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 -- 2421 2022-05-21 06:32:57 |
2 Format Correct Meta information modification 2421 2022-05-23 06:49:35 | |
3 Format Correct + 1 word(s) 2422 2022-05-23 06:50:35 |

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.
Ahmad, S.; Srikulnath, K.; , .; Panthum, T. Transposable Elements during Sex Chromosome Differentiation. Encyclopedia. Available online: https://encyclopedia.pub/entry/23194 (accessed on 14 May 2024).
Ahmad S, Srikulnath K,  , Panthum T. Transposable Elements during Sex Chromosome Differentiation. Encyclopedia. Available at: https://encyclopedia.pub/entry/23194. Accessed May 14, 2024.
Ahmad, Syed, Kornsorn Srikulnath,  , Thitipong Panthum. "Transposable Elements during Sex Chromosome Differentiation" Encyclopedia, https://encyclopedia.pub/entry/23194 (accessed May 14, 2024).
Ahmad, S., Srikulnath, K., , ., & Panthum, T. (2022, May 21). Transposable Elements during Sex Chromosome Differentiation. In Encyclopedia. https://encyclopedia.pub/entry/23194
Ahmad, Syed, et al. "Transposable Elements during Sex Chromosome Differentiation." Encyclopedia. Web. 21 May, 2022.
Transposable Elements during Sex Chromosome Differentiation
Edit

Transposable elements (TEs) comprise a substantial portion of eukaryotic genomes. They have the unique ability to integrate into new locations and serve as the main source of genomic novelties by mediating chromosomal rearrangements and regulating portions of functional genes. Recent studies have revealed that TEs are abundant in sex chromosomes.

sex determination transposable element Ty3/Gypsy

1. Roles of Transposable Elements during Sex Chromosome Differentiation

Sex determination is the process by which organisms develop as either male or female. Sex determination mechanisms differ among species. Genome-level studies in many model and non-model species have elucidated the genetic mechanisms associated with sex chromosome differentiation. The process of sex chromosome degeneration and rearrangement, in which an obvious increase in repetitive non-coding sequences and transposable elements (TEs) is observed for most sex chromosomes [1][2][3][4][5], is an important biological constraint that sheds light on the evolution of sex chromosomes [6][7][8][9]. Using comparative genomic tools, the dynamics of functional transposable elements (TEs) on sex chromosomes can be identified in two aspects: (1) sequence conservation and (2) co-localization with regions with a known genomic function [10][11]. The conservation of TE sequences likely passively contributes to sex chromosome degeneration [12]. TE accumulation on the W/Y sex chromosome may start from several insertions in the close vicinity of crucial heterogametic sex-linked genes/loci where TEs can escape removal from the population due to stochastic processes, such as methylation/heterochromatization, ectopic recombination between TEs, or size reduction on W/Y sex chromosomes [13][14][15][16].
In contrast to mammals, birds and many snakes, which have small and degenerate W/Y chromosomes, the heterogametic sex chromosome (W/Y) is substantially larger than the Z/X chromosome in many fishes, reptiles and amphibians, indicating that sex chromosomes in these groups are usually younger, with frequent turnover [4][17][18][19][20][21][22][23]. In the medaka fish Y chromosome, the male-specific sex-determining region (SDR) has undergone duplication after TE insertion into the proto-Y chromosome [24]. In platyfish (Xiphophorus maculatus, Günther, 1866) [25], a higher repeat content was found in the SDRs of the X (43% repetitive sequences with 36% TEs) and Y (49% repetitive sequences with 32% TEs) chromosomes than in the whole genome (23% repeats with 21% TEs). The X and Y SDRs have higher densities of repeats and TEs than the whole X chromosome [26][27]. This suggests that TEs occupy the whole Y SDR, but are more compartmentalized on the X SDR [28]. A similar phenomenon occurs in the W chromosome of the half-smooth tongue sole (Cynoglossus semilaevis, Günther, 1873) [29], which has a substantially higher TE content than the Z chromosome (Chen et al. 2014) [30]. This indicates that TEs are recently strongly active in the X and Y or Z and W SDRs, thus potentially increasing the size and divergence of the currently small non-pseudo-autosomal region [31][32]. In the genome of the African clawed frog (Xenopus laevis, Daudin 1802) [33], recombination between the W and Z sex chromosomes stopped recently, with a strong accumulation of TEs in W-specific regions [34]. Potentially, the accumulation of TEs interfered with chromosome pairing during prophase I of meiosis and suppressed recombination, leading to an increase in sex chromosome size during the early phase of their differentiation, whereas a size reduction occurred later in their evolution and thus the W/Y sex chromosomes became smaller [35][36][37][38][39].
The transposition of a pre-existing SD locus onto a new chromosome may lead to the emergence of a new sex chromosome [40]. TEs displace DNA and may promote the emergence of new SD loci [41][42][43]. Recent studies have shown that the master sex-determining (SD) gene, sdY, is conserved in many species. However, it is not consistently located on the same linkage homology but seems to behave like a “jumping gene” [42][44][45]. Analysis of the boundaries of the moving region that harbors sdY revealed the presence of several TE sequences, based on which a mechanism involving TE-associated transduction has been proposed [46]. Similarly, in medaka, TEs co-localize with regions that have a known genomic function and play an active role in the evolution of sex chromosomes, and the integration of DNA via TEs contributed to the regulation of the newly emerged SD gene, dmrt1bY. The SD gene of medaka arose from a duplication event of the autosomal dmrt1a gene [47]. The two dmrt1 genes exert their functions at different times during gonad development: dmrt1bY in the SD stage and dmrt1a in the differentiating and adult testes [48]. This phenomenon may be linked to a rapid turnover of the sex chromosomes in the lineage, e.g., the genus Oncorhynchus has six independent sex chromosome pairs that originated 6–8 million years ago (MYA) [45].
