liu
S
,
Li
F
,
Kong
L
,
Sun
Y
,
Qin
L
,
Chen
S
,
Cui
H
,
Huang
Y
,
Xia
G
.
2015
.
Genetic and epigenetic changes in somatic hybrid introgression lines between wheat and tall wheatgrass
.
Genetics
199
,
1035
1045
.

Liu
Z
,
Wang
Y
,
Shen
Y
,
Guo
W
,
Hao
S
,
Liu
B
.
2004
.
Extensive alterations in DNA methylation and transcription in rice caused by introgression from Zizania latifolia
.
Plant Molecular Biology
54
,
571
582
.

Llopart
A
,
Elwyn
S
,
Lachaise
D
,
Coyne
JA
.
2002
.
Genetics of a difference in pigmentation between Drosophila yakuba and Drosophila santomea
.
Evolution
56
,
2262
2277
.

Llopart
A
,
Herrig
D
,
Brud
E
,
Stecklein
Z
.
2014
.
Sequential adaptive introgression of the mitochondrial genome in Drosophila yakuba and Drosophila santomea
.
Molecular Ecology
23
,
1124
1136
.

Long
A
,
Liti
G
,
Luptak
A
,
Tenaillon
O
.
2015
.
Elucidating the molecular architecture of adaptation via evolve and resequence experiments
.
Nature reviews. Genetics
16
,
567
582
.

Lotterhos
KE
,
Whitlock
MC
.
2015
.
The relative power of genome scans to detect local adaptation depends on sampling design and statistical method
.
Molecular Ecology
24
,
1031
1046
.

Mahelka
V
,
Krak
K
,
Kopecký
D
,
Fehrer
J
,
Šafář
J
,
Bartoš
J
,
Hobza
R
,
Blavet
N
,
Blattner
FR
.
2017
.
Multiple horizontal transfers of nuclear ribosomal genes between phylogenetically distinct grass lineages
.
Proceedings of the National Academy of Sciences,USA
114
,
1726
1731
.

Malinsky
M
,
Challis
RJ
,
Tyers
AM
et al. 
2015
.
Genomic islands of speciation separate cichlid ecomorphs in an East African crater lake
.
Science
350
,
1493
1498
.

Mallet
J
.
2005
.
Hybridization as an invasion of the genome
.
Trends in ecology & evolution
20
,
229
237
.

Mallet
J
.
2007
.
Hybrid speciation
.
Nature
446
,
279
283
.

Mallet
J
,
Barton
N
,
Lamas
G
,
Santisteban
J
,
Muedas
M
,
Eeley
H
.
1990
.
Estimates of selection and gene flow from measures of cline width and linkage disequilibrium in heliconius hybrid zones
.
Genetics
124
,
921
936
.

Mallet
J
,
Besansky
N
,
Hahn
MW
.
2016
.
How reticulated are species?
BioEssays
38
,
140
149
.

Marques
DA
,
Lucek
K
,
Meier
JI
,
Mwaiko
S
,
Wagner
CE
,
Excoffier
L
,
Seehausen
O
.
2016
.
Genomics of Rapid Incipient Speciation in Sympatric Threespine Stickleback
.
PLOS Genetics
12
,
e1005887
.

Martin
SH
,
Dasmahapatra
KK
,
Nadeau
NJ
et al. 
2013
.
Genome-wide evidence for speciation with gene flow in Heliconius butterflies
.
Genome Research
23
,
1817
1828
.

Martin
SH
,
Davey
JW
,
Jiggins
CD
.
2015
.
Evaluating the use of ABBA-BABA statistics to locate introgressed loci
.
Molecular Biology and Evolution
32
,
244
257
.

Martin
SH
,
Van Belleghem
SM
.
2017
.
Exploring evolutionary relationships across the genome using topology weighting
.
Genetics
206
,
429
438
.

