Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2556 word(s) 2556 2021-12-30 04:17:25 |
2 Done Meta information modification 2556 2022-01-06 07:25:46 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Capy, P. Taming, Domestication and Exaptation. Encyclopedia. Available online: https://encyclopedia.pub/entry/17779 (accessed on 11 September 2024).
Capy P. Taming, Domestication and Exaptation. Encyclopedia. Available at: https://encyclopedia.pub/entry/17779. Accessed September 11, 2024.
Capy, Pierre. "Taming, Domestication and Exaptation" Encyclopedia, https://encyclopedia.pub/entry/17779 (accessed September 11, 2024).
Capy, P. (2022, January 05). Taming, Domestication and Exaptation. In Encyclopedia. https://encyclopedia.pub/entry/17779
Capy, Pierre. "Taming, Domestication and Exaptation." Encyclopedia. Web. 05 January, 2022.
Taming, Domestication and Exaptation
Edit

During evolution, several types of sequences pass through genomes. Along with mutations and internal genetic tinkering, they are a useful source of genetic variability for adaptation and evolution. Most of these sequences are acquired by horizontal transfers (HT), but some of them may come from the genomes themselves. If they are not lost or eliminated quickly, they can be tamed, domesticated, or even exapted.

transposable elements domestication exaptation taming

1. Introduction

Transposable elements (TE), frequently called “selfish genes[1], “selfish DNA[2], or junk or garbage DNA, according to the categories defined by Graur et al. [3], present several genetic characteristics that allow them to rapidly invade genomes and populations, as well as to sometime to settle there permanently. Generally, after their arrival in a naïve genome and an invasion phase, their overall activity decreases drastically, leading to the maintenance of very few autonomous copies. However, several non-autonomous or dead copies, or even pieces of TE, can be preserved with non-neutral effects on individual fitness, due to their particular insertion site or the acquisition of new characteristics after a more or less lengthy coevolution with genomes.
During this coevolution process between TE and genomes, various interactions and trajectories can lead to the emergence of relatively stable evolutionary states, usually described as taming, domestication, or exaptation. Although these different terms seem to be closely related, they cover different phenomena, as briefly described below.
Taming. This interaction tends to rapidly reduce and limit the negative fitness impact of an excessively high transposition rate of a new invading TE on both genome structure and function. This is not an irreversible phenomenon because, sometimes, it must be reset at each generation, especially if it is due to non-transgenerational epigenetic marks. Moreover, in stressful conditions, an element can escape and have an intensive transposition activity. This can be illustrated by the regulation of TE activity, with occasional wake-up and bursts [4][5][6][7][8][9]. In this respect, the epigenetic regulation of TE activity plays an important role, and a few autonomous and silenced copies present in the genome can be reactivated occasionally by biotic, abiotic, genomic, or demographic stress. At the populational level, this is crucial for creating new genetic variability to cope with stress and adapt to new environmental conditions.
Domestication. The general definition of domestication is: a sustainable interaction, maintained over generations, resulting from a hierarchical relationship, based on a directional transformation of one entity by another for its own benefit. This leads to deep modifications of genetic material of the domesticated entities, like acquisition, loss, or transformation of one or several traits. In a genomic context, TEs are the domesticated entities and genomes of the “hierarchical superior”. Moreover, while there is no emergence of a new function, they can have an impact on the genome’s functioning. Indeed, a copy, through its genomic insertion site, can impact individual fitness and rapidly invade and settle in the population if it provides an advantage. In this respect, work based on populational analyses has reported many examples [10][11][12][13].
Exaptation. This term, introduced by Gould and Vbra in 1982 [14], refers to the emergence of a new function that enhances the fitness of individuals. More precisely, it (in Table 1 of their publication) suggests two different processes: “1—character, previously shaped by natural selection for a particular function (an adaptation), is coopted for a new use-cooptation; 2—A character whose origin cannot be ascribed to the direct action of natural selection (a nonaptation), is coopted for a current use-cooptation”. It is, therefore, a sequential evolution of a trait that was initially shaped (or not) by natural selection to a trait today shaped by natural selection and adapted to a new function.
Numerous biological examples, at the morphological, physiological, and molecular levels, can illustrate such an evolutionary trajectory, such as the feathers of birds originally “designed” for thermoregulation and today exapted in flight. At the molecular level and, more particularly, in the TE world, several examples will be detailed below.
During evolution, genetic tinkering is a major source for the emergence of new regulation systems, genome reorganization, and new functions [15][16][17]. Within species, this tinkering may be due to the shuffling and association of different parts of a genome by ectopic recombination, transposition, gene duplication, frameshift mutation, translocation, or, again, autopolyploidy in plants. However, this dynamic can also be fueled by the acquisition of external genetic material, as a result, for example, of interspecific hybridizations or horizontal transfers (HT). Such phenomena are responsible for the emergence of genetic novelties, as, for instance, the acquisition of new genes, paralogs of existing genes, and xenologous gene displacement [18]. In addition, they can occur in distantly related species, from different kingdoms within eukaryotes, or even between prokaryotes and eukaryotes. Many example of adaptive horizontal transfers are reviewed by Crisp et al. [19]. According to these authors, 2% of the foreign genes of primates come from archaea, 25% from bacteria, 57.6% from protists, 5.4% from plants, and 10% from fungi.