The variety of Y chromosomes probably results from TEs (TC1-like transposase and non-LTR retrotransposons) in the flanking regions of the SDRs, which can move throughout the genome [43]. The movement of male/female SDRs among autosomes may prevent sex chromosome degradation and deleterious TE accumulation [28]. Alternatively, TEs have been shown to play a role in heterochromatization of the W/Y chromosomes and dosage compensation mechanisms [30][49][50][51][52][53][54][55]
TEs may play a role in resolving dosage-related gene expression problems in several species by promoting the silencing and condensation of sex chromatin in X (or Z) chromosomes, known as “the hitchhiking effect of favorable mutations” [6]. Interspersed repeat elements such as L1s in humans and mice have been suggested to enhance the inactivation process, thus promoting the heterochromatin state [56]. X chromosome inactivation in mammals, also termed “lyonization”, is a dosage compensation process in which one of the two X chromosomes is inactivated in XX females during early embryogenesis, preventing gene overexpression in comparison to males, which have a single X chromosome [57]. In marsupials, such as opossums, the paternal X chromosome is inactivated, whereas in placental mammals, a random X chromosome is inactivated by a long non-coding RNA produced by Xist [58]. Interestingly, in opossums, L1s show equally frequent interruptions on the X chromosome and autosomes, whereas in humans, L1s are less frequently interrupted on the X chromosome than on the autosomes [59]. On the human X chromosome, L1-poor regions contain genes that escape X inactivation and are physically distant from Xist-silenced regions [60]. This suggests that L1s play a role in the spreading of X chromosome silencing by recruiting Xist RNAs, which play a general role in the regulation of X-gene expression. This hypothesis has been tested in the spiny rat (Tokudaia osimensis, Abe, 1934) [61], of which males and females are XO [62][63]. A similar high concentration of LINEs was observed on both male and female X chromosomes, whereas no dosage compensation by X inactivation is required. This suggests that LINEs are not required on the X chromosome [64]. It is unclear whether the unique TE distributions patterns on the X chromosome are the cause or consequence of inactivation (or both). One possibility is that L1 accumulation on the X chromosome may be only a by-product of reduced recombination. Further investigation of the role of LINEs in X chromosome inactivation in mammals and other organisms is necessary.
Recently, the TE drbx1 was found to be inserted in an intron of the X-linked region encoding the SD gene dmrt1 in Siamese fighting fish (Betta splendens, Regan, 1910 [65]) [66]. This structural change was associated with a shift in the epigenetic silencing of X-dmrt1 during the critical sex determination stage. Similar mechanisms have been found in plants, e.g., in melon (Cucumis melo L.), and TE-induced methylation of the promoter of the transcription factor CmWIP1 has been shown to suppress expression and thus realize sex determination [67].
Some TE families are differentially expressed in specific tissues or conditions between sexes [12]. In platyfish, an accumulation of Texim genes is observed on the Y chromosome [68]. These genes are physically associated with a Helitron, which may have spread the Texim sequences on Y, but not X. The higher density of TEs on Y may suggest that they regulate some key sexual developmental genes and, consequently, impact sexual development [12]. In the medaka genome, a dmrt1-binding element in the promoter of the SD gene dmrt1bY mediates downregulation through its own gene product and autosomal dmrt1a. This dmrt1-binding silencer was introduced in the dmrt1bY promoter through the insertion of a novel TE, termed Izanagi, which is present in multiple copies in the genome and acts as a transcription factor-binding site [69]. This event contributed to the transcriptional rewiring of the new SD gene that created evolutionary innovations [70][71]. A recent study assessed TE expression on sex chromosomes of Drosophila melanogaster (Meigen, 1830) [72] and found specific enrichment of expressed TEs on the Y chromosome that were depleted on the X chromosome [73]. This result suggested that high-level expression of Y-specific TEs was associated with the activation of spermatocyte-specific and Y chromosome-specific transcriptional pathways.
The discovery of sex determination in the domestic silk moth (Bombyx mori, Linnaeus, 1758 [74]) demonstrated the regulatory function of the W chromosome, where TEs are involved in physical and biochemical interactions with thousands of autosomal protein-coding genes [50][52][75]. In the house fly (Musca domestica, Linnaeus, 1758 [53]), approximately two-thirds of the Y-linked scaffolds contain sequence similarities with TEs (Meisel et al. 2017 [54]). Neo-X chromosome formation through a domesticated non-autonomous Helitron has been identified in Drosophila miranda (Dobzhansky, 1935 [55]), and its role in the expression of X-linked genes has been revealed [51]. By contrast, the formation of male-specific lethal binding sites contributes to the dosage compensation process [51]. In mammalian genomes, TEs are hypomethylated in females compared to males, implying that oocytes are more resilient to TE transposition than the male germline. This may be linked to the lifelong division and numerous cell divisions of spermatogonial cells in contrast to oocytes. More cell divisions may allow for many deleterious insertions in the male germline [76].
In addition to the effects of TEs on host gene expression, there exist genomic differences between males and females in terms of TE profiles that impact sexual development [12]. Subsequent cycles of W/Y chromosome degeneration and rejuvenation may differ in length of evolutionary time. During W/Y chromosome degeneration, the appearance of a large number of TE transcripts in a cell may be more or less sudden [28]. In species/populations with a fixed W/Y chromosome loss or with frequent turnover, the period of TE transcript increase may be too short to induce trans-generational and constant preparedness of the genome defense system for TE invasion. When genome defense is inefficient, W/Y chromosome degradation may proceed at a faster rate, and the cycle of chromosome rejuvenation may be shorter [28]. The W/Y chromosome may become a substantial source of functional TE transcripts in the cell and threaten whole-genome stability [77][78][79][80]. The high density of TEs on the W/Y chromosome may serve as a hallmark for heterochromatin marks, which result in different levels of chromatin repression in the rest of the genome and in differential gene expression between males and females [6][81]. TE accumulation has been observed on specific Y or W sex chromosomes and on the corresponding regions of the X and Z chromosomes [15][82]. It is well established that TEs are abundant and play roles in both heterogametic and homogametic sex chromosomes; however, the question remains as to whether TEs are passive or active drivers of sex chromosome differentiation and evolution.

