Skip to main content
Interface Focus logoLink to Interface Focus
. 2017 Aug 18;7(5):20160115. doi: 10.1098/rsfs.2016.0115

Biological action in Read–Write genome evolution

James A Shapiro 1,
PMCID: PMC5566801  PMID: 28839913

Abstract

Many of the most important evolutionary variations that generated phenotypic adaptations and originated novel taxa resulted from complex cellular activities affecting genome content and expression. These activities included (i) the symbiogenetic cell merger that produced the mitochondrion-bearing ancestor of all extant eukaryotes, (ii) symbiogenetic cell mergers that produced chloroplast-bearing ancestors of photosynthetic eukaryotes, and (iii) interspecific hybridizations and genome doublings that generated new species and adaptive radiations of higher plants and animals. Adaptive variations also involved horizontal DNA transfers and natural genetic engineering by mobile DNA elements to rewire regulatory networks, such as those essential to viviparous reproduction in mammals. In the most highly evolved multicellular organisms, biological complexity scales with ‘non-coding’ DNA content rather than with protein-coding capacity in the genome. Coincidentally, ‘non-coding’ RNAs rich in repetitive mobile DNA sequences function as key regulators of complex adaptive phenotypes, such as stem cell pluripotency. The intersections of cell fusion activities, horizontal DNA transfers and natural genetic engineering of Read–Write genomes provide a rich molecular and biological foundation for understanding how ecological disruptions can stimulate productive, often abrupt, evolutionary transformations.

Keywords: symbiogenesis, hybrid speciation, horizontal DNA transfer, mobile DNA, network rewiring, non-coding RNA regulation

1. Living organisms regularly facilitate their own evolution

The most basic questions in evolutionary biology concern the origins of taxonomic and adaptive innovations. How do new kinds of organisms arise and acquire heritable functionalities that enable them to proliferate in a changing ecology? While the details of each innovation are unique, genome analysis shows that diverse evolutionary novelties repeatedly originated through core generic activities:

  • — horizontal DNA transfers between unrelated cell lineages [13];

  • — symbiogenetic cell mergers [4,5];

  • — interspecific hybridizations [6,7];

  • — biochemical activities that cleave, splice, polymerize and otherwise modify cell DNA molecules (collectively, ‘natural genetic engineering’ or NGE activities) to generate novel sequence configurations and amplify mobile DNA elements [8].

The following sections present examples of these generic processes repeatedly achieving evolutionary innovations. These cases exemplify broader implications of biological agency (i.e. direct biochemical and biomechanical activity) in generating evolutionary variation:

  • (1.a.) Major classes of genome modification (horizontal transfer, symbiogenesis, hybridization) involve the interaction of organisms with distinct evolutionary histories. Consequently, a full account of many evolutionary innovations involves more than a single branching vertical line of descent, as commonly depicted [9]. A major fraction of evolutionary pedigrees have an obligatory networked or reticulate architecture [10].

  • (1.b.) Since cells have agency in generating organisms with new genome configurations, the genome functions as a read–write (RW) storage system in evolution [11]. The RW genome is subject to a range of generic inscriptions (e.g. acquisition of DNA molecules encoding complex metabolic capabilities, relocation of information-rich sequence cassettes) with greater functional consequences than the random copying errors postulated to occur in the conventional read-only memory (ROM) genome [12,13].

  • (1.c.) Like all classes of cellular biochemistry, NGE DNA transport and restructuring functions are subject to control by regulatory circuits and respond to changing conditions (electronic supplementary material, table S1). Among the conditions known to activate NGE functions are cell mergers, such as interspecific hybridizations [14,15]. By altering the regulatory context, one kind of genome-modification activity can stimulate other kinds in a positive feedback loop that amplifies evolutionary innovation.

  • (1.d.) Genome changes by symbiogenetic cell fusions, interspecific hybridizations and multilocus NGE activities typically affect multiple characters of the variant cell and organism. Consequently, major phenotypic transformations can occur in a single evolutionary episode and are not restricted to a gradual accumulation of ‘numerous, successive, slight modifications’ [9]. In other words, there is an empirical molecular–cellular basis for saltationist views of the evolutionary process, similar to those proposed by certain early evolutionists [1619].

2. Many lineages acquire adaptive innovations through horizontal DNA transfers rather than reinventing each function de novo

A major fraction of all biomass on Earth consists of prokaryotic cells without a nucleus carrying out most biogeochemical transformations in a complex web of interactions with nucleated eukaryotic microbes (protists) [20,21]. Thanks to the pioneering work of Carl Woese and colleagues on ribosomal rRNA sequences characterizing fundamental cell protein synthesis organelles, we know that there are two distinct evolutionary lineages of prokaryotes, Eubacteria and Archaea [22,23]. In addition to ribosomal differences, the two prokaryotic lineages differ in other core cell functionalities, such as DNA replication machinery, RNA transcription complexes and membrane structure.

Well before the discovery of Archaea in the 1970s, it was clear from basic bacterial genetics that prokaryotic cells have multiple molecular mechanisms for horizontal DNA transfer: uptake of naked DNA from the environment (‘transformation’), encapsidation of cellular DNA in infectious virus particles (‘transduction’), and direct contact-mediated transfer (‘conjugation’) [1,24]. Rapid appreciation of bacterial virtuosity in horizontal DNA transfer resulted from the need to understand the evolution and spread of infectious multiple antibiotic resistance in pathogenic bacteria [25] (electronic supplementary material, table S2). The discovery of transmissible drug resistance plasmids provided an early real-world object lesson on the limitations of evolutionary thinking based solely on spontaneous mutation and vertical inheritance of novel traits.

The lessons of antibiotic resistance evolution proved applicable to other bacterial adaptations. Diverse bacteria possess extended DNA structures (labelled ‘genome islands’ or ‘integrons’) encoding complex functionalities, like catabolic pathways or pathogenicity determinants, that they can transfer to unrelated cells, where they integrate into the genome and abruptly extend the recipient's adaptive capabilities [2628].

The existence of transmissible DNA cassettes that generate sudden adaptive changes prompted rethinking of how an ecologically suitable genome evolves in the prokaryotic realm. The prevalence of horizontal DNA exchange gave rise to the idea of a vast supra-cellular pan-genome that bacterial cells sample in a facultative manner to evolve novel genomes suitable for adaptation to particular ecological niches [29]. This initially controversial notion gained wider credibility as metagenomic analysis revealed that extracellular environmental DNA samples contain unexpectedly large numbers of potentially adaptive coding elements in virus particles [30]. These environmental DNA elements include sequences encoding previously unknown functionalities [31]. Detailed single cell analysis of soil bacteria indicates virus infections and horizontal DNA transfers are active ongoing events [32].

Although the discrete origins of Eubacteria and Archaea remain highly speculative [33,34], patterns of biochemical specializations in the two kingdoms provide evidence for significant inter-kingdom DNA transfer. Horizontal transfer was initially apparent between hyperthermophilic archaeal and eubacterial cells that shared a common high-temperature ecology [35]. Subsequent genomics-based analysis of Archaea from more temperate ecologies showed that transfer of DNA encoding hundreds of metabolic functions from eubacteria correlated with the origination of novel mesophilic archaeal clades [3639]. While there is some debate about the number and speed of Eubacteria–Archaea horizontal transfers [40], the sequence evidence clearly indicates that evolutionary innovation among Archaea only makes sense by invoking DNA exchanges with Eubacteria. These discoveries also fit with the Eubacteria–Archaea transfers now recognized as necessary to explain the origin of nucleated eukaryotic cells from prokaryotic ancestors [38,41,42].

Horizontal exchange of adaptively important DNA has also been important in eukaryotic evolution, although less pervasive than in prokaryotes [4346] (electronic supplementary material, table S2). Many bacteria-to-eukaryote transfers originate from endosymbiotic bacteria [47,48]. Intriguing horizontal exchanges include the transfer from bacteria to primitive land plants (mosses) of xylem synthesis and other functions needed for land colonization [49] and transfers from both bacteria and fungi to plant parasitic nematode worms of biochemical activities for phytopolymer digestion [5052]. In the latter case, we know there were multiple independent transfers because related nematode species use distinct microbe-derived enzymes to break down plant material.