Based on all genome analyses during the last decade, it has been evidenced that the exchange of genetic material between closely or distantly related species is probably much more frequent than previously assumed. Concerning TE, HT are possible both after an interspecific hybridization or between distantly related species. Nowadays, such transfers do not appear to be rare evolutionary events, and the number of descriptions or suspicions continues to increase [20][21][22][23]. For instance, in insects, Peccoud et al. [22] found that out of 195 genomes, 4500 HT can be detected.
More precisely, inter-specific hybridizations occur between closely related species, which can hybridize and are able to produce fertile offspring. In plants, such a phenomenon is frequent and leads to the emergence of allopolyploids [24]. This favors the addition of genetic material in both species and the introduction of new variants, which can become the raw material for new genetic tinkering. In animals, inter-specific hybridization can also be observed between species with sexual reproduction. In such a case, and according to Haldane’s rule, only the homogametic sex is fertile (for instance, XX females in the XY system and ZZ males in the ZW system). The fertile sex can then be backcrossed with individuals (males XY or females ZW) of one of the two parental species, leading to the transfer of genetic material of one species to the other (introgression).
On the other hand, horizontal transfers also occur between distantly related species when no sexual reproduction is possible. They were probably very frequent during the early steps of life [25] and were at the origin of important evolutionary steps, such as the exchanges between prokaryotes and eukaryotes or between bacteria/archaea and extremophilic eukaryotes [26][27]. This also occurs during the endosymbioses of proteobacteria and cyanobacteria, leading to the emergence of mitochondria and chloroplasts [28] or between prokaryotes (see for instance the numerous examples in Escudero et al. [29], or San Millan et al. [30]), where they frequently promote the exchange of resistance to environmental stress via conjugation, transduction, and transformation, whether or not they use TE as vectors [31][32].

2. Short-Term Co-Evolution of Transposable Elements and Genomes: Taming

While in prokaryotes, the HT mechanisms are known and responsible for rapid diffusion of resistance to environmental stresses [30], the transfer mechanisms remain unknown in eukaryotes, and several scenarios have been proposed [21][33][34]. However, it is likely that the arrival of a new TE in a eukaryotic genome probably occurs in most cases by horizontal transfer [21][35]. At this point in the TE life cycle, there is only one copy in a single individual. Therefore, the probability of losing this copy through genetic drift is very high. To maintain it and allow genome and populational invasion, the impact on fitness must be positive and very high or more likely, TEs have to adopt a parasitic strategy, i.e., a low phenotypic effect, with a relatively high transposition rate [36]. In addition, it seems that several TE, among which some members of the Tc1-mariner superfamily, such as Bari1, Bari3, and Sleeping Beauty would facilitate their genomic diffusion after a horizontal transfer, might have evolved as “blurry promoters” [37][38].
After this more or less lengthy invasion phase, a plateau is reached, during which the number of copies is stabilized. Few copies of this element will then remain autonomous, while the others will become non-autonomous but trans-mobilizable, with the remaining copies degenerating. In this context, it is interesting to observe that a competition can take place between the different types of copies from the same family (between autonomous vs. non-autonomous but trans-mobilizable copies), leading to a dynamic similar to that described by Lokta [39] and Volterra [40] for the prey-predator relationship in population biology [41][42].
This basic TE life cycle can be viewed as a parasitic strategy in the invaded genome. However, the golden rule of many parasitic entities is to be as “silent” as possible. In other words, to be maintained over long evolutionary periods, the TE copy number must be neither too low to avoid elimination by genetic drift or ectopic recombination nor too high to avoid a negative impact on individual fitness.
In this phase, TE silencing may be promoted by epigenetic regulation. The term “epigenetics” generally refers to several mechanisms, such as cytosine methylation in Arabidopsis thaliana, where most copies are methylated and inactivated [43]; small RNA (siRNA and piRNA), as described in different tissues in D. melanogaster (in germline to control I and P element transposition [44], testes and ovaries [45][46][47][48]), and somatic and germinal tissues of arthropods [49] (as a stress response in A. thaliana [50]), as well as long non-coding RNA in plants with differential expression in tissue and depending on environmental conditions [51]. While the epigenetic regulation seems to be dominant, other mechanisms of TE-silencing can be evoked, such as those involving a self-encoded repressor (such as the internally deleted KP element, derived from the P element), [52] or to splicing events, such as for the Bari1 element [53].
One of the evolutionary interests of such silencing is its reversibility. This has two main effects. First, when epigenetic marks are removed, a transposition burst can be observed [54][55] and second, genes located near the TE insertion site can also be reactivated because the methylated area may be larger than the TE itself and can encompass neighboring sequences [56][57][58][59]. Therefore, this type of reversible interaction between TE and genomes can be useful for the genome, insofar as it allows it to maintain a functional “genetic toolbox”, which can be reactivated when necessary to generate new genetic variability and evolve rapidly in a changing environment.