2. Ty3/Gypsy Transposable Elements and Sex Chromosome Differentiation

Substantial evidence suggests a close relationship between retroviruses and certain LTR retrotransposons, such as Ty1/Copia and Ty3/Gypsy [12][83][84][85][86][87]. Ty3/Gypsy TEs are phylogenetically more closely related to vertebrate retroviruses (Retroviridae) than to the Ty1/Copia class (Pseudoviridae) [88]. A few Ty3/Gypsy elements have a third open reading frame, putatively encoding an envelope protein, such as in the fruit fly (D. melanogaster) [85]. Thus, some Ty3/Gypsy-related elements display similarity to retroviruses in that they have an envelope (env) gene, which may indicate their infectious nature [89][90][91]. Several Ty3/Gypsy elements have occasionally been transmitted horizontally. For example, the envelope-encoding Gypsy element has jumped between species of the D. melanogaster subgroup [92], while the sea urchin retroviral-like SURL elements (no apparent envelope gene) have been transferred between echinoid species [93][94]. The Sushi retrotransposon, which belongs to the Ty3/Gypsy class, was originally observed in the Japanese pufferfish (Fugu rubripes Temminck and Schlegel, 1850 [95]) [96]. Other partial elements (mostly partial pol sequences) have been identified in lampreys, fishes, amphibians and reptiles [97][98][99][100]. The Hsr1 element in terrestrial salamanders (Hydromantes spp.) shows a high degree of similarity to the pufferfish Sushi element [95][99]. This suggests that Ty3/Gypsy elements are widely distributed across living organisms [100][101]. However, they are present in extremely low copy numbers in mammalian genomes [27].
Ty3/Gypsy elements are stable in nearly all fruit fly (D. melanogaster) strains studied to date, and their copies are generally located in centromeric and/or pericentromeric regions, as also observed in D. virilis and D. subobscura [102]. The papaya HSY region occupies only 13% of the Y chromosome [83]; however, analysis of HSY bacterial artificial chromosomes revealed that papaya HSY has a high repeat content [83]. Comparative genomics revealed that Ty3/Gypsy elements are highly accumulated in the SDR and account for 46.3% of the HSY and 37.7% of the corresponding X chromosome. By contrast, Ty1/Copia elements are less abundant than Ty3/Gypsy in both the HSY and the corresponding X chromosome. The Ty1/Copia content in the corresponding X chromosome is 1.3% higher than that in the HSY, suggesting that Ty1/Copia elements are not a major contributor to repeat accumulation in both the HSY and the corresponding X chromosome. This differs from the TE accumulation scenario in the broadleaf arrowhead (Sagittaria latifolia, Willd, 1805 [103]) Y chromosome, where Ty1/Copia elements are more abundant than Ty3/Gypsy elements [104].
In animal lineages, genome-wide SNP analyses of fishes and squamate reptiles have shown that most sex-specific or sex-linked loci are more strongly associated with Ty3/Gypsy than with other TEs [105][106][107][108][109][110][111]. This might be because of the role of Ty3/Gypsy elements in sex chromosome differentiation or the novel genomic impacts. It is also possible that partial Ty3/Gypsy accumulation on sex-specific or sex-linked loci follows a stochastic pattern within a restricted set of species, representing random homologies, given that only small sets of genomic regions in a restricted set of species are involved. Investigating the possible existence of Ty3/Gypsy elements in other organisms using fluorescence in situ hybridization mapping and whole-genome sequencing may further substantiate this hypothesis. The cumulative load of functional TEs from different classes on W/Y or other chromosomes has not yet been investigated in detail in many non-model species because of difficulties in the sequencing of W/Y chromosomes with highly repetitive sequences [8][79][112][113] and in the assignment of functional TE transcripts to particular loci [114]. Given these limitations,  the transposition rates of particular TE families may have been underestimated and/or artificially standardized over evolutionary time [6][115].
The hypothesis of how TE accumulation impacted SD genes in different biological functions related to sex determination should be further investigated. SNP-based studies can also be integrated with transcriptomics and epigenomic level analyses to reveal the candidate SD genes regulated by accumulated TEs in the SDR to promote a better understanding. This may allow a better understanding of W/Y sex chromosome evolution and function.

References

  1. Singchat, W.; O’Connor, R.E.; Tawichasri, P.; Suntronpong, A.; Sillapaprayoon, S.; Suntrarachun, S.; Muangmai, N.; Baicharoen, S.; Peyachoknagul, S.; Chanhome, L.; et al. Chromosome map of the Siamese cobra: Did partial synteny of sex chromosomes in the amniote represent “a hypothetical ancestral super-sex chromosome” or random distribution? BMC Genom. 2018, 19, 939.
  2. Singchat, W.; Ahmad, S.F.; Sillapaprayoon, S.; Muangmai, N.; Duengkae, P.; Peyachoknagul, S.; O’Connor, R.E.; Griffin, D.K.; Srikulnath, K. Partial amniote sex chromosomal linkage homologies shared on snake W sex chromosomes support the ancestral super-sex chromosome evolution in amniotes. Front. Genet. 2020, 11, 948.
  3. Singchat, W.; Sillapaprayoon, S.; Muangmai, N.; Baicharoen, S.; Indananda, C.; Duengkae, P.; Peyachoknagul, S.; O’Connor, R.E.; Griffin, D.K.; Srikulnath, K. Do sex chromosomes of snakes, monitor lizards, and iguanian lizards result from multiple fission of an “ancestral amniote super-sex chromosome”? Chromosome Res. 2020, 28, 209–228.