3. Symbiogenetic cell fusions formed the ancestral eukaryote and major photosynthetic eukaryotic clades

We know that evolution of the ancestor of all nucleated eukaryotic cells involved a symbiogenetic event. Eukaryotic cells all possess an oxidative mitochondrion organelle or a defective non-oxidative derivative (hydrogenosome or mitosome) [42,5355]. While mitochondria and derivatives are quite diverse, all contain organelle genomes encoding organelle-specific ribosomal rRNA molecules. These rRNA molecules clearly identify the organelles as descendants of an ancestral alpha-proteobacteria endosymbiont [56,57]. Thus, a saltatory cell fusion event has to be added to horizontal DNA transfers as contributing to one of the most important evolutionary innovations in life history.

The ancestral eukaryote had two genomic compartments: (i) the nucleus, encoding eukaryotic rRNA and most proliferative functions, and (ii) the mitochondrion, encoding mitochondrial rRNA and oxidative metabolic functions from the alpha-proteobacterium endosymbiont. A major feature of subsequent eukaryotic evolutionary diversification has been loss from mitochondrial genomes of sequences encoding functions needed for autonomous cell reproduction as well as intracellular horizontal transfer to the nuclear genome of DNA sequences encoding mitochondrial proteins [56,58]. The processes of mitochondrial genome reduction and restructuring have been different in distinct eukaryotic lineages (including those that have lost oxidative metabolism), producing a great variety of taxonomically specific organelle genotypes [54,5961]. Eukaryotic diversification thus resulted from biochemical NGE functions executing deletions, transfers and rearrangements of mitochondrial DNA.

Symbiogenetic cell fusions have been ongoing features of major steps in eukaryotic evolution. The origin of photosynthetic eukaryotes harbouring additional chloroplast/plastid organelles can be traced by rRNA sequencing to an ancestral cyanobacterial endosymbiont [53,62]. The cell product of this primordial symbiogenesis contained three genome compartments—nucleus, mitochondrion and plastid—and subsequent plastid–nuclear transfers and deletions gave rise to four primary photosynthetic eukaryotic lineages: (i) green algae, (ii) land plants (embryophytes), (iii) glaucophytes and (iv) red algae [53,62]. The evolutionary history of photosynthetic eukaryotes diversified further by higher-level symbiogenetic events, in which red or green algae became endosymbionts of distinct eukaryotic lineages (table 1). The cell arising from such a secondary endosymbiosis has four distinct genome compartments: (i) nucleus, (ii) mitochondrion, (iii) plastid and (iv) nucleomorph (derived from the nucleus of the algal endosymbiont). As with mitochondria, deletions, intracellular transfers and rearrangements of the DNA in all these compartments contribute to evolutionary innovation.

Table 1.

Photosynthetic eukaryotic lineages resulting from symbiogenesis [53,62].

taxonomic group symbiogenetic origin
green algae (Chlorophyta) primary cyanobacterial endosymbiosis
land plants (Embryophyta) primary cyanobacterial endosymbiosis
Glaucophytes (order Chlorococcales) primary cyanobacterial endosymbiosis
red algae (Rhodophyta) primary cyanobacterial endosymbiosis
Euglenids (flagellated algae) secondary green algal endosymbiosis
Chlorarachniophytes (marine algae) secondary green algal endosymbiosis
Chromalveolates (multiple lineages including major photosynthetic organisms responsible for a large fraction of atmospheric oxygen, such as brown algae, dinoflagellates and diatoms) secondary red algal endosymbiosis

Among the dinoflagellates, there is a remarkable photosynthetic taxon labelled warnowiids, which comprises individual photosynthetic cells containing a light-harvesting organelle (the ‘ocelloid’) that resembles the camera eyes of multicellular animals [63]. The ocelloid has analogues to the cornea, lens, iris and retina. A noteworthy recent paper reports genomic analysis showing that two of these structures resulted from distinct endosymbiogenetic events, the cornea composed of mitochondria and the retina of red algal plastids [64]. The role of cell fusion events in the evolution of the ocelloid eye-like structure stands in sharp contrast to Darwin's conception of eyes evolving ‘by numerous, successive, slight modifications’ (p. 189) [9].

4. Interspecific hybridizations and whole genome duplications accelerate eukaryotic speciation and taxonomic diversification

Although Darwin emphasized repeated selection of individual traits as a source of new species, using human selective breeding as an illustration [9], major domesticated crop species with beneficial agricultural characters like wheat, cotton and tobacco actually arose by a completely different ‘cataclysmic evolution’ process: cross-breeding of different species to form genotypically and phenotypically novel hybrids [65,66]. For example, several thousand years ago in the Middle East, ‘the combination of chromosomes of a moderately useful plant, emmer wheat, and those of a completely useless and noxious weed {wild goat grass, Aegilops squarrosa} produced the world's most valuable crop plant,’ high-yielding bread wheat, and this hybrid speciation process can be repeated in real time [65].

Because organisms with two different haploid chromosome sets will not proceed through meiosis to reproduce sexually, successful fertile interspecific hybridizations also involve whole genome duplication (WGD) events so that the merged ‘allopolyploid’ genome has the necessary two copies of each parental chromosome. Genome doubling is a common response to interspecific hybridization, at least in plants where it has most been studied [6769]. The cytogenetic and genomic evidence for hybrid speciation among yeasts, plants and animals is abundant (table 2) [133136]. Two rapidly evolving species groups in table 2 merit special emphasis: Darwin's Galapagos finches (Geospiza), often taken as models of gradualist selection-driven speciation, and East African dichlids (fresh water fish), recent examples of explosive adaptive radiations [137].

Table 2.

Selected examples of speciation and adaptive radiation by interspecific hybridization and changes in chromosome number.

taxon references
fungi [70]
Saccharomyces yeast [7177]
plants [69,7881]
Tragopogon (Asteraceae) [82,83]
 irises (Iris fulva, I. hexagona, and I. nelsonii) [8486]
Nicotiana (Solanaceae) [8789]
Brassica napus [9092]
Arabidopsis [9395]
 potatoes (Solanum stoloniferum and S. hjertingii) [9698]
 wheats (Aegilops–Triticum group) [65,66,99,100]
 cotton (Gossypium) [101,102]
animals [103105]
 tephritid fruit flies [106,107]
Heliconius butterflies [108112]
 sailfin silversides (Teleostei) [113,114]
 East African cichlids [115121]
 sparrows [122,123]
 Galapagos finches (Geospiza) [124130]
 Clymene dolphin (Delphininae (Cetacea, Mammalia)) [131,132]

An important feature of evolutionary innovation by interspecific hybridization is that the transformational event involves the entire genomes of both parental species. Thus, all organismal traits are simultaneously subject to modification. Molecular studies find dramatically altered genome-wide expression patterns in newly synthesized interspecific hybrids, resulting at least in part from epigenetic modifications across the hybrid genomes [138,139]. The same regulatory and epigenetic changes that alter coding sequence expression in hybrid nuclei also de-repress transcription of mobile DNA elements and other NGE functions [140]. Repetitive and mobile DNA elements play special roles in chromosome restructuring [141144], as exemplified in primate evolution by the recently published gibbon genome [145]. Thus, in addition to global effects on genome expression, interspecific hybridization serves as a major stimulus to genome variability, including chromosome rearrangements and mobile DNA activation [14,15,82,146149] (electronic supplementary material, table S3). Chromosome rearrangements have long been recognized as major features of speciation and taxonomic divergence [150152]. Accordingly, hybrid organisms have a markedly elevated potential for generating novel DNA sequence configurations not present in either parent species genome, an added impact of biological agency on evolutionary innovation [45].

5. Amplification of mobile DNA accompanies increased organismal complexity and provides information-rich cassettes for adaptive innovations

The fact that interspecific hybridization activates the spread and amplification of mobile DNA elements (MEs) fits into a larger set of connections between this particular class of genome sequences and the origins of novel adaptations. All organisms contain repetitive DNA elements capable of transposing to novel genome locations [153]. Both transposition purely at the DNA level and retrotransposition through RNA intermediates tend to increase the copy number of a particular element, and MEs are collectively referred to as dispersed repetitive DNA.