3. Long Term Co-Evolution of Transposable Elements and Genomes: Domestication and Exaptation

Two common characteristics are shared by the processes of domestication and exaptation. The first is the “capture” of a copy in a specific genomic location, and the second its maintenance, which can go as far as fixing itself in a population or a species. Regarding the genomic location, this raises the question of the distribution of TE copies in a genome. Is there a random distribution or a patchy distribution with hot insertion regions?
For more than 30 years, it has been observed that TE distribution is patchy [60]. On a coarse scale, this distribution can vary from one chromosome to another, but also within a chromosome, and again among the main TE Classes. For instance, in the human genome, the Alu distribution is not similar between chromosomes 21 and 22 [61], and L1 elements are not randomly distributed, although they seem able to target all genomic regions [62]. A similar distribution bias is also observed in Drosophila [63], catfish [64], and woodpeckers [65], among others. All these results suggest that even if TEs are potentially capable of jumping everywhere in the genomes, purifying selection against new insertion and ectopic recombination can remove several of them and reshape distribution [66][67]. However, the alternative hypothesis, assuming that TEs insert into peculiar regions, cannot be ruled out.
With the accumulation of complete genome sequences and the new molecular tools recently developed to explore them, it is now clear that this distribution is patchy. In addition to the evolutionary forces previously mentioned, new parameters must be taken into account, such as the status (condensation vs. decondensation) of chromatin [68] or “DNA sequence, chromatin and nuclear context and cellular proteins” because they are also involved in TE integration [69], showing that peculiar genomic territories are more prone to TE insertions than others.
More precisely, several results show that regions with a specific chromatin structure seem to be more “attractive”, such as the regulatory regions of genes or heterochromatin, whether they are centromeric, telomeric or interspersed in euchromatin [70][71][72][73][74]. Insertions of TEs in gene-rich regions have also been frequently described in numerous species, such as Drosophila for retrotransposons [75], for the retroposon Accord in 5′ of a gene involved in resistance to insecticides [76][77][78], and, more recently, for diverse TE families, frequently associated with stress-related genes [79]. Similar observations have been reported in mice [80] and wheat [81]. Moreover, the existence of nested accumulation of TEs in euchromatin [82], useful for TE “paleontology” [83], must also be considered. Especially, since they could be at the origin of Pi clusters, involved in regulation of TE activity by small RNA [48][84][85][86][87].
Some regions are the main targets of TEs, probably because of their accessibility [88][89][90]. In addition, patchy distribution due to the accessibility effect could be reinforced by the existence of low recombination rates, leading us to consider some of these regions as TE graveyards [91][92][93][94].
Therefore, patchy TE distribution is the result of multiple factors, and two steps must be considered: first, an insertion phase with random or non-random insertions, and second, a differential elimination or maintenance phase, due to selection against deleterious insertions, positive selection on insertion with beneficial host impact and elimination in regions with a high recombination rate.
In this review, I will differentiate domestication and exaptation. Can an insertion close to a gene and modifying its expression profile not be considered as an exaptation? Although such insertions have an impact on the host genome, as illustrated by many examples such as Mendel’s wrinkled pea [95], the industrial melanism of Biston betularia [96], the resistance to insecticides [77][97][98] or to xenobiotics [78] in D. melanogaster. Their frequency may increase in natural populations more or less rapidly, depending on their effect on host fitness [99] and/or the genetic drift, due to the effective population size (Ne). Domestication applies to a whole TE copy or a part of it, and frequently a copy is completely domesticated as soon as its mobility and its capacity to encode a functional transposition machinery is lost. Whatever the situation, these copies have an impact on the expression profile of the surrounding genes, but they are not initially the source of new functions or new genes. However, domestication can be a step towards exaptation.
On the other hand, in an exaptation process, all or part of the sequence of a copy is fixed in the population or species. This is the source of new functions and sometimes new genes, which significantly increase host fitness. Such novelties are present in a single species when exaptation is recent or in a group of phylogenetically related species for an older exaptation that occurred before the speciation events. Several examples detailed later will illustrate such a phenomenon, such as the telomeric element in arthropods [100], the vertebrate immune system [101], or placenta development in mammals [102].
Domestication and exaptation can be detected from analysis of the evolution of polymorphism along the chromosome by the existence of regions with low variability due to the effect of selective sweep or background selection. As recently suggested in very interesting articles [103][104], these phenomena require several successive stages. Here, I would just like to summarize this process and add several considerations.

References

  1. Dawkins, R. The Selfish Gene; Oxford University Press: Oxford, UK, 1976.
  2. Doolittle Sapienza, C. Selfish genes, the phenotype paradigm and genome evolution. Nature 1980, 284, 601–603.
  3. Graur, D.; Zheng, Y.; Azevedo, R.B.R. An Evolutionary Classification of Genomic Function. Genome Biol. Evol. 2015, 7, 642–645.
  4. Boissinot, S.; Bourgeois, Y.; Manthey, J.D.; Ruggiero, R.P. The Mobilome of Reptiles: Evolution, Structure, and Function. Cytogenet. Genome Res. 2019, 157, 21–33.