Mayr
E
.
1942
.
Systematics and the origin of species
.
New York
:
Columbia University Press
.

McMullan
M
,
Gardiner
A
,
Bailey
K
et al. 
2015
.
Evidence for suppression of immunity as a driver for genomic introgressions and host range expansion in races of Albugo candida, a generalist parasite
.
eLife
4
:
e04550
.

Meier
JI
,
Marques
DA
,
Mwaiko
S
,
Wagner
CE
,
Excoffier
L
,
Seehausen
O
.
2017
.
Ancient hybridization fuels rapid cichlid fish adaptive radiations
.
Nature Communications
8
,
14363
.

Meyer
M
,
Kircher
M
,
Gansauge
M-T
et al. 
2012
.
A high coverage genome sequence from an archaic Denisovan individual
.
Science
338
,
222
226
.

Michel
AP
,
Sim
S
,
Powell
TH
,
Taylor
MS
,
Nosil
P
,
Feder
JL
.
2010
.
Widespread genomic divergence during sympatric speciation
.
Proceedings of the National Academy of Sciences, USA
107
,
9724
9729
.

Moore
WS
.
1977
.
An evaluation of narrow hybrid zones in vertebrates
.
The Quarterly Review of Biology
52
,
263
277
.

Morales
AE
,
Jackson
ND
,
Dewey
TA
,
O’Meara
BC
,
Carstens
BC
.
2017
.
Speciation with gene flow in north american Myotis bats
.
Systematic Biology
66
,
440
452
.

Mower
JP
,
Stefanović
S
,
Hao
W
,
Gummow
JS
,
Jain
K
,
Ahmed
D
,
Palmer
JD
.
2010
.
Horizontal acquisition of multiple mitochondrial genes from a parasitic plant followed by gene conversion with host mitochondrial genes
.
BMC Biology
8
,
150
.

Muhlfeld
CC
,
Kovach
RP
,
Jones
LA
,
Al-Chokhachy
R
,
Boyer
MC
,
Leary
RF
,
Lowe
WH
,
Luikart
G
,
Allendorf
FW
.
2014
.
Invasive hybridization in a threatened species is accelerated by climate change
.
Nature Climate Change
4
,
620
624
.

Nadeau
NJ
,
Martin
SH
,
Kozak
KM
et al. 
2013
.
Genome-wide patterns of divergence and gene flow across a butterfly radiation
.
Molecular Ecology
22
,
814
826
.

Nadeau
NJ
,
Ruiz
M
,
Salazar
P
et al. 
2014
.
Population genomics of parallel hybrid zones in the mimetic butterflies, H. melpomene and H. erato
.
Genome Research
24
,
1316
1333
.

Nielsen
R
,
Williamson
S
,
Kim
Y
,
Hubisz
MJ
,
Clark
AG
,
Bustamante
C
.
2005
.
Genomic scans for selective sweeps using SNP data
.
Genome Research
15
,
1566
1575
.

Noor
MA
,
Bennett
SM
.
2009
.
Islands of speciation or mirages in the desert? Examining the role of restricted recombination in maintaining species
.
Heredity
103
,
439
444
.

Noor
MAF
,
Grams
KL
,
Bertucci
LA
,
Reiland
J
.
2001
.
Chromosomal inversions and the reproductive isolation of species
.
Proceedings of the National Academy of Sciences, USA
98
,
12084
12088
.

Norris
LC
,
Main
BJ
,
Lee
Y
,
Collier
TC
,
Fofana
A
,
Cornel
AJ
,
Lanzaro
GC
.
2015
.
Adaptive introgression in an African malaria mosquito coincident with the increased usage of insecticide-treated bed nets
.
Proceedings of the National Academy of Sciences, USA
112
,
815
820
.

Nosil
P
.
2008
.
Speciation with gene flow could be common
.
Molecular Ecology
17
,
2103
2106
.