The class of mobile repetitive DNA has typically been considered functionally distinct from the relatively fixed, low copy DNA encoding the vast majority of organismal proteins. Hence, it was significant when the initial draft human genome sequence revealed that at least 44% of our genomes consisted of MEs [154]. (Today, our fraction of repetitive DNA is estimated to be as high as 67% [155].) In fact, the repetitive, so-called ‘non-coding’ content of the genome scales more closely with organismal complexity defined by the number of distinct cell types than does the protein-coding content, which plateaus in organisms of about a dozen different cell types [156]. In other words, MEs and other ‘non-coding’ sequences are especially prevalent in genomes of the most highly evolved organisms.

Focusing our attention on mammals, where the most in-depth genomic analysis has been performed, we can discern at least three major evolutionary roles for MEs [157]:

  • (1) MEs provide distributed signals, such as binding site for structural proteins and replication factors, to format the genome as a physically organized self-replicating sequence database [158160];

  • (2) Although characterized by some as non-functional ‘junk DNA’ [161,162], MEs provide cis-regulatory signals to rewire transcriptional networks determining multiple higher-level traits, including key mammalian innovations on both uterine and placental sides of the maternal–fetal interface in viviparous reproduction [163166];

  • (3) MEs provide the majority of conserved sequences comprising taxonomically specific ‘long non-coding lncRNAs' [167169], a class of molecule that we increasingly recognize as modulating epigenetic control [170] and regulating complex traits like stem cell pluripotency [171174], internal organ and nervous system development [175177], and innate and adaptive immunity [178,179].

These three categories, plus dozens of other examples cited in the various tables of [157], provide robust support for the RW genome view of MEs as NGE tools (plug-in formatting cassettes) for adaptive genome restructuring. Viewing MEs as genome modification tools has been in the literature ever since McClintock showed transposable ‘controlling elements’ to alter developmental patterns [180] and molecular studies in bacteria and yeast revealed MEs to act as sources of regulatory innovation [181183]. Significantly, ecological stress triggers mobile DNA activity [8] (electronic supplementary material, table S1).

6. Core biological capacities rewrite genomes for evolutionary innovation

The preceding discussion illustrates how generic biological activities have regularly played decisive roles in major episodes of evolutionary innovation:

  • — horizontal DNA transfers in the origination of mesophilic archaeal taxa;

  • — recurring symbiogenetic cell fusions in the origination of the ancestral eukaryotic cell and the major clades of photosynthetic eukaryotes;

  • — interspecific hybridizations and changes in genome ploidy in speciation and adaptive radiations in yeast, plants and animals;

  • — amplification and relocalization of mobile DNA elements in formatting mammalian genomes for replication, viviparous reproduction, and lncRNA regulation of nervous and immune system functions.

These examples show us that core biological capacities for self-modification in response to ecological challenge have been integral to the history of life on earth. That conclusion should not surprise us because extant organisms are descendants of multiple evolutionary episodes. Considering potential interactions between dynamic ecological conditions and the biological engines of cell and genome variation raises important questions about control and specificity in evolutionary innovation. The years to come likely hold surprising lessons about how cell fusions, genome doublings, and natural genetic engineering may operate non-randomly to enhance the probabilities of evolutionary success.

Supplementary Material

Supplementary Table 1. Various stimuli documented to activate natural genetic engineering
rsfs20160115supp1.pdf (186.9KB, pdf)

Supplementary Material

Supplementary Table 2. Examples of horizontal DNA transfer
rsfs20160115supp2.pdf (281.7KB, pdf)

Supplementary Material

Supplementary Table 3. Genomic responses to changes in ploidy and interspecific hybridization in fungi, plants and animals.pdf
rsfs20160115supp3.pdf (124.5KB, pdf)

Acknowledgements

The author is grateful for the invitation to speak at the Royal Society-British Academy meeting on new perspectives in evolutionary thinking.

Data accessibility

This article has no additional data.

Authors' contributions

J.A.S. is the sole author of this review.

Funding

The author received no outside funds for preparing this manuscript.

Competing interests

The author declares no competing interests.