  5. Carotti, E.; Carducci, F.; Canapa, A.; Barucca, M.; Greco, S.; Gerdol, M.; Biscotti, M.A. Transposable Elements and Teleost Migratory Behaviour. Int. J. Mol. Sci. 2021, 22, 602.
  6. El Baidouri, M.; Panaud, O. Comparative Genomic Paleontology across Plant Kingdom Reveals the Dynamics of TE-Driven Genome Evolution. Genome Biol. Evol. 2013, 5, 954–965.
  7. Giraud, D.; Lima, O.; Huteau, V.; Coriton, O.; Boutte, J.; Kovarik, A.; Leitch, A.R.; Leitch, I.J.; Aïnouche, M.; Salmon, A. Evolutionary dynamics of transposable elements and satellite DNAs in polyploid Spartina species. Plant Sci. 2021, 302, 110671.
  8. Horvath, V.; Merenciano, M.; Gonzalez, J. Revisiting the Relationship between Transposable Elements and the Eukaryotic Stress Response. Trends Genet. 2017, 33, 832–841.
  9. Oliveira, D.S.; Rosa, M.T.; Vieira, C.; Loreto, E.L.S. Oxidative and radiation stress induces transposable element transcription in Drosophila melanogaster. J. Evol. Biol. 2021, 34, 628–638.
  10. Gonzalez, J.; Macpherson, J.M.; Petrov, D.A. A Recent Adaptive Transposable Element Insertion Near Highly Conserved Developmental Loci in Drosophila melanogaster. Mol. Biol. Evol. 2009, 26, 1949–1961.
  11. Merenciano, M.; Iacometti, C.; Gonzalez, J. A unique cluster of roo insertions in the promoter region of a stress response gene in Drosophila melanogaster. Mob. DNA 2019, 10, 10.
  12. Petrov, D.A.; Fiston-Lavier, A.-S.; Lipatov, M.; Lenkov, K.; Gonzalez, J. Population Genomics of Transposable Elements in Drosophila melanogaster. Mol. Biol. Evol. 2011, 28, 1633–1644.
  13. Wos, G.; Choudhury, R.R.; Kolář, F.; Parisod, C. Transcriptional activity of transposable elements along an elevational gradient in Arabidopsis arenosa. Mob. DNA 2021, 12, 7.
  14. Gould, S.J.; Vrba, E.S. Exaptation-A Missing Term in the Science of Form. Paleobiology 1982, 8, 4–15.
  15. Jacob, F. Molecular tinkering in evolution. In Evolution from Molecules to Men; Bendall, D.S., Ed.; Cambridge University Press: London, UK, 1983; pp. 131–144.
  16. Schlötterer, C. Genes from scratch—The evolutionary fate of de novo genes. Trends Genet. 2015, 31, 215–219.
  17. Wissler, L.; Gadau, J.; Simola, D.F.; Helmkampf, M.; Bornberg-Bauer, E. Mechanisms and Dynamics of Orphan Gene Emergence in Insect Genomes. Genome Biol. Evol. 2013, 5, 439–455.
  18. Koonin, E.V.; Makarova, K.S.; Aravind, L. Horizontal gene transfer in prokaryotes: Quantification and classification. Annu. Rev. Microbiol. 2001, 55, 709–742.
  19. Crisp, A.; Boschetti, C.; Perry, M.; Tunnacliffe, A.; Micklem, G. Expression of multiple horizontally acquired genes is a hallmark of both vertebrate and invertebrate genomes. Genome Biol. 2015, 16, 50.
  20. Dotto, B.R.; Carvalho, E.L.; Silva, A.F.; Duarte Silva, L.F.; Pinto, P.M.; Ortiz, M.F.; Wallau, G.L. HTT-DB: Horizontally transferred transposable elements database. Bioinformatics 2015, 31, 2915–2917.
  21. Loreto, E.L.S.; Carareto, C.M.A.; Capy, P. Revisiting horizontal transfer of transposable elements in Drosophila. Heredity 2008, 100, 545–554.
  22. Peccoud, J.; Loiseau, V.; Cordaux, R.; Gilbert, C. Massive horizontal transfer of transposable elements in insects. Proc. Natl. Acad. Sci. USA 2017, 114, 4721–4726.
  23. Wallau, G.L.; Capy, P.; Loreto, E.; Le Rouzic, A.; Hua-Van, A. VHICA, a New Method to Discriminate between Vertical and Horizontal Transposon Transfer: Application to the Mariner Family within Drosophila. Mol. Biol. Evol. 2016, 33, 1094–1109.
  24. Barker, M.; Arrigo, N.; Baniaga, A.; Li, Z.; Levin, D. On the relative abundance of autopolyploids and allopolyploids. New Phytol. 2016, 210, 391–398.
  25. Doolittle, W.F. Phylogenetic classification and the universal tree. Science 1999, 284, 2124–2129.
  26. Cordaux, R.; Gilbert, C. Evolutionary Significance of Wolbachia-to-Animal Horizontal Gene Transfer: Female Sex Determination and the f Element in the Isopod Armadillidium vulgare. Genes 2017, 8, 186.