Nosil
P
,
Schluter
D
.
2011
.
The genes underlying the process of speciation
.
Trends in Ecology & Evolution
26
,
160
167
.

Olson-Manning
CF
,
Wagner
MR
,
Mitchell-Olds
T
.
2012
.
Adaptive evolution: evaluating empirical support for theoretical predictions
.
Nature reviews. Genetics
13
,
867
877
.

Orr
HA
,
Masly
JP
,
Presgraves
DC
.
2004
.
Speciation genes
.
Current Opinion in Genetics & Development
14
,
675
679
.

Ottoni
C
,
Van Neer
W
,
De Cupere
B
et al. 
2017
.
The palaeogenetics of cat dispersal in the ancient world
.
Nature Ecology and Evolution
1
,
0139
.

Ozerov
MY
,
Gross
R
,
Bruneaux
M
,
Vähä
JP
,
Burimski
O
,
Pukk
L
,
Vasemägi
A
.
2016
.
Genomewide introgressive hybridization patterns in wild Atlantic salmon influenced by inadvertent gene flow from hatchery releases
.
Molecular Ecology
25
,
1275
1293
.

Paaby
AB
,
Rockman
MV
.
2014
.
Cryptic genetic variation: evolution’s hidden substrate
.
Nature reviews. Genetics
15
,
247
258
.

Pallares
LF
,
Harr
B
,
Turner
LM
,
Tautz
D
.
2014
.
Use of a natural hybrid zone for genomewide association mapping of craniofacial traits in the house mouse
.
Molecular Ecology
23
,
5756
5770
.

Pallares
LF
,
Turner
LM
,
Tautz
D
.
2016
.
Craniofacial shape transition across the house mouse hybrid zone: implications for the genetic architecture and evolution of between-species differences
.
Development Genes and Evolution
226
,
173
186
.

Pardo-Diaz
C
,
Salazar
C
,
Baxter
SW
,
Merot
C
,
Figueiredo-Ready
W
,
Joron
M
,
McMillan
WO
,
Jiggins
CD
.
2012
.
Adaptive introgression across species boundaries in Heliconius butterflies
.
PLOS Genetics
8
,
e1002752
.

Parisod
C
,
Definod
C
,
Sarr
A
,
Arrigo
N
,
Felber
F
.
2013
.
Genome-specific introgression between wheat and its wild relative Aegilops triuncialis
.
Journal of Evolutionary Biology
26
,
223
228
.

Park
JM
,
Manen
JF
,
Schneeweiss
GM
.
2007
.
Horizontal gene transfer of a plastid gene in the non-photosynthetic flowering plants Orobanche and Phelipanche (Orobanchaceae)
.
Molecular Phylogenetics and Evolution
43
,
974
985
.

Pease
JB
,
Haak
DC
,
Hahn
MW
,
Moyle
LC
.
2016
.
Phylogenomics Reveals Three Sources of Adaptive Variation during a Rapid Radiation
.
PLOS Biology
14
,
e1002379
.

Pease
JB
,
Hahn
MW
.
2015
.
Detection and polarization of introgression in a five-taxon phylogeny
.
Systematic Biology
64
,
651
662
.

Pickrell
JK
,
Pritchard
JK
.
2012
.
Inference of population splits and mixtures from genome-wide allele frequency data
.
PLOS Genetics
8
,
e1002967
.

Pinheiro
F
,
De Barros
F
,
Palma-Silva
C
,
Meyer
D
,
Fay
MF
,
Suzuki
RM
,
Lexer
C
,
Cozzolino
S
.
2010
.
Hybridization and introgression across different ploidy levels in the Neotropical orchids Epidendrum fulgens and E. puniceoluteum (Orchidaceae)
.
Molecular Ecology
19
,
3981
3994
.

Vaz Pinto
P
,
Beja
P
,
Ferrand
N
,
Godinho
R
.
2016
.
Hybridization following population collapse in a critically endangered antelope
.
Scientific Reports
6
,
18788
.