References

  • 1.Syvanen M, Kado CI. 2002. Horizontal gene transfer, 2nd edn London, UK: Academic Press. [Google Scholar]
  • 2.Syvanen M. 2012. Evolutionary implications of horizontal gene transfer. Annu. Rev. Genet. 46, 341–358. ( 10.1146/annurev-genet-110711-155529) [DOI] [PubMed] [Google Scholar]
  • 3.Fournier GP, Huang J, Gogarten JP. 2009. Horizontal gene transfer from extinct and extant lineages: biological innovation and the coral of life. Phil. Trans. R. Soc. B 364, 2229–2239. ( 10.1098/rstb.2009.0033) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Margulis L, Sagan D. 2002. Acquiring genomes: a theory of the origins of species. Amherst, MA: Perseus Books Group. [Google Scholar]
  • 5.Archibald J. 2014. One plus one equals one: symbiosis and the evolution of complex life. Oxford, UK: Oxford University Press. [Google Scholar]
  • 6.Arnold ML. 2006. Evolution through genetic exchange. Oxford, UK: Oxford University Press. [Google Scholar]
  • 7.Arnold ML. 2015. Divergence with genetic exchange. Oxford, UK: Oxford University Press. [Google Scholar]
  • 8.Shapiro JA. 2011. Evolution: a view from the 21st century. Upper Saddle River, NJ: FT Press Science. [Google Scholar]
  • 9.Darwin C. 1859. Origin of species. London, UK: John Russel. [Google Scholar]
  • 10.Gontier N. (ed.). 2015. Reticulate evolution: symbiogenesis, lateral gene transfer, hybridization and infectious heredity. Heidelberg, Germany: Springer International Publishing AG Switzerland. [Google Scholar]
  • 11.Shapiro JA. 2013. How life changes itself: the read–write (RW) genome. Phys. Life Rev. 10, 287–323. ( 10.1016/j.plrev.2013.07.001) [DOI] [PubMed] [Google Scholar]
  • 12.Dawkins R. 1976. The selfish gene. Oxford, UK: Oxford University Press. [Google Scholar]
  • 13.Dennett D. 1996. Darwin's dangerous idea: evolution and the meanings of life. London, UK: Penguin Books Limited. [Google Scholar]
  • 14.Guerreiro MP. 2014. Interspecific hybridization as a genomic stressor inducing mobilization of transposable elements in Drosophila. Mob. Genet. Elements 4, e34394 ( 10.4161/mge.34394) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Parisod C, Alix K, Just J, Petit M, Sarilar V, Mhiri C, Ainouche M, Chalhoub B, Grandbastien MA. 2010. Impact of transposable elements on the organization and function of allopolyploid genomes. New Phytol. 186, 37–45. ( 10.1111/j.1469-8137.2009.03096.x) [DOI] [PubMed] [Google Scholar]
  • 16.Bateson W. 1894. Materials for the study of variation treated with especial regard to discontinuity in the origin of species. London, UK: Macmillan. [Google Scholar]
  • 17.de Vries H. 1905. Species and varieties, their origin by mutation; lectures delivered at the University of California. Chicago, IL: Open Court Publishing Co. [Google Scholar]
  • 18.Goldschmidt R. 1940. The material basis of evolution, reissued (The Silliman Memorial Lectures Series), 1982. New Haven, CT: Yale Univ. Press. [Google Scholar]
  • 19.Dittrich-Reed DR, Fitzpatrick BM. 2013. Transgressive hybrids as hopeful monsters. Evol. Biol. 40, 310–315. ( 10.1007/s11692-012-9209-0) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Whitman WB, Coleman DC, Wiebe WJ. 1998. Prokaryotes: the unseen majority. Proc. Natl Acad. Sci. USA 95, 6578–6583. ( 10.1073/pnas.95.12.6578) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Gast RJ, Sanders RW, Caron DA. 2009. Ecological strategies of protists and their symbiotic relationships with prokaryotic microbes. Trends Microbiol. 17, 563–569. ( 10.1016/j.tim.2009.09.001) [DOI] [PubMed] [Google Scholar]
  • 22.Woese CR, Fox GE. 1977. Phylogenetic structure of the prokaryotic domain: the primary kingdoms. Proc. Natl Acad. Sci. USA 74, 5088–5090. ( 10.1073/pnas.74.11.5088) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Woese CR, Kandler O, Wheelis ML. 1990. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc. Natl Acad. Sci. USA 87, 4576–4579. ( 10.1073/pnas.87.12.4576) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Hayes W. 1968. The genetics of bacteria and their viruses, 2nd edn London, UK: Blackwell. [Google Scholar]
  • 25.Watanabe T. 1963. Infective heredity of multiple drug resistance in bacteria. Bacteriol. Rev. 27, 87–115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Daccord A, Ceccarelli D, Rodrigue S, Burrus V. 2013. Comparative analysis of mobilizable genomic islands. J. Bacteriol. 195, 606–614. ( 10.1128/jb.01985-12) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Bellanger X, Payot S, Leblond-Bourget N, Guedon G. 2014. Conjugative and mobilizable genomic islands in bacteria: evolution and diversity. FEMS Microbiol. Rev. 38, 720–760. ( 10.1111/1574-6976.12058) [DOI] [PubMed] [Google Scholar]
  • 28.Escudero JA, Loot C, Nivina A, Mazel D. 2015. The integron: adaptation on demand. Microbiol. Spectr. 3, MDNA3–0019–2014 ( 10.1128/microbiolspec.MDNA3-0019-2014) [DOI] [PubMed] [Google Scholar]
  • 29.Sonea S, Panisset M. 1983. A new bacteriology. Boston, MA: Jones and Batlett. [Google Scholar]
  • 30.Sharon I, et al. 2011. Comparative metagenomics of microbial traits within oceanic viral communities. ISME J. 5, 1178–1190. ( 10.1038/ismej.2011.2) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Ufarte L, Potocki-Veronese G, Laville E. 2015. Discovery of new protein families and functions: new challenges in functional metagenomics for biotechnologies and microbial ecology. Front. Microbiol. 6, 563 ( 10.3389/fmicb.2015.00563) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Labonte JM, Field EK, Lau M, Chivian D, Van Heerden E, Wommack KE, Kieft TL, Onstott TC, Stepanauskas R. 2015. Single cell genomics indicates horizontal gene transfer and viral infections in a deep subsurface Firmicutes population. Front. Microbiol. 6, 349 ( 10.3389/fmicb.2015.00349) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Woese CR. 2002. On the evolution of cells. Proc. Natl Acad. Sci. USA 99, 8742–8747. ( 10.1073/pnas.132266999) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Bapteste E, et al. 2009. Prokaryotic evolution and the tree of life are two different things. Biol. Direct 4, 34 ( 10.1186/1745-6150-4-34) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Aravind L, Tatusov RL, Wolf YI, Walker DR, Koonin EV. 1998. Evidence for massive gene exchange between archaeal and bacterial hyperthermophiles. Trends Genet. 14, 442–444. ( 10.1016/S0168-9525(98)01553-4) [DOI] [PubMed] [Google Scholar]
  • 36.Nelson-Sathi S, et al. 2015. Origins of major archaeal clades correspond to gene acquisitions from bacteria. Nature 517, 77–80. ( 10.1038/nature13805) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Lopez-Garcia P, Zivanovic Y, Deschamps P, Moreira D. 2015. Bacterial gene import and mesophilic adaptation in archaea. Nat. Rev. Microbiol. 13, 447–456. ( 10.1038/nrmicro3485) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Akanni WA, Siu-Ting K, Creevey CJ, McInerney JO, Wilkinson M, Foster PG, Pisani D. 2015. Horizontal gene flow from Eubacteria to Archaebacteria and what it means for our understanding of eukaryogenesis. Phil. Trans. R. Soc. B 370, 20140337 ( 10.1098/rstb.2014.0337) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Meheust R, Lopez P, Bapteste E. 2015. Metabolic bacterial genes and the construction of high-level composite lineages of life. Trends Ecol. Evol. 30, 127–129. ( 10.1016/j.tree.2015.01.001) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Groussin M, Boussau B, Szollosi G, Eme L, Gouy M, Brochier-Armanet C, Daubin V. 2016. Gene acquisitions from bacteria at the origins of major archaeal clades are vastly overestimated. Mol. Biol. Evol. 33, 305–310. ( 10.1093/molbev/msv249) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Williams TA, Embley TM. 2015. Changing ideas about eukaryotic origins. Phil. Trans. R. Soc. B 370, 20140318 ( 10.1098/rstb.2014.0318) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Martin WF, Garg S, Zimorski V. 2015. Endosymbiotic theories for eukaryote origin. Phil. Trans. R. Soc. B 370, 20140330 ( 10.1098/rstb.2014.0330) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Swithers KS, Soucy SM, Gogarten JP. 2012. The role of reticulate evolution in creating innovation and complexity. Int. J. Evol. Biol. 2012, 418964 ( 10.1155/2012/418964) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Soucy SM, Huang J, Gogarten JP. 2015. Horizontal gene transfer: building the web of life. Nat. Rev. Genet. 16, 472–482. ( 10.1038/nrg3962) [DOI] [PubMed] [Google Scholar]