  27. Schoenknecht, G.; Chen, W.-H.; Ternes, C.M.; Barbier, G.G.; Shrestha, R.P.; Stanke, M.; Braeutigam, A.; Baker, B.J.; Banfield, J.F.; Garavito, R.M.; et al. Gene Transfer from Bacteria and Archaea Facilitated Evolution of an Extremophilic Eukaryote. Science 2013, 339, 1207–1210.
  28. Zimorski, V.; Ku, C.; Martin, W.F.; Gould, S.B. Endosymbiotic theory for organelle origins. Curr. Opin. Microbiol. 2014, 22, 38–48.
  29. Escudero, J.A.; Loot, C.; Nivina, A.; Mazel, D. The Integron: Adaptation On Demand. Microbiol. Spectr. 2015, 3, MDNA3-0019-2014.
  30. San Millan, A.; Escudero, J.A.; Gifford, D.R.; Mazel, D.; MacLean, R.C. Multicopy plasmids potentiate the evolution of antibiotic resistance in bacteria. Nat. Ecol. Evol. 2017, 1, 0010.
  31. García-Aljaro, C.; Ballesté, E.; Muniesa, M. Beyond the canonical strategies of horizontal gene transfer in prokaryotes. Curr. Opin. Microbiol. 2017, 38, 95–105.
  32. von Wintersdorff, C.J.H.; Penders, J.; van Niekerk, J.M.; Mills, N.D.; Majumder, S.; van Alphen, L.B.; Savelkoul, P.H.M.; Wolffs, P.F.G. Dissemination of Antimicrobial Resistance in Microbial Ecosystems through Horizontal Gene Transfer. Front. Microbiol. 2016, 7, 173.
  33. Gao, C.; Ren, X.; Mason, A.S.; Liu, H.; Xiao, M.; Li, J.; Fu, D. Horizontal gene transfer in plants. Funct. Integr. Genomics 2014, 14, 23–29.
  34. Gilbert, C.; Cordaux, R. Viruses as vectors of horizontal transfer of genetic material in eukaryotes. Curr. Opin. Virol. 2017, 25, 16–22.
  35. Zhang, H.-H.; Jean Peccoud Xu, M.-R.-X.; Zhang, X.-G.; Gilbert, C. Horizontal transfer and evolution of transposable elements in vertebrates. Nat. Commun. 2020, 11, 1362.
  36. Le Rouzic, A.; Capy, P. The first steps of transposable elements invasion: Parasitic strategy vs. genetic drift. Genetics 2005, 169, 1033–1043.
  37. Palazzo, A.; Caizzi, R.; Viggiano, L.; Marsano, R.M. Does the Promoter Constitute a Barrier in the Horizontal Transposon Transfer Process? Insight from Bari Transposons. Genome Biol. Evol. 2017, 9, 1637–1645.
  38. Palazzo, A.; Lorusso, P.; Miskey, C.; Walisko, O.; Gerbino, A.; Marobbio, C.M.T.; Ivics, Z.; Marsano, R.M. Transcriptionally promiscuous “blurry” promoters in Tc1/mariner transposons allow transcription in distantly related genomes. Mob. DNA 2019, 10, 13.
  39. Lokta, A.J. Elements of Physical Biology; Williams & Wilkins: Baltimore, MD, USA, 1925.
  40. Volterra, V. Fluctuations in the Abundance of a Species considered Mathematically. Nature 1926, 118, 558–560.
  41. Le Rouzic, A.; Boutin, T.S.; Capy, P. Long-term evolution of transposable elements. Proc. Natl. Acad. Sci. USA 2007, 104, 19375–19380.
  42. Robillard, É.; Le Rouzic, A.; Zhang, Z.; Capy, P.; Hua-Van, A. Experimental evolution reveals hyperparasitic interactions among transposable elements. Proc. Natl. Acad. Sci. USA 2016, 113, 14763–14768.
  43. Bouyer, D.; Kramdi, A.; Kassam, M.; Heese, M.; Schnittger, A.; Roudier, F.; Colot, V. DNA methylation dynamics during early plant life. Genome Biol. 2017, 18, 179.
  44. Brennecke, J.; Malone, C.D.; Aravin, A.A.; Sachidanandam, R.; Stark, A.; Hannon, G.J. An Epigenetic Role for Maternally Inherited piRNAs in Transposon Silencing. Science 2008, 322, 1387–1392.
  45. Cappucci, U.; Noro, F.; Casale, A.M.; Fanti, L.; Berloco, M.; Alagia, A.A.; Grassi, L.; Le Pera, L.; Piacentini, L.; Pimpinelli, S. The Hsp70 chaperone is a major player in stress-induced transposable element activation. Proc. Natl. Acad. Sci. USA 2019, 116, 17943–17950.
  46. Gonzalez, J.; Qi, H.; Liu, N.; Lin, H. Piwi Is a Key Regulator of Both Somatic and Germline Stem Cells in the Drosophila Testis. Cell Rep. 2015, 12, 150–161.