Prentice
HC
,
Li
Y
,
Lönn
M
,
Tunlid
A
,
Ghatnekar
L
.
2015
.
A horizontally transferred nuclear gene is associated with microhabitat variation in a natural plant population
.
Proceedings of the Royal Society B
282
,
20152453
.

Pritchard
JK
,
Stephens
M
,
Donnelly
P
.
2000
.
Inference of population structure using multilocus genotype data
.
Genetics
155
,
945
959
.

Racimo
F
,
Marnetto
D
,
Huerta-Sánchez
E
.
2017
.
Signatures of archaic adaptive introgression in present-day human populations
.
Molecular Biology and Evolution
34
,
296
317
.

Raj
A
,
Stephens
M
,
Pritchard
JK
.
2014
.
fastSTRUCTURE: variational inference of population structure in large SNP data sets
.
Genetics
197
,
573
589
.

Ramsey
J
,
Schemske
DW
.
1998
.
Pathways, mechanisms, and rates of polyploid formation in flowering plants
.
Annual Review of Ecology, Evolution, and Systematics
29
,
467
501
.

Renaut
S
,
Grassa
CJ
,
Yeaman
S
,
Moyers
BT
,
Lai
Z
,
Kane
NC
,
Bowers
JE
,
Burke
JM
,
Rieseberg
LH
.
2013
.
Genomic islands of divergence are not affected by geography of speciation in sunflowers
.
Nature Communications
4
,
1827
.

Rice
DW
,
Alverson
AJ
,
Richardson
AO
et al. 
2013
.
Horizontal transfer of entire genomes via mitochondrial fusion in the angiosperm Amborella
.
Science
342
,
1468
1473
.

Rice
DW
,
Palmer
JD
.
2006
.
An exceptional horizontal gene transfer in plastids: gene replacement by a distant bacterial paralog and evidence that haptophyte and cryptophyte plastids are sisters
.
BMC Biology
4
,
31
.

Richardson
AO
,
Palmer
JD
.
2007
.
Horizontal gene transfer in plants
.
Journal of Experimental Botany
58
,
1
9
.

Rieseberg
LH
.
2001
.
Chromosomal rearrangements and speciation
.
Trends in Ecology & Evolution
16
,
351
358
.

Rieseberg
LH
,
Blackman
BK
.
2010
.
Speciation genes in plants
.
Annals of Botany
106
,
439
455
.

Rieseberg
LH
,
Buerkle
CA
.
2002
.
Genetic mapping in hybrid zones
.
The American Naturalist
159
(
Suppl 3
),
S36
S50
.

Rougeux
C
,
Bernatchez
L
,
Gagnaire
P-A
.
2016
.
Modeling the multiple facets of speciation-with-gene-flow towards inferring the divergence history of lake whitefish species pairs (Coregonus clupeaformis)
. bioRxiv doi:
10.1101/068932
.

Roux
C
,
Fraïsse
C
,
Castric
V
,
Vekemans
X
,
Pogson
GH
,
Bierne
N
.
2014
.
Can we continue to neglect genomic variation in introgression rates when inferring the history of speciation? A case study in a Mytilus hybrid zone
.
Journal of Evolutionary Biology
27
,
1662
1675
.

Royer
AM
,
Streisfeld
MA
,
Smith
CI
.
2016
.
Population genomics of divergence within an obligate pollination mutualism: Selection maintains differences between Joshua tree species
.
American Journal of Botany
103
,
1730
1741
.

Saarman
NP
,
Pogson
GH
.
2015
.
Introgression between invasive and native blue mussels (genus Mytilus) in the central California hybrid zone
.
Molecular Ecology
24
,
4723
4738
.

Sankararaman
S
,
Mallick
S
,
Dannemann
M
,
Prüfer
K
,
Kelso
J
,
Pääbo
S
,
Patterson
N
,
Reich
D
.
2014
.
The landscape of Neandertal ancestry in present-day humans
.
Nature
507
,
354
357
.