  • 45.Mallet J, Besansky N, Hahn MW. 2016. How reticulated are species? Bioessays 38, 140–149. ( 10.1002/bies.201500149) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Crisp A, Boschetti C, Perry M, Tunnacliffe A, Micklem G. 2015. Expression of multiple horizontally acquired genes is a hallmark of both vertebrate and invertebrate genomes. Genome Biol. 16, 50 ( 10.1186/s13059-015-0607-3) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Ku C, et al. 2015. Endosymbiotic origin and differential loss of eukaryotic genes. Nature 524, 427–432. ( 10.1038/nature14963) [DOI] [PubMed] [Google Scholar]
  • 48.Keeling PJ, Palmer JD. 2008. Horizontal gene transfer in eukaryotic evolution. Nat. Rev. Genet. 9, 605–618. ( 10.1038/nrg2386) [DOI] [PubMed] [Google Scholar]
  • 49.Yue J, Hu X, Sun H, Yang Y, Huang J. 2012. Widespread impact of horizontal gene transfer on plant colonization of land. Nat. Commun. 3, 1152 ( 10.1038/ncomms2148) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Haegeman A, Jones JT, Danchin EG. 2011. Horizontal gene transfer in nematodes: a catalyst for plant parasitism? Mol. Plant Microbe Interact. 24, 879–887. ( 10.1094/mpmi-03-11-0055) [DOI] [PubMed] [Google Scholar]
  • 51.Mayer WE, Schuster LN, Bartelmes G, Dieterich C, Sommer RJ. 2011. Horizontal gene transfer of microbial cellulases into nematode genomes is associated with functional assimilation and gene turnover. BMC Evol. Biol. 11, 13 ( 10.1186/1471-2148-11-13) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Mitreva M, Smant G, Helder J. 2009. Role of horizontal gene transfer in the evolution of plant parasitism among nematodes. Methods Mol. Biol. 532, 517–535. ( 10.1007/978-1-60327-853-9_30) [DOI] [PubMed] [Google Scholar]
  • 53.Embley TM, Martin W. 2006. Eukaryotic evolution, changes and challenges. Nature 440, 623–630. ( 10.1038/nature04546) [DOI] [PubMed] [Google Scholar]
  • 54.van der Giezen M. 2009. Hydrogenosomes and mitosomes: conservation and evolution of functions. J. Eukaryot. Microbiol. 56, 221–231. ( 10.1111/j.1550-7408.2009.00407.x) [DOI] [PubMed] [Google Scholar]
  • 55.McInerney J, Pisani D, O'Connell MJ. 2015. The ring of life hypothesis for eukaryote origins is supported by multiple kinds of data. Phil. Trans. R. Soc. B 370, 20140323 ( 10.1098/rstb.2014.0323) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Gray MW, Burger G, Lang BF. 1999. Mitochondrial evolution. Science 283, 1476–1481. ( 10.1126/science.283.5407.1476) [DOI] [PubMed] [Google Scholar]
  • 57.Esser C, et al. 2004. A genome phylogeny for mitochondria among alpha-proteobacteria and a predominantly eubacterial ancestry of yeast nuclear genes. Mol. Biol. Evol. 21, 1643–1660. ( 10.1093/molbev/msh160) [DOI] [PubMed] [Google Scholar]
  • 58.Lang BF, Gray MW, Burger G. 1999. Mitochondrial genome evolution and the origin of eukaryotes. Annu. Rev. Genet. 33, 351–397. ( 10.1146/annurev.genet.33.1.351) [DOI] [PubMed] [Google Scholar]
  • 59.Burger G, Gray MW, Lang BF. 2003. Mitochondrial genomes: anything goes. Trends Genet. 19, 709–716. ( 10.1016/j.tig.2003.10.012) [DOI] [PubMed] [Google Scholar]
  • 60.Bullerwell CE, Gray MW. 2004. Evolution of the mitochondrial genome: protist connections to animals, fungi and plants. Curr. Opin. Microbiol. 7, 528–534. ( 10.1016/j.mib.2004.08.008) [DOI] [PubMed] [Google Scholar]
  • 61.Hackstein JH, Tjaden J, Huynen M. 2006. Mitochondria, hydrogenosomes and mitosomes: products of evolutionary tinkering! Curr. Genet. 50, 225–245. ( 10.1007/s00294-006-0088-8) [DOI] [PubMed] [Google Scholar]
  • 62.Keeling PJ. 2010. The endosymbiotic origin, diversification and fate of plastids. Phil. Trans. R. Soc. B 365, 729–748. ( 10.1098/rstb.2009.0103) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Hoppenrath M, Bachvaroff TR, Handy SM, Delwiche CF, Leander BS. 2009. Molecular phylogeny of ocelloid-bearing dinoflagellates (Warnowiaceae) as inferred from SSU and LSU rDNA sequences. BMC Evol. Biol. 9, 116 ( 10.1186/1471-2148-9-116) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Gavelis GS, Hayakawa S, White Iii RA, Gojobori T, Suttle CA, Keeling PJ, Leander BS. 2015. Eye-like ocelloids are built from different endosymbiotically acquired components. Nature 523, 204–207. ( 10.1038/nature14593) [DOI] [PubMed] [Google Scholar]
  • 65.Stebbins JGL. 1951. Cataclysmic evolution. Sci. Am. 184, 54–59. ( 10.1038/scientificamerican0451-54) [DOI] [Google Scholar]
  • 66.Anderson E, Stebbins GL Jr. 1954. Hybridization as an evolutionary stimulus. Evolution 8, 378–388. ( 10.1111/j.1558-5646.1954.tb01504.x) [DOI] [Google Scholar]
  • 67.Hegarty M, Hiscock S. 2007. Polyploidy: doubling up for evolutionary success. Curr. Biol. 17, R927–R929. ( 10.1016/j.cub.2007.08.060) [DOI] [PubMed] [Google Scholar]
  • 68.Soltis DE. 2009. Polyploidy and angiosperm diversification. Am. J. Bot. 96, 336–348. ( 10.3732/ajb.0800079) [DOI] [PubMed] [Google Scholar]
  • 69.Doyle JJ, Flagel LE, Paterson AH, Rapp RA, Soltis DE, Soltis PS, Wendel JF. 2008. Evolutionary genetics of genome merger and doubling in plants. Annu. Rev. Genet. 42, 443–461. ( 10.1146/annurev.genet.42.110807.091524) [DOI] [PubMed] [Google Scholar]
  • 70.Albertin W, Marullo P. 2012. Polyploidy in fungi: evolution after whole-genome duplication. Proc. R. Soc. B 279, 2497–2509. ( 10.1098/rspb.2012.0434) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Greig D, Louis EJ, Borts RH, Travisano M. 2002. Hybrid speciation in experimental populations of yeast. Science 298, 1773–1775. ( 10.1126/science.1076374) [DOI] [PubMed] [Google Scholar]
  • 72.Stelkens RB, Brockhurst MA, Hurst GD, Greig D. 2014. Hybridization facilitates evolutionary rescue. Evol. Appl. 7, 1209–1217. ( 10.1111/eva.12214) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Muller LA, McCusker JH. 2009. A multispecies-based taxonomic microarray reveals interspecies hybridization and introgression in Saccharomyces cerevisiae. FEMS Yeast Res. 9, 143–152. ( 10.1111/j.1567-1364.2008.00464.x) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Sipiczki M. 2008. Interspecies hybridization and recombination in Saccharomyces wine yeasts. FEMS Yeast Res. 8, 996–1007. ( 10.1111/j.1567-1364.2008.00369.x) [DOI] [PubMed] [Google Scholar]
  • 75.Wolfe KH. 2015. Origin of the yeast whole-genome duplication. PLoS Biol. 13, e1002221 ( 10.1371/journal.pbio.1002221) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Marcet-Houben M, Gabaldon T. 2015. Beyond the whole-genome duplication: phylogenetic evidence for an ancient interspecies hybridization in the baker's yeast lineage. PLoS Biol. 13, e1002220 ( 10.1371/journal.pbio.1002220) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.da Silva T, et al. 2015. Hybridization within Saccharomyces genus results in homoeostasis and phenotypic novelty in winemaking conditions. PLoS ONE 10, e0123834 ( 10.1371/journal.pone.0123834) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Soltis PS, Soltis DE. 2009. The role of hybridization in plant speciation. Annu. Rev. Plant Biol. 60, 561–588. ( 10.1146/annurev.arplant.043008.092039) [DOI] [PubMed] [Google Scholar]
  • 79.Hegarty M, Coate J, Sherman-Broyles S, Abbott R, Hiscock S, Doyle J. 2013. Lessons from natural and artificial polyploids in higher plants. Cytogenet. Genome Res. 140, 204–225. ( 10.1159/000353361) [DOI] [PubMed] [Google Scholar]
  • 80.Cui L. 2006. Widespread genome duplications throughout the history of flowering plants. Genome Res. 16, 738–749. ( 10.1101/gr.4825606) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.De Bodt S, Maere S, Van de Peer Y. 2005. Genome duplication and the origin of angiosperms. Trends Ecol. Evol. 20, 591–597. ( 10.1016/j.tree.2005.07.008) [DOI] [PubMed] [Google Scholar]