  47. Saint-Leandre, B.; Capy, P.; Hua-Van, A.; Filee, J. piRNA and Transposon Dynamics in Drosophila: A Female Story. Genome Biol. Evol. 2020, 12, 931–947.
  48. Saint-Leandre, B.; Clavereau, I.; Hua-Van, A.; Capy, P. Transcriptional polymorphism of pi RNA regulatory genes underlies the mariner activity in Drosophila simulans testes. Mol. Ecol. 2017, 26, 3715–3731.
  49. Lewis, S.H.; Quarles, K.A.; Yang, Y.; Tanguy, M.; Frézal, L.; Smith, S.A.; Sharma, P.P.; Cordaux, R.; Gilbert, C.; Giraud, I.; et al. Pan-arthropod analysis reveals somatic piRNAs as an ancestral defence against transposable elements. Nat. Ecol. Evol. 2018, 2, 174–181.
  50. McCue, A.D.; Nuthikattu, S.; Reeder, S.H.; Slotkin, R.K. Gene Expression and Stress Response Mediated by the Epigenetic Regulation of a Transposable Element Small RNA. PLoS Genet. 2012, 8, 18.
  51. Wang, D.; Qu, Z.; Yang, L.; Zhang, Q.; Liu, Z.-H.; Do, T.; Adelson, D.L.; Wang, Z.-Y.; Searle, I.; Zhu, J.-K. Transposable elements (TEs) contribute to stress-related long intergenic noncoding RNAs in plants. Plant J. 2017, 90, 133–146.
  52. Josse, T.; Boivin, A.; Anxolabehere, D.; Ronsseray, S. P element-encoded regulatory products enhance Repeat-Induced Gene Silencing (RIGS) of P-lacZ-white clusters in Drosophila melanogaster. Mol. Genet. Genom. 2002, 268, 311–320.
  53. Palazzo, A.; Marconi, S.; Specchia, V.; Bozzetti, M.P.; Ivics, Z.; Caizzi, R.; Marsano, R.M. Functional Characterization of the Bari1 Transposition System. PLoS ONE 2013, 8, e79385.
  54. Hirochika, H.; Okamoto, H.; Kakutani, T. Silencing of Retrotransposons in Arabidopsis and Reactivation by the ddm1 Mutation. Plant Cell 2000, 12, 357–368.
  55. Tsukahara, S.; Kobayashi, A.; Kawabe, A.; Mathieu, O.; Miura, A.; Kakutani, T. Bursts of retrotransposition reproduced in Arabidopsis. Nature 2009, 461, 423–426.
  56. Choi, J.Y.; Lee, Y.C.G. Double-edged sword: The evolutionary consequences of the epigenetic silencing of transposable elements. PLoS Genet. 2020, 16, e1008872.
  57. Hollister, J.D.; Gaut, B.S. Epigenetic silencing of transposable elements: A trade-off between reduced transposition and deleterious effects on neighboring gene expression. Genome Res. 2009, 19, 1419–1428.
  58. Lippman, Z.; Gendrel, A.V.; Black, M.; Vaughn, M.W.; Dedhia, N.; McCombie, W.R.; Lavine, K.; Mittal, V.; May, B.; Kasschau, K.D.; et al. Role of transposable elements in heterochromatin and epigenetic control. Nature 2004, 430, 471–476.
  59. Quadrana, L.; Silveira, A.B.; Mayhew, G.F.; LeBlanc, C.; Martienssen, R.A.; Jeddeloh, J.A.; Colot, V. The Arabidopsis thaliana mobilome and its impact at the species level. eLife 2016, 5, e15716.
  60. Wright, S.I.; Agrawal, N.; Bureau, T.E. Effects of Recombination Rate and Gene Density on Transposable Element Distributions in Arabidopsis thaliana. Genome Res. 2003, 13, 1897–1903.
  61. Grover, D.; Majumder, P.P.; Rao, C.B.; Brahmachari, S.K.; Mukerji, M. Nonrandom distribution of Alu elements in genes of various functional categories: Insight from analysis of human chromosomes 21 and 22. Mol. Biol. Evol. 2003, 20, 1420–1424.
  62. Sultana, T.; van Essen, D.; Siol, O.; Bailly-Bechet, M.; Philippe, C.; Zine El Aabidine, A.; Pioger, L.; Nigumann, P.; Saccani, S.; Andrau, J.-C.; et al. The Landscape of L1 Retrotransposons in the Human Genome Is Shaped by Pre-insertion Sequence Biases and Post-insertion Selection. Mol. Cell 2019, 74, 555–570.e7.
  63. Cridland, J.M.; Macdonald, S.J.; Long, A.D.; Thornton, K.R. Abundance and Distribution of Transposable Elements in Two Drosophila QTL Mapping Resources. Mol. Biol. Evol. 2013, 30, 2311–2327.
  64. Favarato, R.M.; Ribeiro, L.B.; Feldberg, E.; Matoso, D.A. Chromosomal mapping of transposable elements of the Rex family in the bristlenose catfish, Ancistrus (Siluriformes, Loricariidae), from the Amazonian region. J. Hered. 2017, 108, 254–261.