Sankararaman
S
,
Patterson
N
,
Li
H
,
Pääbo
S
,
Reich
D
.
2012
.
The date of interbreeding between Neandertals and modern humans
.
PLOS Genetics
8
,
e1002947
.

Scascitelli
M
,
Whitney
KD
,
Randell
RA
,
King
M
,
Buerkle
CA
,
Rieseberg
LH
.
2010
.
Genome scan of hybridizing sunflowers from Texas (Helianthus annuus and H. debilis) reveals asymmetric patterns of introgression and small islands of genomic differentiation
.
Molecular Ecology
19
,
521
541
.

Schiffels
S
,
Durbin
R
.
2014
.
Inferring human population size and separation history from multiple genome sequences
.
Nature Genetics
46
,
919
925
.

Schiffels
S
,
Haak
W
,
Paajanen
P
,
Llamas
B
,
Popescu
E
,
Loe
L
,
Clarke
R
,
Lyons
A
,
Mortimer
R
,
Sayer
D
,
Tyler-Smith
C
,
Cooper
A
,
Durbin
R
.
2016
.
Iron Age and Anglo-Saxon genomes from East England reveal British migration history
.
Nature Communications
7
,
10408
.

Schneider
H
,
Schuettpelz
E
,
Pryer
KM
,
Cranfill
R
,
Magallón
S
,
Lupia
R
.
2004
.
Ferns diversified in the shadow of angiosperms
.
Nature
428
,
553
557
.

Schuettpelz
E
,
Pryer
KM
.
2009
.
Evidence for a Cenozoic radiation of ferns in an angiosperm-dominated canopy
.
Proceedings of the National Academy of Sciences of the United States of America
106
,
11200
11205
.

Schmickl
R
,
Koch
MA
.
2011
.
Arabidopsis hybrid speciation processes
.
Proceedings of the National Academy of Sciences of the United States of America
108
,
14192
14197
.

Seehausen
O
.
2004
.
Hybridization and adaptive radiation
.
Trends in ecology & evolution
19
,
198
207
.

Shapiro
BJ
,
Leducq
JB
,
Mallet
J
.
2016
.
What Is Speciation?
PLOS Genetics
12
,
e1005860
.

Smith
J
,
Kronforst
MR
.
2013
.
Do Heliconius butterfly species exchange mimicry alleles?
Biology Letters
9
,
20130503
.

Song
Y
,
Endepols
S
,
Klemann
N
,
Richter
D
,
Matuschka
FR
,
Shih
CH
,
Nachman
MW
,
Kohn
MH
.
2011
.
Adaptive introgression of anticoagulant rodent poison resistance by hybridization between old world mice
.
Current Biology: CB
21
,
1296
1301
.

Soucy
SM
,
Huang
J
,
Gogarten
JP
.
2015
.
Horizontal gene transfer: building the web of life
.
Nature Reviews. Genetics
16
,
472
482
.

Stankowski
S
,
Streisfeld
MA
.
2015
.
Introgressive hybridization facilitates adaptive divergence in a recent radiation of monkeyflowers
.
Proceedings of the Royal Society B
282
,
20151666
.

Starcevic
A
,
Akthar
S
,
Dunlap
WC
,
Shick
JM
,
Hranueli
D
,
Cullum
J
,
Long
PF
.
2008
.
Enzymes of the shikimic acid pathway encoded in the genome of a basal metazoan, Nematostella vectensis, have microbial origins
.
Proceedings of the National Academy of Sciences, USA
105
,
2533
2537
.

Staubach
F
,
Lorenc
A
,
Messer
PW
,
Tang
K
,
Petrov
DA
,
Tautz
D
.
2012
.
Genome patterns of selection and introgression of haplotypes in natural populations of the house mouse (Mus musculus)
.
PLOS Genetics
8
,
e1002891
.