  • 82.Lim KY, et al. 2008. Rapid chromosome evolution in recently formed polyploids in Tragopogon (Asteraceae). PLoS ONE 3, e3353 ( 10.1371/journal.pone.0003353) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Symonds VV, Soltis PS, Soltis DE. 2010. Dynamics of polyploid formation in Tragopogon (Asteraceae): recurrent formation, gene flow, and population structure. Evolution 64, 1984–2003. ( 10.1111/j.1558-5646.2010.00978.x) [DOI] [PubMed] [Google Scholar]
  • 84.Arnold ML, Buckner CM, Robinson JJ. 1991. Pollen-mediated introgression and hybrid speciation in Louisiana irises. Proc. Natl Acad. Sci. USA 88, 1398–1402. ( 10.1073/pnas.88.4.1398) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Arnold ML, Tang S, Knapp SJ, Martin NH. 2010. Asymmetric introgressive hybridization among louisiana iris species. Genes (Basel) 1, 9–22. ( 10.3390/genes1010009) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hamlin JA, Arnold ML. 2014. Determining population structure and hybridization for two iris species. Ecol. Evol. 4, 743–755. ( 10.1002/ece3.964) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.McCarthy EW, et al. 2015. The effect of polyploidy and hybridization on the evolution of floral colour in Nicotiana (Solanaceae). Ann. Bot. 115, 1117–1131. ( 10.1093/aob/mcv048) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Kelly LJ, Leitch AR, Clarkson JJ, Hunter RB, Knapp S, Chase MW. 2010. Intragenic recombination events and evidence for hybrid speciation in Nicotiana (Solanaceae). Mol. Biol. Evol. 27, 781–799. ( 10.1093/molbev/msp267) [DOI] [PubMed] [Google Scholar]
  • 89.Fuentes I, Stegemann S, Golczyk H, Karcher D, Bock R. 2014. Horizontal genome transfer as an asexual path to the formation of new species. Nature 511, 232–235. ( 10.1038/nature13291) [DOI] [PubMed] [Google Scholar]
  • 90.Albertin W, Alix K, Balliau T, Brabant P, Davanture M, Malosse C, Valot B, Thiellement H. 2007. Differential regulation of gene products in newly synthesized Brassica napus allotetraploids is not related to protein function nor subcellular localization. BMC Genomics 8, 56 ( 10.1186/1471-2164-8-56) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Chalhoub B, et al. 2014. Plant genetics. Early allopolyploid evolution in the post-Neolithic Brassica napus oilseed genome. Science 345, 950–953. ( 10.1126/science.1253435) [DOI] [PubMed] [Google Scholar]
  • 92.Albertin W, Balliau T, Brabant P, Chevre AM, Eber F, Malosse C, Thiellement H. 2006. Numerous and rapid nonstochastic modifications of gene products in newly synthesized Brassica napus allotetraploids. Genetics 173, 1101–1113. ( 10.1534/genetics.106.057554) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Casneuf T, De Bodt S, Raes J, Maere S, Van de Peer Y. 2006. Nonrandom divergence of gene expression following gene and genome duplications in the flowering plant Arabidopsis thaliana. Genome Biol. 7, R13 ( 10.1186/gb-2006-7-2-r13) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Schmickl R, Paule J, Klein J, Marhold K, Koch MA. 2012. The evolutionary history of the Arabidopsis arenosa complex: diverse tetraploids mask the Western Carpathian center of species and genetic diversity. PLoS ONE 7, e42691 ( 10.1371/journal.pone.0042691) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Schmickl R, Koch MA. 2011. Arabidopsis hybrid speciation processes. Proc. Natl Acad. Sci. USA 108, 14 192–14 197. ( 10.1073/pnas.1104212108) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Marfil CF, Camadro EL, Masuelli RW. 2009. Phenotypic instability and epigenetic variability in a diploid potato of hybrid origin, Solanum ruiz-lealii. BMC Plant Biol. 9, 21 ( 10.1186/1471-2229-9-21) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Pendinen G, Spooner DM, Jiang J, Gavrilenko T. 2012. Genomic in situ hybridization reveals both auto- and allopolyploid origins of different North and Central American hexaploid potato (Solanum sect. Petota) species. Genome 55, 407–415. ( 10.1139/g2012-027) [DOI] [PubMed] [Google Scholar]
  • 98.Pendinen G, Gavrilenko T, Jiang J, Spooner DM. 2008. Allopolyploid speciation of the Mexican tetraploid potato species Solanum stoloniferum and S. hjertingii revealed by genomic in situ hybridization. Genome 51, 714–720. ( 10.1139/G08-052) [DOI] [PubMed] [Google Scholar]
  • 99.Ozkan H, Levy AA, Feldman M. 2001. Allopolyploidy-induced rapid genome evolution in the wheat (Aegilops-Triticum) group. Plant Cell 13, 1735–1747. ( 10.1105/tpc.13.8.1735) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Ozkan H, Tuna M, Arumuganathan K. 2003. Nonadditive changes in genome size during allopolyploidization in the wheat (Aegilops–Triticum) group. J. Hered. 94, 260–264. ( 10.1093/jhered/esg053) [DOI] [PubMed] [Google Scholar]
  • 101.Flagel L, Udall J, Nettleton D, Wendel J. 2008. Duplicate gene expression in allopolyploid Gossypium reveals two temporally distinct phases of expression evolution. BMC Biol. 6, 16 ( 10.1186/1741-7007-6-16) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Flagel LE, Wendel JF. 2010. Evolutionary rate variation, genomic dominance and duplicate gene expression evolution during allotetraploid cotton speciation. New Phytol. 186, 184–193. ( 10.1111/j.1469-8137.2009.03107.x) [DOI] [PubMed] [Google Scholar]
  • 103.Dowling TS. 1997. The role of hybridization and introgression in the diversification of animals. Annu. Rev. Ecol. Syst. 28, 593–619. ( 10.1146/annurev.ecolsys.28.1.593) [DOI] [Google Scholar]
  • 104.Schwenk K, Brede N, Streit B. 2008. Introduction. Extent, processes and evolutionary impact of interspecific hybridization in animals. Phil. Trans. R. Soc. B 363, 2805–2811. ( 10.1098/rstb.2008.0055) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Mallet J. 2008. Hybridization, ecological races and the nature of species: empirical evidence for the ease of speciation. Phil. Trans. R. Soc. B 363, 2971–2986. ( 10.1098/rstb.2008.0081) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Schwarz D, Matta BM, Shakir-Botteri NL, McPheron BA. 2005. Host shift to an invasive plant triggers rapid animal hybrid speciation. Nature 436, 546–549. ( 10.1038/nature03800) [DOI] [PubMed] [Google Scholar]
  • 107.Schwarz D, Shoemaker KD, Botteri NL, McPheron BA. 2007. A novel preference for an invasive plant as a mechanism for animal hybrid speciation. Evolution 61, 245–256. ( 10.1111/j.1558-5646.2007.00027.x) [DOI] [PubMed] [Google Scholar]
  • 108.Mallet J, Beltran M, Neukirchen W, Linares M. 2007. Natural hybridization in heliconiine butterflies: the species boundary as a continuum. BMC Evol. Biol. 7, 28 ( 10.1186/1471-2148-7-28) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.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 Genet. 8, e1002752 ( 10.1371/journal.pgen.1002752) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Heliconius GC. 2012. Butterfly genome reveals promiscuous exchange of mimicry adaptations among species. Nature 487, 94–98. ( 10.1038/nature11041) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Nadeau NJ, et al. 2012. Genomic islands of divergence in hybridizing Heliconius butterflies identified by large-scale targeted sequencing. Phil. Trans. R. Soc. B 367, 343–353. ( 10.1098/rstb.2011.0198) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Zhang W, Dasmahapatra KK, Mallet J, Moreira GR, Kronforst MR. 2016. Genome-wide introgression among distantly related Heliconius butterfly species. Genome Biol. 17, 25 ( 10.1186/s13059-016-0889-0) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Stelbrink B, Stoger I, Hadiaty RK, Schliewen UK, Herder F. 2014. Age estimates for an adaptive lake fish radiation, its mitochondrial introgression, and an unexpected sister group: sailfin silversides of the Malili Lakes system in Sulawesi. BMC Evol. Biol. 14, 94 ( 10.1186/1471-2148-14-94) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Herder F, Nolte AW, Pfaender J, Schwarzer J, Hadiaty RK, Schliewen UK. 2006. Adaptive radiation and hybridization in Wallace's Dreamponds: evidence from sailfin silversides in the Malili Lakes of Sulawesi. Proc. R. Soc. B 273, 2209–2217. ( 10.1098/rspb.2006.3558) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Keller I, Wagner CE, Greuter L, Mwaiko S, Selz OM, Sivasundar A, Wittwer S, Seehausen O. 2013. Population genomic signatures of divergent adaptation, gene flow and hybrid speciation in the rapid radiation of Lake Victoria cichlid fishes. Mol. Ecol. 22, 2848–2863. ( 10.1111/mec.12083) [DOI] [PubMed] [Google Scholar]