  4. Singchat, W.; Ahmad, S.F.; Laopichienpong, N.; Suntronpong, A.; Panthum, T.; Griffin, D.K.; Srikulnath, K. Snake W sex chromosome: The shadow of ancestral amniote super-sex chromosome. Cells 2020, 9, 2386.
  5. Singchat, W.; Panthum, T.; Ahmad, S.F.; Baicharoen, S.; Muangmai, N.; Duengkae, P.; Griffin, D.K.; Srikulnath, K. Remnant of unrelated amniote sex chromosomal linkage sharing on the same chromosome in house gecko lizards, providing a better understanding of the ancestral super-sex chromosome. Cells 2021, 10, 2969.
  6. Charlesworth, B.; Charlesworth, D. The degeneration of Y chromosomes. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2000, 355, 1563–1572.
  7. Bachtrog, D. The temporal dynamics of processes underlying Y-chromosome degeneration. Genetics 2008, 179, 1513–1525.
  8. Carvalho, B.; Koerich, L.B.; Clark, A.G. Origin and evolution of Y chromosomes: Drosophila tales. Trends Genet. 2009, 25, 270–277.
  9. Rifkin, J.L.; Beaudry, F.E.; Humphries, Z.; Choudhury, B.I.; Barrett, S.C.; Wright, S.I. Widespread recombination suppression facilitates plant sex chromosome evolution. Mol. Biol. Evol. 2021, 38, 1018–1030.
  10. Arkhipova, I.R. Neutral theory, transposable elements, and eukaryotic genome evolution. Mol. Biol. Evol. 2018, 35, 1332–1337.
  11. Makałowski, W.; Gotea, V.; Pande, A.; Makałowska, I. Transposable elements: Classification, identification, and their use as a tool for comparative genomics. Evol. Genom. 2019, 1910, 177–207.
  12. Dechaud, C.; Volff, J.N.; Schartl, M.; Naville, M. Sex and the TEs: Transposable elements in sexual development and function in animals. Mob. DNA 2019, 10, 42.
  13. VanBuren, R.; Zeng, F.C.; Chen, C.X.; Zhang, J.S.; Wai, C.M.; Han, J.; Aryal, R.; Gschwend, A.R.; Wang, J.P.; Na, J.K.; et al. Originanddomestication of papaya Yh chromosome. Genome Res. 2015, 25, 524–533.
  14. 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.
  15. Hobza, R.; Cegan, R.; Jesionek, W.; Kejnovsky, E.; Vyskot, B.; Kubat, Z. Impact of repetitive elements on the Y chromosome formation in plants. Genes 2017, 8, 302.
  16. Muyle, A.; Bachtrog, D.; Marais, G.; Turner, J. Epigenetics drive the evolution of sex chromosomes in animals and plants. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2021, 376, 20200124.
  17. Ahmad, S.F.; Singchat, W.; Jehangir, M.; Panthum, T.; Srikulnath, K. Consequence of paradigm shift with repeat landscapes in reptiles: Powerful facilitators of chromosomal rearrangements for diversity and evolution. Genes 2020, 11, 827.
  18. Matsubara, K.; Tarui, H.; Toriba, M.; Yamada, K.; Nishida, C.; Agata, K.; Matsuda, Y. Evidence for different origin of sex chromosomes in snakes, birds, and mammals and step-wise differentiation of snake sex chromosomes. Proc. Natl. Acad. Sci. USA 2006, 103, 18190–18195.
  19. Bellott, D.W.; Hughes, J.F.; Skaletsky, H.; Brown, L.G.; Pyntikova, T.; Cho, T.J.; Koutseva, N.; Zaghlul, S.; Graves, T.; Rock, S.; et al. Mammalian Y chromosomes retain widely expressed dosage-sensitive regulators. Nature 2014, 508, 494–499.
  20. Hughes, J.F.; Skaletsky, H.; Pyntikova, T.; Graves, T.A.; van Daalen, S.K.; Minx, P.J.; Fulton, R.S.; McGrath, S.D.; Locke, D.P.; Friedman, C.; et al. Chimpanzee and human Y chromosomes are remarkably divergent in structure and gene content. Nature 2010, 463, 536–539.
  21. Hughes, J.F.; Skaletsky, H.; Brown, L.G.; Pyntikova, T.; Graves, T.; Fulton, R.S.; Dugan, S.; Ding, Y.; Buhay, C.J.; Kremitzki, C.; et al. Strict evolutionary conservation followed rapid gene loss on human and rhesus Y chromosomes. Nature 2012, 482, 82–86.
  22. Soh, Y.S.; Alföldi, J.; Pyntikova, T.; Brown, L.G.; Graves, T.; Minx, P.J.; Page, D.C. Sequencing the mouse Y chromosome reveals convergent gene acquisition and amplification on both sex chromosomes. Cell 2014, 159, 800–813.
  23. Ahmad, S.F.; Singchat, W.; Panthum, T.; Srikulnath, K. Impact of repetitive DNA elements on snake genome biology and evolution. Cells 2021, 10, 1707.
  24. Kondo, M.; Hornung, U.; Nanda, I.; Imai, S.; Sasaki, T.; Shimizu, A.; Asakawa, S.; Hori, H.; Schmid, M.; Shimizu, N.; et al. Genomic organization of the sex-determining and adjacent regions of the sex chromosomes of medaka. Genome Res. 2006, 16, 815–826.
  25. Günther . 1866. Available online: https://www.gbif.org/species/2350164 (accessed on 5 January 2022).
  26. Kallman, K.D.; Schreibman, M.P. A sex-linked gene controlling gonadotrop differentiation and its significance in determining the age of sexual maturation and size of the platyfish, Xiphophorus maculatus. Gen. Comp. Endocrinol. 1973, 21, 287–304.