Stegemann
S
,
Bock
R
.
2009
.
Exchange of genetic material between cells in plant tissue grafts
.
Science
324
,
649
651
.

Stegemann
S
,
Keuthe
M
,
Greiner
S
,
Bock
R
.
2012
.
Horizontal transfer of chloroplast genomes between plant species
.
Proceedings of the National Academy of Sciences of the United States of America
109
,
2434
2438
.

Stebbins
GL
.
1971
.
Chromosomal evolution in higher plants
.
London
:
Edward Arnold Ltd
.

Suarez-Gonzalez
A
,
Hefer
CA
,
Christe
C
,
Corea
O
,
Lexer
C
,
Cronk
QC
,
Douglas
CJ
.
2016
.
Genomic and functional approaches reveal a case of adaptive introgression from Populus balsamifera (balsam poplar) in P. trichocarpa (black cottonwood)
.
Molecular Ecology
25
,
2427
2442
.

Sun
BF
,
Li
T
,
Xiao
JH
,
Jia
LY
,
Liu
L
,
Zhang
P
,
Murphy
RW
,
He
SM
,
Huang
DW
.
2015
.
Horizontal functional gene transfer from bacteria to fishes
.
Scientific Reports
5
,
18676
.

Supple
MA
,
Papa
R
,
Hines
HM
,
McMillan
WO
,
Counterman
BA
.
2015
.
Divergence with gene flow across a speciation continuum of Heliconius butterflies
.
BMC Evolutionary Biology
15
,
204
.

Taylor
SA
,
Larson
EL
,
Harrison
RG
.
2015
.
Hybrid zones: windows on climate change
.
Trends in Ecology & Evolution
30
,
398
406
.

Than
C
,
Ruths
D
,
Nakhleh
L
.
2008
.
PhyloNet: a software package for analyzing and reconstructing reticulate evolutionary relationships
.
BMC Bioinformatics
9
,
322
.

The Heliconius Genome Consortium
.
2012
.
Butterfly genome reveals promiscuous exchange of mimicry adaptations among species
.
Nature
487
,
94
98
.

Todesco
M
,
Pascual
MA
,
Owens
GL
et al. 
2016
.
Hybridization and extinction
.
Evolutionary Applications
9
,
892
908
.

Turner
TL
,
Hahn
MW
.
2010
.
Genomic islands of speciation or genomic islands and speciation?
Molecular Ecology
19
,
848
850
.

Turner
TL
,
Hahn
MW
,
Nuzhdin
SV
.
2005
.
Genomic islands of speciation in Anopheles gambiae
.
PLOS Biology
3
,
e285
.

Turner
LM
,
Harr
B
.
2014
.
Genome-wide mapping in a house mouse hybrid zone reveals hybrid sterility loci and Dobzhansky-Muller interactions
.
eLife
3
,
e02504
.

van Wyk
AM
,
Dalton
DL
,
Hoban
S
,
Bruford
MW
,
Russo
IM
,
Birss
C
,
Grobler
P
,
van Vuuren
BJ
,
Kotzé
A
.
2017
.
Quantitative evaluation of hybridization and the impact on biodiversity conservation
.
Ecology and Evolution
7
,
320
330
.

Vernot
B
,
Akey
JM
.
2014
.
Resurrecting surviving Neandertal lineages from modern human genomes
.
Science
343
,
1017
1021
.

Via
S
.
2012
.
Divergence hitchhiking and the spread of genomic isolation during ecological speciation-with-gene-flow
.
Philosophical Transactions of the Royal Society B
367
,
451
460
.

Wang
GX
,
Lv
J
,
Zhang
J
,
Han
S
,
Zong
M
,
Guo
N
,
Zeng
XY
,
Zhang
YY
,
Wang
YP
,
Liu
F
.
2016
.
Genetic and epigenetic alterations of Brassica nigra introgression lines from somatic hybridization: a resource for cauliflower improvement
.
Frontiers in Plant Science
7
,
1258
.