  • 116.Brawand D, et al. 2014. The genomic substrate for adaptive radiation in African cichlid fish. Nature 513, 375–381. ( 10.1038/nature13726) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Selz OM, Lucek K, Young KA, Seehausen O. 2014. Relaxed trait covariance in interspecific cichlid hybrids predicts morphological diversity in adaptive radiations. J. Evol. Biol. 27, 11–24. ( 10.1111/jeb.12283) [DOI] [PubMed] [Google Scholar]
  • 118.Selz OM, Thommen R, Maan ME, Seehausen O. 2014. Behavioural isolation may facilitate homoploid hybrid speciation in cichlid fish. J. Evol. Biol. 27, 275–289. ( 10.1111/jeb.12287) [DOI] [PubMed] [Google Scholar]
  • 119.Svensson O, Smith A, Garcia-Alonso J, van Oosterhout C. 2016. Hybridization generates a hopeful monster: a hermaphroditic selfing cichlid. R. Soc. Open Sci. 3, 150684 ( 10.1098/rsos.150684) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Hulsey CD, Garcia-de-Leon FJ. 2013. Introgressive hybridization in a trophically polymorphic cichlid. Ecol. Evol. 3, 4536–4547. ( 10.1002/ece3.841) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Smith PF, Konings A, Kornfield I. 2003. Hybrid origin of a cichlid population in Lake Malawi: implications for genetic variation and species diversity. Mol. Ecol. 12, 2497–2504. ( 10.1046/j.1365-294X.2003.01905.x) [DOI] [PubMed] [Google Scholar]
  • 122.Hermansen JS, Saether SA, Elgvin TO, Borge T, Hjelle E, Saetre GP. 2011. Hybrid speciation in sparrows I: phenotypic intermediacy, genetic admixture and barriers to gene flow. Mol. Ecol. 20, 3812–3822. ( 10.1111/j.1365-294X.2011.05183.x) [DOI] [PubMed] [Google Scholar]
  • 123.Eroukhmanoff F, Hermansen JS, Bailey RI, Saether SA, Saetre GP. 2013. Local adaptation within a hybrid species. Heredity (Edinb) 111, 286–292. ( 10.1038/hdy.2013.47) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Grant PR, Grant BR, Petren K. 2005. Hybridization in the recent past. Am. Nat. 166, 56–67. ( 10.1086/430331) [DOI] [PubMed] [Google Scholar]
  • 125.Grant P, Grant BR. 1992. Hybridization of bird species. Science 256, 193–197. ( 10.1126/science.256.5054.193) [DOI] [PubMed] [Google Scholar]
  • 126.Grant PR, Grant BR, Markert JA, Keller LF, Petren K. 2004. Convergent evolution of Darwin's finches caused by introgressive hybridization and selection. Evolution 58, 1588–1599. ( 10.1111/j.0014-3820.2004.tb01738.x) [DOI] [PubMed] [Google Scholar]
  • 127.Grant BR, Grant PR. 2008. Fission and fusion of Darwin's finches populations. Phil. Trans. R. Soc. B 363, 2821–2829. ( 10.1098/rstb.2008.0051) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Grant PR, Grant BR. 2014. 40 years of evolution. Darwin’s finches on Daphne Major Island. Princeton, NJ: Princeton University Press. [Google Scholar]
  • 129.Lamichhaney S, et al. 2015. Evolution of Darwin's finches and their beaks revealed by genome sequencing. Nature 518, 371–375. ( 10.1038/nature14181) [DOI] [PubMed] [Google Scholar]
  • 130.Palmer DH, Kronforst MR. 2015. Divergence and gene flow among Darwin's finches: a genome-wide view of adaptive radiation driven by interspecies allele sharing. Bioessays 37, 968–974. ( 10.1002/bies.201500047) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Amaral AR, Jackson JA, Moller LM, Beheregaray LB, Manuela Coelho M. 2012. Species tree of a recent radiation: the subfamily Delphininae (Cetacea, Mammalia). Mol. Phylogenet. Evol. 64, 243–253. ( 10.1016/j.ympev.2012.04.004) [DOI] [PubMed] [Google Scholar]
  • 132.Amaral AR, Lovewell G, Coelho MM, Amato G, Rosenbaum HC. 2014. Hybrid speciation in a marine mammal: the clymene dolphin (Stenella clymene). PLoS ONE 9, e83645 ( 10.1371/journal.pone.0083645) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Arnold ML. 2004. Transfer and origin of adaptations through natural hybridization: were Anderson and Stebbins right? Plant Cell 16, 562–570. ( 10.1105/tpc.160370) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Abbott R, et al. 2013. Hybridization and speciation. J. Evol. Biol. 26, 229–246. ( 10.1111/j.1420-9101.2012.02599.x) [DOI] [PubMed] [Google Scholar]
  • 135.Mallet J. 2007. Hybrid speciation. Nature 446, 279–283. ( 10.1038/nature05706) [DOI] [PubMed] [Google Scholar]
  • 136.Payseur BA, Rieseberg LH. 2016. A genomic perspective on hybridization and speciation. Mol. Ecol. 25, 2337–2360. ( 10.1111/mec.13557) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Seehausen O. 2015. Process and pattern in cichlid radiations—inferences for understanding unusually high rates of evolutionary diversification. New Phytol. 207, 304–312. ( 10.1111/nph.13450) [DOI] [PubMed] [Google Scholar]
  • 138.Hegarty MJ, Barker GL, Brennan AC, Edwards KJ, Abbott RJ, Hiscock SJ. 2008. Changes to gene expression associated with hybrid speciation in plants: further insights from transcriptomic studies in Senecio. Phil. Trans. R. Soc. B 363, 3055–3069. ( 10.1098/rstb.2008.0080) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Greaves IK, Groszmann M, Ying H, Taylor JM, Peacock WJ, Dennis ES. 2012. Trans chromosomal methylation in Arabidopsis hybrids. Proc. Natl Acad. Sci. USA 109, 3570–3575. ( 10.1073/pnas.1201043109) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Parisod C, Salmon A, Zerjal T, Tenaillon M, Grandbastien MA, Ainouche M. 2009. Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytol. 184, 1003–1015. ( 10.1111/j.1469-8137.2009.03029.x) [DOI] [PubMed] [Google Scholar]
  • 141.Lim JK, Simmons MJ. 1994. Gross chromosome rearrangements mediated by transposable elements in Drosophila melanogaster. Bioessays 16, 269–275. ( 10.1002/bies.950160410) [DOI] [PubMed] [Google Scholar]
  • 142.Mieczkowski PA, Lemoine FJ, Petes TD. 2006. Recombination between retrotransposons as a source of chromosome rearrangements in the yeast Saccharomyces cerevisiae. DNA Repair (Amst) 5, 1010–1020. ( 10.1016/j.dnarep.2006.05.027) [DOI] [PubMed] [Google Scholar]
  • 143.Lonnig WE, Saedler H. 2002. Chromosome rearrangements and transposable elements. Annu. Rev. Genet. 36, 389–410. ( 10.1146/annurev.genet.36.040202.092802) [DOI] [PubMed] [Google Scholar]
  • 144.Zhang J, Yu C, Krishnaswamy L, Peterson T. 2011. Transposable elements as catalysts for chromosome rearrangements. Methods Mol. Biol. 701, 315–326. ( 10.1007/978-1-61737-957-4_18) [DOI] [PubMed] [Google Scholar]
  • 145.Carbone L, et al. 2014. Gibbon genome and the fast karyotype evolution of small apes. Nature 513, 195–201. ( 10.1038/nature13679) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Senerchia N, Felber F, Parisod C. 2015. Genome reorganization in F1 hybrids uncovers the role of retrotransposons in reproductive isolation. Proc. R. Soc. B 282, 20142874 ( 10.1098/rspb.2014.2874) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Metcalfe CJ, Bulazel KV, Ferreri GC, Schroeder-Reiter E, Wanner G, Rens W, Obergfell C, Eldridge MD, O'Neill RJ. 2007. Genomic instability within centromeres of interspecific marsupial hybrids. Genetics 177, 2507–2517. ( 10.1534/genetics.107.082313) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Marfil CF, Masuelli RW, Davison J, Comai L. 2006. Genomic instability in Solanum tuberosum×Solanum kurtzianum interspecific hybrids. Genome 49, 104–113. [DOI] [PubMed] [Google Scholar]
  • 149.Han FP, Fedak G, Ouellet T, Liu B. 2003. Rapid genomic changes in interspecific and intergeneric hybrids and allopolyploids of Triticeae. Genome 46, 716–723. ( 10.1139/g03-049) [DOI] [PubMed] [Google Scholar]