  65. Bertocchi, N.A.; de Oliveira, T.D.; del Garnero, A.V.; Buogo Coan, R.L.; Gunski, R.J.; Martins, C.; Torres, F.P. Distribution of CR1-like transposable element in woodpeckers (Aves Piciformes): Z sex chromosomes can act as a refuge for transposable elements. Chromosome Res. 2018, 26, 333–343.
  66. Charlesworth, B.; Langley, C.H. The population genetics of Drosophila transposable elements. Annu. Rev. Genet. 1989, 23, 251–287.
  67. Langley, C.H.; Montgomery, E.; Hudson, R.; Kaplan, N.; Charlesworth, B. On the role of unequal exchange in the containment of transposable element copy number. Genet. Res. 1988, 52, 223–235.
  68. Fontanillas, P.; Hartl, D.L.; Reuter, M. Genome Organization and Gene Expression Shape the Transposable Element Distribution in the Drosophila melanogaster Euchromatin. PLoS Genet. 2007, 3, e210.
  69. Sultana, T.; Zamborlini, A.; Cristofari, G.; Lesage, P. Integration site selection by retroviruses and transposable elements in eukaryotes. Nat. Rev. Genet. 2017, 18, 292–308.
  70. Auvinet, J.; Graça, P.; Ghigliotti, L.; Pisano, E.; Dettaï, A.; Ozouf-Costaz, C.; Higuet, D. Insertion Hot Spots of DIRS1 Retrotransposon and Chromosomal Diversifications among the Antarctic Teleosts Nototheniidae. Int. J. Mol. Sci. 2019, 20, 701.
  71. Biemont, C.; Vieira, C. Genetics—Junk DNA as an evolutionary force. Nature 2006, 443, 521–524.
  72. Mazzuchelli, J.; Martins, C. Genomic organization of repetitive DNAs in the cichlid fish Astronotus ocellatus. Genetica 2009, 136, 461–469.
  73. Pardue, M.L.; DeBaryshe, P.G. Drosophila telomeres: Two transposable elements with important roles in chromosomes. Genetica 1999, 107, 189–196.
  74. Pimpinelli, S.; Berloco, M.; Fanti, L.; Dimitri, P.; Bonaccorsi, S.; Marchetti, E.; Caizzi, R.; Caggese, C.; Gatti, M. Transposable elements are stable structural components of Drosophila melanogaster heterochromatin. Proc. Natl. Acad. Sci. USA 1995, 92, 3804–3808.
  75. Ganko, E.W.; Greene, C.S.; Lewis, J.A.; Bhattacharjee, V.; McDonald, J.F. LTR retrotransposon-gene associations in Drosophila melanogaster. J. Mol. Evol. 2006, 62, 111–120.
  76. Chung, H.; Bogwitz, M.R.; McCart, C.; Andrianopoulos, A.; Ffrench-Constant, R.H.; Batterham, P.; Daborn, P.J. Cis-regulatory elements in the Accord retrotransposon result in tissue-specific expression of the Drosophila melanogaster insecticide resistance gene Cyp6g1. Genetics 2007, 175, 1071–1077.
  77. Daborn, P.J.; Yen, J.L.; Bogwitz, M.R.; Le Goff, G.; Feil, E.; Jeffers, S.; Tijet, N.; Perry, T.; Heckel, D.; Batterham, P.; et al. A single P450 allele associated with insecticide resistance in Drosophila. Science 2002, 297, 2253–2256.
  78. Mateo, L.; Ullastres, A.; González, J. A Transposable Element Insertion Confers Xenobiotic Resistance in Drosophila. PLoS Genet. 2014, 10, e1004560.
  79. Villanueva-Cañas, J.L.; Horvath, V.; Aguilera, L.; González, J. Diverse families of transposable elements affect the transcriptional regulation of stress-response genes in Drosophila melanogaster. Nucleic Acids Res. 2019, 47, 6842–6857.
  80. DeBarry, J.D.; Ganko, E.W.; McCarthy, E.M.; McDonald, J.F. The contribution of LTR retrotransposon sequences to gene evolution in Mus musculus. Mol. Biol. Evol. 2006, 23, 479–481.
  81. Wicker, T.; Gundlach, H.; Spannagl, M.; Uauy, C.; Borrill, P.; Ramirez-Gonzalez, R.H.; De Oliveira, R.; Mayer, K.F.X.; Paux, E.; Choulet, F. Impact of transposable elements on genome structure and evolution in bread wheat. Genome Biol. 2018, 19, 103.
  82. SanMiguel, P.; Tikhonov, A.; Jin, Y.-K.; Motchoulskaia, N.; Zakharov, D.; Melake-Berhan, A.; Springer, P.S.; Edwards, K.J.; Lee, M.; Avramova, Z.; et al. Nested Retrotransposons in the Intergenic Regions of the Maize Genome. Science 1996, 274, 765–768.
  83. SanMiguel, P.; Gaut, B.S.; Tikhonov, A.; Nakajima, Y.; Bennetzen, J.L. The paleontology of intergene retrotransposons of maize. Nat. Genet. 1998, 20, 43–45.