Wang
H
,
Vieira
FG
,
Crawford
JE
,
Chu
C
,
Nielsen
R
.
2017
.
Asian wild rice is a hybrid swarm with extensive gene flow and feralization from domesticated rice
.
Genome Research
27
,
1029
1038
.

Ward
BJ
,
van Oosterhout
C
.
2016
.
HYBRIDCHECK: software for the rapid detection, visualization and dating of recombinant regions in genome sequence data
.
Molecular Ecology Resources
16
,
534
539
.

Weill
M
,
Chandre
F
,
Brengues
C
,
Manguin
S
,
Akogbeto
M
,
Pasteur
N
,
Guillet
P
,
Raymond
M
.
2000
.
The kdr mutation occurs in the Mopti form of Anopheles gambiae s.s. through introgression
.
Insect Molecular Biology
9
,
451
455
.

Wen
D
,
Yu
Y
,
Hahn
MW
,
Nakhleh
L
.
2016
.
Reticulate evolutionary history and extensive introgression in mosquito species revealed by phylogenetic network analysis
.
Molecular Ecology
25
,
2361
2372
.

White
BJ
,
Cheng
C
,
Simard
F
,
Costantini
C
,
Besansky
NJ
.
2010
.
Genetic association of physically unlinked islands of genomic divergence in incipient species of Anopheles gambiae
.
Molecular Ecology
19
,
925
939
.

Whitney
KD
,
Broman
KW
,
Kane
NC
,
Hovick
SM
,
Randell
RA
,
Rieseberg
LH
.
2015
.
QTL mapping identifies candidate alleles involved in adaptive introgression and range expansion in a wild sunflower
.
Molecular Ecology
24
,
2194
2211
.

Whitney
KD
,
Randell
RA
,
Rieseberg
LH
.
2010
.
Adaptive introgression of abiotic tolerance traits in the sunflower Helianthus annuus
.
New Phytologist
187
,
230
239
.

Wilcox
JT
,
Padgett-Flohr
GE
,
Alvarez
JA
,
Johnson
JR
.
2015
.
Possible phenotypic influence of superinvasive alleles on larval California tiger salamanders (Ambystoma californiense)
.
The American Midland Naturalist
173
,
168
175
.

Wolf
JB
,
Ellegren
H
.
2017
.
Making sense of genomic islands of differentiation in light of speciation
.
Nature reviews. Genetics
18
,
87
100
.

Wolf
DE
,
Takebayashi
N
,
Rieseberg
LH
.
2001
.
Predicting the risk of extinction through hybridization
.
Conservation Biology
15
,
1039
1053
.

Won
H
,
Renner
SS
.
2003
.
Horizontal gene transfer from flowering plants to Gnetum
.
Proceedings of the National Academy of Sciences, USA
100
,
10824
10829
.

Wu
CI
,
Ting
CT
.
2004
.
Genes and speciation
.
Nature reviews. Genetics
5
,
114
122
.

Wybouw
N
,
Pauchet
Y
,
Heckel
DG
,
Van Leeuwen
T
.
2016
.
Horizontal Gene Transfer Contributes to the Evolution of Arthropod Herbivory
.
Genome Biology and Evolution
8
,
1785
1801
.

Xi
Z
,
Bradley
RK
,
Wurdack
KJ
,
Wong
K
,
Sugumaran
M
,
Bomblies
K
,
Rest
JS
,
Davis
CC
.
2012
.
Horizontal transfer of expressed genes in a parasitic flowering plant
.
BMC Genomics
13
,
227
.

Yang
Z
,
Liu
L
,
Fang
H
,
Li
P
,
Xu
S
,
Cao
W
,
Xu
C
.
2016
.
Origin of the plant Tm-1-like gene via two independent horizontal transfer events and one gene fusion event
.
Scientific Reports
6
,
33691
.