  • 150.King M. 1995. Species evolution: the role of chromosome change. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 151.White MJ. 1949. Chromosomes of the vertebrates. Evolution 3, 379–381. ( 10.1111/j.1558-5646.1949.tb00043.x) [DOI] [PubMed] [Google Scholar]
  • 152.Nie W, Wang J, Su W, Wang D, Tanomtong A, Perelman PL, Graphodatsky AS, Yang F. 2012. Chromosomal rearrangements and karyotype evolution in carnivores revealed by chromosome painting. Heredity (Edinb) 108, 17–27. ( 10.1038/hdy.2011.107) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Craig N, Craigie R, Gellert M, Lambowitz AM. 2002. Mobile DNA II. Washington, DC: American Society for Microbiology Press. [Google Scholar]
  • 154.Lander ES, et al. 2001. Initial sequencing and analysis of the human genome. Nature 409, 860–921. ( 10.1038/35057062) [DOI] [PubMed] [Google Scholar]
  • 155.de Koning AP, Gu W, Castoe TA, Batzer MA, Pollock DD. 2011. Repetitive elements may comprise over two-thirds of the human genome. PLoS Genet. 7, e1002384 ( 10.1371/journal.pgen.1002384) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Liu G, Mattick JS, Taft RJ. 2013. A meta-analysis of the genomic and transcriptomic composition of complex life. Cell Cycle 12, 2061–2072. ( 10.4161/cc.25134) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Shapiro JA. 2016. Exploring the read–write genome: mobile DNA and mammalian adaptation. Crit. Rev. Biochem. Mol. Biol. 52, 1–17. ( 10.1080/10409238.2016.1226748) [DOI] [PubMed] [Google Scholar]
  • 158.Shapiro JA, Sternberg RV. 2005. Why repetitive DNA is essential to genome function. Biol. Rev. (Camb.). 80, 227–250. ( 10.1017/S1464793104006657) [DOI] [PubMed] [Google Scholar]
  • 159.Bartholdy B, Mukhopadhyay R, Lajugie J, Aladjem MI, Bouhassira EE. 2015. Allele-specific analysis of DNA replication origins in mammalian cells. Nat. Commun. 6, 7051 ( 10.1038/ncomms8051) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Cournac A, Koszul R, Mozziconacci J. 2016. The 3D folding of metazoan genomes correlates with the association of similar repetitive elements. Nucleic Acids Res. 44, 245–255. ( 10.1093/nar/gkv1292) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Elliott TA, Linquist S, Gregory TR. 2014. Conceptual and empirical challenges of ascribing functions to transposable elements. Am. Nat. 184, 14–24. ( 10.1086/676588) [DOI] [PubMed] [Google Scholar]
  • 162.Doolittle WF. 2013. Is junk DNA bunk? A critique of ENCODE. Proc. Natl Acad. Sci. USA 110, 5294–5300. ( 10.1073/pnas.1221376110) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Macaulay EC, Weeks RJ, Andrews S, Morison IM. 2011. Hypomethylation of functional retrotransposon-derived genes in the human placenta. Mamm. Genome. 22, 722–735. ( 10.1007/s00335-011-9355-1) [DOI] [PubMed] [Google Scholar]
  • 164.Emera D, Wagner GP. 2012. Transposable element recruitments in the mammalian placenta: impacts and mechanisms. Brief. Funct. Genomics 11, 267–276. ( 10.1093/bfgp/els013) [DOI] [PubMed] [Google Scholar]
  • 165.Chuong EB. 2013. Retroviruses facilitate the rapid evolution of the mammalian placenta. Bioessays 35, 853–861. ( 10.1002/bies.201300059) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Lynch VJ, et al. 2015. Ancient transposable elements transformed the uterine regulatory landscape and transcriptome during the evolution of mammalian pregnancy. Cell Rep. 10, 551–561. ( 10.1016/j.celrep.2014.12.052) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Kapusta A, Kronenberg Z, Lynch VJ, Zhuo X, Ramsay L, Bourque G, Yandell M, Feschotte C. 2013. Transposable elements are major contributors to the origin, diversification, and regulation of vertebrate long noncoding RNAs. PLoS Genet. 9, e1003470 ( 10.1371/journal.pgen.1003470) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Kapusta A, Feschotte C. 2014. Volatile evolution of long noncoding RNA repertoires: mechanisms and biological implications. Trends Genet. 30, 439–452. ( 10.1016/j.tig.2014.08.004) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Johnson R, Guigo R. 2014. The RIDL hypothesis: transposable elements as functional domains of long noncoding RNAs. RNA 20, 959–976. ( 10.1261/rna.044560.114) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Peschansky VJ, Wahlestedt C. 2014. Non-coding RNAs as direct and indirect modulators of epigenetic regulation. Epigenetics 9, 3–12. ( 10.4161/epi.27473) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Guttman M, et al. 2011. lincRNAs act in the circuitry controlling pluripotency and differentiation. Nature 477, 295–300. ( 10.1038/nature10398) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Kelley D, Rinn J. 2012. Transposable elements reveal a stem cell-specific class of long noncoding RNAs. Genome Biol. 13, R107 ( 10.1186/gb-2012-13-11-r107) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Wang Y, et al. 2013. Endogenous miRNA sponge lincRNA-RoR regulates Oct4, Nanog, and Sox2 in human embryonic stem cell self-renewal. Dev. Cell 25, 69–80. ( 10.1016/j.devcel.2013.03.002) [DOI] [PubMed] [Google Scholar]
  • 174.Durruthy-Durruthy J, et al. 2016. The primate-specific noncoding RNA HPAT5 regulates pluripotency during human preimplantation development and nuclear reprogramming. Nat. Genet. 48, 44–52. ( 10.1038/ng.3449) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Sauvageau M, et al. 2013. Multiple knockout mouse models reveal lincRNAs are required for life and brain development. Elife 2, e01749 ( 10.7554/eLife.01749) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Lin N, et al. 2014. An evolutionarily conserved long noncoding RNA TUNA controls pluripotency and neural lineage commitment. Mol. Cell 53, 1005–1019. ( 10.1016/j.molcel.2014.01.021) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Aprea J, Calegari F. 2015. Long non-coding RNAs in corticogenesis: deciphering the non-coding code of the brain. EMBO J. 34, 2865–2884. ( 10.15252/embj.201592655) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Aune TM, Spurlock CF 3rd. 2016. Long non-coding RNAs in innate and adaptive immunity. Virus Res. 212, 146–160. ( 10.1016/j.virusres.2015.07.003) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Chuong EB, Elde NC, Feschotte C. 2016. Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science 351, 1083–1087. ( 10.1126/science.aad5497) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.McClintock B. 1952. Controlling elements and the gene. Cold Spring Harb. Symp. Quant. Biol. 21, 197–216. ( 10.1101/SQB.1956.021.01.017) [DOI] [PubMed] [Google Scholar]
  • 181.Pilacinski W, Mosharrafa E, Edmundson R, Zissler J, Fiandt M, Szybalski W. 1977. Insertion sequence IS2 associated with int-constitutive mutants of bacteriophage lambda. Gene 2, 61–74. ( 10.1016/0378-1119(77)90073-7) [DOI] [PubMed] [Google Scholar]
  • 182.Fiandt M, Szybalski W, Ahmed A. 1977. Identification of the gal3 insertion in Escherichia coli AS IS2. Gene 2, 55–58. ( 10.1016/0378-1119(77)90021-X) [DOI] [PubMed] [Google Scholar]
  • 183.Errede B, Cardillo TS, Wever G, Sherman F, Stiles JI, Friedman LR, Sherman F. 1981. Studies on transposable elements in yeast. I. ROAM mutations causing increased expression of yeast genes: their activation by signals directed toward conjugation functions and their formation by insertion of Ty1 repetitive elements. II. Deletions, duplications, and transpositions of the COR segment that encompasses the structural gene of yeast iso-1-cytochrome c. Cold Spring Harb. Symp. Quant. Biol. 45, 593–607. ( 10.1101/SQB.1981.045.01.077) [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Table 1. Various stimuli documented to activate natural genetic engineering
rsfs20160115supp1.pdf (186.9KB, pdf)
Supplementary Table 2. Examples of horizontal DNA transfer
rsfs20160115supp2.pdf (281.7KB, pdf)
Supplementary Table 3. Genomic responses to changes in ploidy and interspecific hybridization in fungi, plants and animals.pdf
rsfs20160115supp3.pdf (124.5KB, pdf)

Data Availability Statement

This article has no additional data.


Articles from Interface Focus are provided here courtesy of The Royal Society

RESOURCES