  27. Volff, J.N.; Schartl, M. Variability of genetic sex determination in poeciliid fishes. Genetica 2001, 111, 101–110.
  28. Śliwińska, E.B.; Martyka, R.; Tryjanowski, P. Evolutionary interaction between W/Y chromosome and transposable elements. Genetica 2016, 144, 267–278.
  29. Günther . 1873. Available online: https://www.gbif.org/species/2409626 (accessed on 6 January 2022).
  30. Chen, S.; Zhang, G.; Shao, C.; Huang, Q.; Liu, G.; Zhang, P.; Song, W.; An, N.; Chalopin, D.; Volff, J.N.; et al. Whole-genome sequence of a flatfish provides insights into ZW sex chromosome evolution and adaptation to a benthic lifestyle. Nat. Genet. 2014, 46, 253–260.
  31. Blavet, N.; Blavet, H.; Čegan, R.; Zemp, N.; Zdanska, J.; Janoušek, B.; Hobza, R.; Widmer, A. Comparative analysis of a plant pseudoautosomal region (PAR) in Silene latifolia with the corresponding S. vulgaris autosome. BMC Genom. 2012, 13, 226.
  32. Ishii, K.; Nishiyama, R.; Shibata, F.; Kazama, Y.; Abe, T.; Kawano, S. Rapid degeneration of noncoding DNA regions surrounding SlAP3X/Y after recombination suppression in the dioecious plant Silene latifolia. G3 2013, 3, 2121–2130.
  33. Daudin . 1802. Available online: https://www.gbif.org/species/5217334 (accessed on 6 January 2022).
  34. Mawaribuchi, S.; Takahashi, S.; Wada, M.; Uno, Y.; Matsuda, Y.; Kondo, M.; Fukui, A.; Takamatsu, N.; Taira, M.; Ito, M. Sex chromosome differentiation and the W-and Z-specific loci in Xenopus laevis. Dev. Biol. 2017, 426, 393–400.
  35. Bachtrog, D. Adaptation shapes patterns of genome evolution in sexual and asexual genomes in Drosophila. Nat. Genet. 2003, 34, 215–219.
  36. Hoehn, K.B.; Gall, A.; Bashford-Rogers, R.; Fidler, S.J.; Kaye, S.; Weber, J.N.; McClure, M.O.; Kellam, P.; Pybus, O.G. Dynamics of immunoglobulin sequence diversity in HIV-1 infected individuals. Philos. Trans. R. Soc. B 2015, 370, 20140241.
  37. Schemberger, M.O.; Nascimento, V.D.; Coan, R.; Ramos, É.; Nogaroto, V.; Ziemniczak, K.; Valente, G.T.; Moreira-Filho, O.; Martins, C.; Vicari, M.R. DNA transposon invasion and microsatellite accumulation guide W chromosome differentiation in a Neotropical fish genome. Chromosoma 2019, 128, 547–560.
  38. Conte, M.A.; Clark, F.E.; Roberts, R.B.; Xu, L.; Tao, W.; Zhou, Q.; Wang, D.; Kocher, T.D. Origin of a giant sex chromosome. Mol. Biol. Evol. 2021, 38, 1554–1569.
  39. Pourrajab, F.; Hekmatimoghaddam, S. Transposable elements, contributors in the evolution of organisms (from an arms race to a source of raw materials). Heliyon 2021, 7, e06029.
  40. Graves, J.A.M.; Peichel, C.L. Are homologies in vertebrate sex determination due to shared ancestry or to limited options? Genome Biol. 2010, 11, 205.
  41. Yano, K.; Takashi, T.; Nagamatsu, S.; Kojima, M.; Sakakibara, H.; Kitano, H.; Matsuoka, M.; Aya, K. Efficacy of microarray profiling data combined with QTL mapping for the identification of a QTL gene controlling the initial growth rate in rice. Plant Cell Physiol. 2012, 53, 729–739.
  42. Yano, A.; Nicol, B.; Jouanno, E.; Quillet, E.; Fostier, A.; Guyomard, R.; Guiguen, Y. The sexually dimorphic on the Y-chromosome gene (sdY) is a conserved male-specific Y-chromosome sequence in many salmonids. Evol. Appl. 2013, 6, 486–496.
  43. Faber-Hammond, J.J.; Phillips, R.B.; Brown, K.H. Comparative analysis of the shared sex-determination region (SDR) among salmonid fishes. Genome Biol. Evol. 2015, 7, 1972–1987.
  44. Eisbrenner, W.D.; Botwright, N.; Cook, M.; Davidson, E.A.; Dominik, S.; Elliott, N.G.; Henshall, J.; Jones, S.L.; Kube, P.D.; Lubieniecki, K.P.; et al. Evidence for multiple sex-determining loci in Tasmanian Atlantic salmon (Salmo salar). Heredity 2014, 113, 86–92.
  45. Lubieniecki, K.P.; Lin, S.; Cabana, E.I.; Li, J.; Lai, Y.Y.; Davidson, W.S. Genomic instability of the sex-determining locus in Atlantic salmon (Salmo salar). G3 2015, 5, 2513–2522.
  46. Larson, J.; Mattu, S.; Kirchner, L.; Angwin, J. How We Analyzed the COMPAS Recidivism Algorithm; ProPublica: New York, NY, USA, 2016; Volume 9, p. 3.
  47. Matsuda, M. Sex determination in the teleost medaka, Oryzias latipes. Annu. Rev. Genet. 2005, 39, 293–307.
  48. Hornung, E.; Vilisics, F.; Szlávecz, K. Szárazföldi ászkarák (Isopoda, Oniscidea) fajok tipizálása hazai előfordulási adatok alapján (különös tekintettel a sikeres megtelepedőkre). Természetvédelmi Közlemények 2007, 13, 47–57.