Yang
Z
,
Wang
Y
,
Zhou
Y
,
Gao
Q
,
Zhang
E
,
Zhu
L
,
Hu
Y
,
Xu
C
.
2013
.
Evolution of land plant genes encoding L-Ala-D/L-Glu epimerases (AEEs) via horizontal gene transfer and positive selection
.
BMC Plant Biology
13
,
34
.

Yang
Z
,
Zhang
Y
,
Wafula
EK
et al. 
2016
.
Horizontal gene transfer is more frequent with increased heterotrophy and contributes to parasite adaptation
.
Proceedings of the National Academy of Sciences, USA
113
,
E7010
E7019
.

Yant
L
,
Hollister
JD
,
Wright
KM
,
Arnold
BJ
,
Higgins
JD
,
Franklin
FCH
,
Bomblies
K
.
2013
.
Meiotic adaptation to genome duplication in Arabidopsis arenosa
.
Current Biology: CB
23
,
2151
2156
.

Yin
Z
,
Zhu
B
,
Feng
H
,
Huang
L
.
2016
.
Horizontal gene transfer drives adaptive colonization of apple trees by the fungal pathogen Valsa mali
.
Scientific Reports
6
,
33129
.

Yoshida
S
,
Maruyama
S
,
Nozaki
H
,
Shirasu
K
.
2010
.
Horizontal gene transfer by the parasitic plant Striga hermonthica
.
Science
328
,
1128
.

Yue
J
,
Hu
X
,
Huang
J
.
2013
.
Horizontal gene transfer in the innovation and adaptation of land plants
.
Plant Signaling & Behavior
8
,
e24130
.

Yue
J
,
Hu
X
,
Sun
H
,
Yang
Y
,
Huang
J
.
2012
.
Widespread impact of horizontal gene transfer on plant colonization of land
.
Nature Communications
3
,
1152
.

Zalapa
JE
,
Brunet
J
,
Guries
RP
.
2009
.
Patterns of hybridization and introgression between invasive Ulmus pumila (Ulmaceae) and native U. rubra
.
American Journal of Botany
96
,
1116
1128
.

Zhang
W
,
Dasmahapatra
KK
,
Mallet
J
,
Moreira
GR
,
Kronforst
MR
.
2016
.
Genome-wide introgression among distantly related Heliconius butterfly species
.
Genome Biology
17
,
25
.

Zhang
Y
,
Fernandez-Aparicio
M
,
Wafula
EK
et al. 
2013
.
Evolution of a horizontally acquired legume gene, albumin 1, in the parasitic plant Phelipanche aegyptiaca and related species
.
BMC Evolutionary Biology
13
,
48
.

Zhang
D
,
Qi
J
,
Yue
J
et al. 
2014
.
Root parasitic plant Orobanche aegyptiaca and shoot parasitic plant Cuscuta australis obtained Brassicaceae-specific strictosidine synthase-like genes by horizontal gene transfer
.
BMC Plant Biology
14
,
19
.

Zhu
B
,
Lou
MM
,
Xie
GL
,
Zhang
GQ
,
Zhou
XP
,
Li
B
,
Jin
GL
.
2011
.
Horizontal gene transfer in silkworm, Bombyx mori
.
BMC Genomics
12
,
248
.

Zohren
J
,
Wang
N
,
Kardailsky
I
,
Borrell
JS
,
Joecker
A
,
Nichols
RA
,
Buggs
RJ
.
2016
.
Unidirectional diploid-tetraploid introgression among British birch trees with shifting ranges shown by restriction site-associated markers
.
Molecular Ecology
25
,
2413
2426
.

Comments

0 Comments
Submit a comment
You have entered an invalid code
Thank you for submitting a comment on this article. Your comment will be reviewed and published at the journal's discretion. Please check for further notifications by email.