  84. Kofler, R. Dynamics of Transposable Element Invasions with piRNA Clusters. Mol. Biol. Evol. 2019, 36, 1457–1472.
  85. Kofler, R. piRNA Clusters Need a Minimum Size to Control Transposable Element Invasions. Genome Biol. Evol. 2020, 12, 736–749.
  86. Mevel-Ninio, M.; Pelisson, A.; Kinder, J.; Campos, A.R.; Bucheton, A. The flamenco locus controls the gypsy and ZAM retroviruses and is required for Drosophila oogenesis. Genetics 2007, 175, 1615–1624.
  87. Zhang, S.; Pointer, B.; Kelleher, E.S. Rapid evolution of piRNA-mediated silencing of an invading transposable element was driven by abundant de novo mutations. Genome Res. 2020, 30, 566–575.
  88. Chuong, E.B.; Elde, N.C.; Feschotte, C. Regulatory activities of transposable elements: From conflicts to benefits. Nat. Rev. Genet. 2017, 18, 71–86.
  89. Cost, G.J.; Golding, A.; Schlissel, M.S.; Boeke, J.D. Target DNA chromatinization modulates nicking by L1 endonuclease. Nucleic Acids Res. 2001, 29, 573–577.
  90. Levin, H.L.; Moran, J.V. Dynamic interactions between transposable elements and their hosts. Nat. Rev. Genet. 2011, 12, 615–627.
  91. Brennecke, J.; Aravin, A.A.; Stark, A.; Dus, M.; Kellis, M.; Sachidanandam, R.; Hannon, G.J. Discrete small RNA-generating loci as master regulators of transposon activity in Drosophila. Cell 2007, 128, 1089–1103.
  92. Genzor, P.; Konstantinidou, P.; Stoyko, D.; Manzourolajdad, A.; Andrews, C.M.; Elchert, A.R.; Stathopoulos, C.; Haase, A.D. Cellular abundance shapes function in piRNA-guided genome defense 2 (Running title: The functional piRNA sequence space). BioRxiv 2021, 31, 2058–2068.
  93. Kjellman, C.; Sjogren, H.; Widegren, B. The Y-Chromosome—A Graveyard For Endogenous Retroviruses. Gene 1995, 161, 163–170.
  94. Sadeq, S.; Al-Hashimi, S.; Cusack, C.M.; Werner, A. Endogenous Double-Stranded RNA. Non-Coding RNA 2021, 7, 15.
  95. Bhattacharyya, M.; Smith, A.; Ellis, T.; Hedley, C.; Martin, C. The wrinkled-seed character of pea described by Mendel is caused by a transposon-like insertion in a gene coding starch-branching enzyme. Cell 1990, 60, 115–122.
  96. van’t Hof, A.E.; Campagne, P.; Rigden, D.J.; Yung, C.J.; Lingley, J.; Quail, M.A.; Hall, N.; Darby, A.C.; Saccheri, I.J. The industrial melanism mutation in British peppered moths is a transposable element. Nature 2016, 534, 102–105.
  97. Bogwitz, M.R.; Chung, H.; Magoc, L.; Rigby, S.; Wong, W.; O’Keefe, M.; McKenzie, J.A.; Batterham, P.; Daborn, P.J. Cyp12a4 confers lufenuron resistance in a natural population of Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2005, 102, 12807–12812.
  98. Marsano, R.M.; Caizzi, R.; Moschetti, R.; Junakovic, N. Evidence for a functional interaction between the Bari1 transposable element and the cytochrome P450 cyp12a4 gene in Drosophila melanogaster. Gene 2005, 357, 122–128.
  99. Barron, M.G.; Fiston-Lavier, A.-S.; Petrov, D.M.; Gonzalez, J. Population Genomics of Transposable Ele-ments in Drosophila. Annu. Rev. Genet. 2014, 48, 561–581.
  100. Garavís, M.; González, C.; Villasante, A. On the Origin of the Eukaryotic Chromosome: The Role of Noncanonical DNA Structures in Telomere Evolution. Genome Biol. Evol. 2013, 5, 1142–1150.
  101. Koonin, E.V.; Krupovic, M. Evolution of adaptive immunity from transposable elements combined with innate immune systems. Nat. Rev. Genet. 2015, 16, 184–192.
  102. Lavialle, C.; Cornelis, G.; Dupressoir, A.; Esnault, C.; Heidmann, O.; Vernochet, C.; Heidmann, T. Paleovirology of ‘syncytins’, retroviral env genes exapted for a role in placentation. Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20120507.
  103. Etchegaray, E.; Naville, M.; Volff, J.-N.; Haftek-Terreau, Z. Transposable element-derived sequences in vertebrate development. Mob. DNA 2021, 12, 1.
  104. Joly-Lopez, Z.; Bureau, T.E. Exaptation of transposable element coding sequences. Curr. Opin. Genet. Dev. 2018, 49, 34–42.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 544
Revisions: 2 times (View History)
Update Date: 06 Jan 2022
1000/1000
ScholarVision Creations