  49. Chaumeil, J.; Waters, P.D.; Koina, E.; Gilbert, C.; Robinson, T.J.; Graves, J.A.M. Evolution from XIST-independent to XIST-controlled X-chromosome inactivation: Epigenetic modifications in distantly related mammals. PLoS ONE 2011, 6, e19040.
  50. Branco, A.T.; Tao, Y.; Hartl, D.L.; Lemos, B. Natural variation of the Y chromosome suppresses sex ratio distortion and modulates testis-specific gene expression in Drosophila simulans. Heredity 2013, 111, 8–15.
  51. Ellison, C.E.; Bachtrog, D. Dosage compensation via transposable element mediated rewiring of a regulatory network. Science 2013, 342, 846–850.
  52. Sackton, T.B.; Hartl, D.L. Meta-analysis reveals that genes regulated by the Y chromosome in Drosophila melanogaster are preferentially localized to repressive chromatin. Genome Biol. Evol. 2013, 5, 255–266.
  53. Vicoso, B.; Emerson, J.J.; Zektser, Y.; Mahajan, S.; Bachtrog, D. Comparative sex chromosome genomics in snakes: Differentiation, evolutionary strata, and lack of global dosage compensation. PLoS Biol. 2013, 11, 1001643.
  54. Zhou, Q.; Ellison, C.E.; Kaiser, V.B.; Alekseyenko, A.A.; Gorchakov, A.A.; Bachtrog, D. The epigenome of evolving Drosophila neo-sex chromosomes: Dosage compensation and heterochromatin formation. PLoS Biol. 2013, 11, e1001711.
  55. White, M.A.; Kitano, J.; Peichel, C.L. Purifying selection maintains dosage-sensitive genes during degeneration of the threespine stickleback Y chromosome. Mol. Biol. Evol. 2015, 32, 1981–1995.
  56. Lyon, M.F. The Lyon and the LINE hypothesis. Semin. Cell Dev. Biol. 2003, 14, 313–318.
  57. Roberts, N.B.; Juntti, S.A.; Coyle, K.P.; Dumont, B.L.; Stanley, M.K.; Ryan, A.Q.; Fernald, R.D.; Roberts, R.B. Polygenic sex determination in the cichlid fish Astatotilapia burtoni. BMC Genom. 2016, 17, 835.
  58. Shevchenko, A.I.; Zakharova, I.S.; Zakian, S.M. The evolutionary pathway of X chromosome inactivation in mammals. Acta Nat. 2013, 5, 40–53.
  59. Abrusán, G.; Giordano, J.; Warburton, P.E. Analysis of transposon interruptions suggests selection for L1 elements on the X chromosome. PLoS Genet. 2008, 4, e1000172.
  60. Lyon, M.F. LINE-1 elements and X chromosome inactivation: A function for “junk” DNA? Proc. Natl. Acad. Sci. USA 2000, 97, 6248–6249.
  61. Abe, Y. On the Amami Oshima spiny rat. Bot. Zool. 1934, 3, 936–942.
  62. Honda, T.; Suzuki, H.; Ithu, M. An unusual sex chromosome constitution found in the amami spinous country-rat, Tokudaia osimensis osimensis. Jpn. Genet. 1977, 52, 247–249.
  63. Kobayashi, T.; Yamada, F.; Hashimoto, A.S.; Matsuda, Y.; Kuroiwa, A. Centromere repositioning in the X chromosome of XO/XO mammals, Ryukyu spiny rat. Chromosome Res. 2008, 16, 587.
  64. Hansen, R.S. X inactivation-specific methylation of LINE-1 elements by DNMT3B: Implications for the Lyon repeat hypothesis. Hum. Mol. Genet. 2003, 12, 2559–2567.
  65. Regan . 1910. Available online: https://www.gbif.org/species/2393998 (accessed on 7 January 2022).
  66. Wang, L.; Sun, F.; Wan, Z.Y.; Yang, Z.; Tay, Y.X.; Lee, M.; Ye, B.; Wen, Y.; Meng, Z.; Fan, B.; et al. Transposon-induced epigenetic silencing in the X chromosome as a novel form of dmrt1 expression regulation during sex determination in the fighting fish. BMC Biol. 2022, 20, 5.
  67. Martin, A.; Troadec, C.; Boualem, A.; Rajab, M.; Fernandez, R.; Morin, H.; Pitrat, M.; Dogimont, C.; Bendahmane, A. A transposon-induced epigenetic change leads to sex determination in melon. Nature 2009, 461, 1135–1138.
  68. Tomaszkiewicz, M.; Chalopin, D.; Schartl, M.; Galiana, D.; Volff, J.N. A multicopy Y-chromosomal SGNH hydrolase gene expressed in the testis of the platyfish has been captured and mobilized by a Helitron transposon. BMC Genet. 2014, 15, 44.
  69. Herpin, A.; Braasch, I.; Kraeussling, M.; Schmidt, C.; Thoma, E.C.; Nakamura, S.; Tanaka, M.; Schartl, M. Transcriptional rewiring of the sex determining dmrt1 gene duplicate by transposable elements. PLoS Genet. 2010, 6, e1000844.
  70. Britten, R.J.; McCormack, T.J.; Mears, T.L.; Davidson, E.H. Gypsy/Ty3-class retrotransposons integrated in the DNA of herring, tunicate, and echinoderms. J. Mol. Evol. 1995, 40, 13–24.
  71. Feschotte, C. Transposable elements and the evolution of regulatory networks. Nat. Rev. Genet. 2008, 9, 397–405.
  72. Meigen . 1830. Available online: https://www.gbif.org/species/11195063 (accessed on 5 January 2022).
  73. Lawlor, M.A.; Cao, W.; Ellison, C.E. A transposon expression burst accompanies the activation of Y-chromosome fertility genes during Drosophila spermatogenesis. Nat Commun. 2021, 12, 6854.
  74. Linnaeus . 1758. Available online: https://www.gbif.org/species/1868664 (accessed on 5 January 2022).
  75. Hara, K.; Fujii, T.; Suzuki, Y.; Sugano, S.; Shimada, T.; Katsuma, S.; Kawaoka, S. Altered expression of testis-specific genes, piRNAs, and transposons in the silkworm ovary masculinized by a W chromosome mutation. BMC Genom. 2012, 13, 119.
  76. Zamudio, N.; Bourc’his, D. Transposable elements in the mammalian germline: A comfortable niche or a deadly trap? Heredity 2010, 105, 92–104.
  77. 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.
  78. Abe, H.; Seki, M.; Ohbayashi, F.; Tanaka, N.; Yamashita, J.; Fujii, T.; Yokoyama, T.; Takahashi, M.; Banno, Y.; Sahara, K.; et al. Partial deletions of the W chromosome due to reciprocal translocation in the silkworm Bombyx mori. Insect Mol. Biol. 2005, 14, 339–352.
  79. Matzke, M.; Kanno, T.; Daxinger, L.; Huettel, B.; Matzke, A.J.M. RNA-mediated chromatin-based silencing in plants. Curr. Opin. Cell. Biol. 2009, 21, 357–376.
  80. Piergentili, R. Multiple roles of the Y chromosome in the biology of Drosophila melanogaster. Sci. World J. 2010, 10, 1749–1767.
  81. Steinemann, S.; Steinemann, M. Y chromosomes: Born to be destroyed. BioEssays 2005, 27, 1076–1083.
  82. Kharchenko, P.V.; Xi, R.; Park, P.J. Evidence for dosage compensation between the X chromosome and autosomes in mammals. Nat. Genet. 2011, 43, 1167–1169.
  83. Na, J.K.; Wang, J.; Ming, R. Accumulation of interspersed and sex-specific repeats in the non-recombining region of papaya sex chromosomes. BMC Genom. 2014, 15, 335.
  84. Presting, G.G.; Malysheva, L.; Fuchs, J.; Schubert, I. A TY3/GYPSY retrotransposon-like sequence localizes to the centromeric regions of cereal chromosomes. Plant J. 1998, 16, 721–728.
  85. Tubio, J.M.C.; Tojo, M.; Bassaganyas, L.; Escaramis, G.; Sharakhov, I.V.; Sharakhova, M.V.; Tornador, C.; Unger, M.F.; Naveira, H.; Costas, J.; et al. Evolutionary dynamics of the Ty3/gypsy LTR retrotransposons in the genome of Anopheles gambiae. PLoS ONE 2011, 6, e16328.
  86. Kralova, T.; Cegan, R.; Kubat, Z.; Vrana, J.; Vyskot, B.; Vogel, I.; Kejnovsky, E.; Hobza, R. Identification of a novel retrotransposon with sex chromosome-specific distribution in Silene latifolia. Cytogenet. Genome Res. 2014, 143, 87–95.
  87. Li, N.; Li, X.; Zhou, J.; Yu, L.A.; Li, S.; Zhang, Y.; Qin, R.; Gao, W.; Deng, C. Genome-wide analysis of transposable elements and satellite DNAs in Spinacia species to shed light on their roles in sex chromosome evolution. Front. Plant Sci. 2021, 11, 575462.
  88. Xiong, Y.; Eickbush, T.H. Origin and evolution of retroelements based upon their reverse transcriptase sequences. EMBO J. 1990, 9, 3353–3362.
  89. Lerat, E.; Capy, P. Retrotransposons and retroviruses: Analysis of the envelope gene. Mol. Biol. Evol. 1999, 16, 1198–1207.
  90. Pantazidis, A.; Labrador, M.; Fontdevila, A. The retrotransposon Osvaldo from D. buzzatii displays all structural features of functional retrovirus. Mol. Biol. Evol. 1999, 16, 909–921.
  91. Peterson-Burch, B.D.; Wright, D.A.; Laten, H.M.; Voytas, D.F. Retroviruses in plants? Trends Genet. 2000, 16, 151–152.
  92. Terzian, C.; Ferraz, C.; Demaille, J.; Bucheton, A. Evolution of the gypsy endogenous retrovirus in the Drosophila melanogaster subgroup. Mol. Biol. Evol. 2000, 17, 908–914.
  93. Gonzalez, P.; Lessios, H.A. Evolution of sea urchin retroviral-like (SURL) elements: Evidence from 40 echinoid species. Mol. Biol. Evol. 1999, 16, 938–952.
  94. Malik, H.S.; Eickbush, T.H. Modular evolution of the integrase domain in the Ty3/Gypsy class of LTR retrotransposons. J. Virol. 1999, 73, 5186–5189.
  95. Temminck and Schlegel . 1850. Available online: https://www.gbif.org/species/2407607 (accessed on 5 January 2022).
  96. Poulter, R.; Butler, M.; Ormandy, J. A LINE element from the pufferfish (fugu) Fugu rubripes which shows similarities to the CR1 family of non-LTR retrotransposons. Gene 1998, 227, 169–179.
  97. Volff, J.N.; Körting, C.; Altschmied, J.; Duschl, J.; Sweeney, K.; Wichert, K.; Froschauer, A.; Schartl, M. Jule from the fish Xiphophorus is the first complete vertebrate Ty3/Gypsy retrotransposon from the Mag family. Mol. Biol. Evol. 2001, 18, 101–111.
  98. Tristem, M.; Kabat, P.; Herniou, E.H.; Karpas, A.; Hill, F. Ease1, a Gypsy LTR retrotransposon in the Salmonidae. Mol. Gen. Genet. 1995, 249, 229–236.
  99. Marracci, S.; Batistoni, R.; Pesole, G.; Citti, L.; Nardi, I. Gypsy/Ty3-like elements in the genome of the terrestrial salamander Hydromantes (Amphibia, Urodela). J. Mol. Evol. 1996, 43, 584–593.
  100. Miller, K.; Lynch, C.; Martin, J.; Herniou, E.; Tristem, M. Identification of multiple Gypsy LTR retrotransposons lineages in vertebrate genomes. J. Mol. Evol. 1999, 49, 358–366.
  101. Marín, I.; Lloréns, C. Ty3/Gypsy retrotransposons: Description of new Arabidopsis thaliana elements and evolutionary perspectives derived from comparative genomic data. Mol. Biol. Evol. 2000, 17, 1040–1049.
  102. Herédia, F.; Loreto, E.L.S.; Valente, V.L.S. Complex evolution of gypsy in Drosophilid species. Mol. Biol. Evol. 2004, 21, 1831–1842.
  103. Willd. In: Sp. Pl. 4: 409, 1805. Available online: https://www.gbif.org/species/5328909 (accessed on 5 January 2022).
  104. Puterova, J.; Kubat, Z.; Kejnovsky, E.; Jesionek, W.; Cizkova, J.; Vyskot, B.; Hobza, R. The slowdown of Y chromosome expansion in dioecious Silene latifolia due to DNA loss and male-specific silencing of retrotransposons. BMC Genom. 2018, 19, 153.
  105. Koomgun, T.; Laopichienpong, N.; Singchat, W.; Panthum, T.; Phatcharakullawarawat, R.; Kraichak, E.; Sillapaprayoon, S.; Ahmad, S.F.; Muangmai, N.; Peyachoknagul, S.; et al. Genome complexity reduction high-throughput genome sequencing of green iguana (Iguana iguana) reveal a paradigm shift in understanding sex-chromosomal linkages on homomorphic X and Y sex chromosomes. Front. Genet. 2020, 11, 1217.
  106. Laopichienpong, N.; Kraichak, E.; Singchat, W.; Sillapaprayoon, S.; Muangmai, N.; Suntrarachun, S.; Baicharoen, S.; Peyachok-nagul, S.; Chanhome, L.; Ezaz, T.; et al. Genome-wide SNP analysis of Siamese cobra (Naja kaouthia) reveals the molecular basis of transitions between Z and W sex chromosomes and supports the presence of an ancestral super-sex chromosome in amniotes. Genomics 2021, 113, 624–636.
  107. Nguyen, D.H.M.; Panthum, T.; Ponjarat, J.; Laopichiengpong, N.; Kraichak, E.; Singchat, W.; Muangmai, N.; Peyachoknagul, S.; Na-Nakorn, U.; Srikulnath, K. An investigation of ZZ/ZW and XX/XY sex determination systems in North African catfish (Clarias gariepinus, Burchell 1822). Front. Genet. 2021, 11, 1719.
  108. Nguyen, D.H.M.; Ponjarat, J.; Laopichienpong, N.; Kraichak, E.; Panthum, T.; Singchat, W.; Ahmad, S.F.; Muangmai, N.; Duengkae, P.; Peyachoknagul, S.; et al. Genome-wide SNP analysis suggests male heterogamety in bighead catfish (Clarias macrocephalus). Aquaculture 2021, 545, 737005.
  109. Nguyen, D.H.M.; Ponjarat, J.; Laopichienpong, N.; Panthum, T.; Singchat, W.; Ahmad, S.F.; Kraichak, E.; Muangmai, N.; Duengkae, P.; Peyachoknagul, S.; et al. Genome-wide SNP analysis of hybrid clariid fish reflects the existence of polygenic sex-determination in the lineage. Front. Genet. 2022, 13, 789573.
  110. Panthum, T.; Laopichienpong, N.; Kraichak, E.; Singchat, W.; Ho My Nguyen, D.; Ariyaraphong, N.; Ahmad, S.F.; Muangmai, N.; Duengkae, P.; Peyachoknagul, S.; et al. The snakeskin gourami (Trichopodus pectoralis) tends to exhibit XX/XY sex determination. Fishes 2021, 6, 43.
  111. Suntronpong, A.; Panthum, T.; Laopichienpong, N.; Nguyen, D.H.M.; Kraichak, E.; Singchat, W.; Ariyaraphong, N.; Ahmad, S.F.; Muangmai, N.; Duengkae, P.; et al. Implications of genome-wide single nucleotide polymorphisms in jade perch (Scortum barcoo) reveals the putative XX/XY sex-determination system, facilitating a new chapter of sex control in aquaculture. Aquaculture 2022, 548, 737587.
  112. Bachtrog, D.; Mank, J.E.; Peichel, C.L.; Kirkpatrick, M.; Otto, S.P.; Ashman, T.L.; Hahn, M.W.; Kitano, J.; Mayrose, I.; Ming, R.; et al. Sex determination: Why so many ways of doing it? PLoS Biol. 2014, 12, e1001899.
  113. Hua-Van, A.; Le Rouzic, A.; Maisonhaute, C.; Capy, P. Abundance, distribution and dynamics of retrotransposable elements and transposons: Similarities and differences. Cytogenet. Genome Res. 2005, 110, 426–440.
  114. Mourier, T.; Willerslev, E. Large-scale transcriptome data reveals transcriptional activity of fission yeast LTR retrotransposons. BMC Genom. 2010, 11, 167.
  115. Nuzhdin, S.V.; Mackay, T. The genomic rate of transposable element movement in Drosophila melanogaster. Mol. Biol. Evol. 1995, 12, 180–181.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , ,
View Times: 406
Revisions: 3 times (View History)
Update Date: 23 May 2022
1000/1000