Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Transposable elements shape the evolution of mammalian development

Abstract

Transposable elements (TEs) promote genetic innovation but also threaten genome stability. Despite multiple layers of host defence, TEs actively shape mammalian-specific developmental processes, particularly during pre-implantation and extra-embryonic development and at the maternal–fetal interface. Here, we review how TEs influence mammalian genomes both directly by providing the raw material for genetic change and indirectly via co-evolving TE-binding Krüppel-associated box zinc finger proteins (KRAB-ZFPs). Throughout mammalian evolution, individual activities of ancient TEs were co-opted to enable invasive placentation that characterizes live-born mammals. By contrast, the widespread activity of evolutionarily young TEs may reflect an ongoing co-evolution that continues to impact mammalian development.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Evolution of mammalian live birth coincided with novel TE insertions and the co-evolution and conservation of abundant TE-binding KRAB-ZFPs.
Fig. 2: Species-specific TEs and KRAB-ZFPs promote the rapid evolution of mammalian pre-implantation development.
Fig. 3: Factors that regulate TEs adopted essential roles during early post-implantation development.
Fig. 4: Co-option of TEs and regulatory mechanisms at the maternal–fetal interface.

Similar content being viewed by others

References

  1. Lander, E. S. et al. Initial sequencing and analysis of the human genome. Nature 409, 860–921 (2001).

    Article  CAS  PubMed  Google Scholar 

  2. Chimpanzee Sequencing and Analysis Consortium. Initial sequence of the chimpanzee genome and comparison with the human genome. Nature 437, 69–87 (2005).

    Article  CAS  Google Scholar 

  3. Waterston, R. H. et al. Initial sequencing and comparative analysis of the mouse genome. Nature 420, 520–562 (2002).

    Article  CAS  PubMed  Google Scholar 

  4. Long, H. K., Prescott, S. L. & Wysocka, J. Ever-changing landscapes: transcriptional enhancers in development and evolution. Cell 167, 1170–1187 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Wells, J. N. & Feschotte, C. A field guide to eukaryotic transposable elements. Annu. Rev. Genet. 54, 539-561 (2020). This article is a comprehensive introduction to features and distribution of major TE groups across eukaryote genomes in the context of their biology and evolution.

  6. Smit, A. F. A., Hubley, R. & Green, P. RepeatMasker Open-4.0 (2013–2015).

  7. Chuong, E. B., Rumi, M. A., Soares, M. J. & Baker, J. C. Endogenous retroviruses function as species-specific enhancer elements in the placenta. Nat. Genet. 45, 325–329 (2013). This study establishes that genome-wide species-specific gene regulatory networks in mouse TSCs rely on ERV-derived sequences, using the example of RLTR13D5 elements that provide binding sites for essential placental transcription factors.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Lynch, V. J. et al. Ancient transposable elements transformed the uterine regulatory landscape and transcriptome during the evolution of mammalian pregnancy. Cell Rep. 10, 551–561 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Lynch, V. J., Leclerc, R. D., May, G. & Wagner, G. P. Transposon-mediated rewiring of gene regulatory networks contributed to the evolution of pregnancy in mammals. Nat. Genet. 43, 1154–1159 (2011). Using transcriptional profiling across mammalian endometrial cells, this work demonstrates that genes expressed in human endometrium gained endometrial expression through eutherian-specific MER20 elements that provide binding sites for transcription factors important for hormone responsiveness and pregnancy.

    Article  CAS  PubMed  Google Scholar 

  10. Trizzino, M. et al. Transposable elements are the primary source of novelty in primate gene regulation. Genome Res. 27, 1623–1633 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Chuong, E. B., Elde, N. C. & Feschotte, C. Regulatory evolution of innate immunity through co-option of endogenous retroviruses. Science 351, 1083–1087 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Bourque, G. et al. Ten things you should know about transposable elements. Genome Biol. 19, 199 (2018). This article is a useful primer by leaders in the field on the importance of mammalian TEs for genome evolution and function that introduces their properties and interactions with host cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Kazazian, H. H. Jr. & Moran, J. V. Mobile DNA in health and disease. N. Engl. J. Med. 377, 361–370 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Deniz, O., Frost, J. M. & Branco, M. R. Regulation of transposable elements by DNA modifications. Nat. Rev. Genet. 20, 417–431 (2019).

    Article  CAS  PubMed  Google Scholar 

  15. Cosby, R. L., Chang, N. C. & Feschotte, C. Host–transposon interactions: conflict, cooperation, and cooption. Genes Dev. 33, 1098–1116 (2019). This study uncovers ~100 fusion proteins combining DNA transposase and host-derived domains in tetrapod genomes, suggesting that novel mammalian transcriptional regulators can arise via this process, which is confirmed for the bat fusion protein KRABINER.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Bruno, M., Mahgoub, M. & Macfarlan, T. S. The arms race between KRAB-zinc finger proteins and endogenous retroelements and its impact on mammals. Annu. Rev. Genet. 53, 393–416 (2019).

    Article  CAS  PubMed  Google Scholar 

  17. Imbeault, M., Helleboid, P. Y. & Trono, D. KRAB zinc-finger proteins contribute to the evolution of gene regulatory networks. Nature 543, 550–554 (2017). By identifying KRAB-ZFP genes in extant vertebrate genomes, performing extensive phylogenetic analysis of these KRAB-ZFPs that allowed dating their evolutionary origin and determining the binding sites of the majority of human KRAB-ZFPs, this study reveals the extent to which the KRAB-ZFP gene family interacts with TEs across vertebrate evolution.

    Article  CAS  PubMed  Google Scholar 

  18. Thomas, J. H. & Schneider, S. Coevolution of retroelements and tandem zinc finger genes. Genome Res. 21, 1800–1812 (2011). This work presents compelling evidence that infiltration with LTR retroelements drives the divergence of tandem zinc finger proteins (which include many KRAB-ZFPs) in vertebrate genomes by showing that the number of tandem zinc finger genes and LTR retroviral elements correlate.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Ecco, G., Imbeault, M. & Trono, D. KRAB zinc finger proteins. Development 144, 2719–2729 (2017).

    Article  CAS  PubMed  Google Scholar 

  20. Sheng, G. & Foley, A. C. Diversification and conservation of the extraembryonic tissues in mediating nutrient uptake during amniote development. Ann. N. Y. Acad. Sci. 1271, 97–103 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Rodriguez-Terrones, D. & Torres-Padilla, M. E. Nimble and ready to mingle: transposon outbursts of early development. Trends Genet. 34, 806–820 (2018).

    Article  CAS  PubMed  Google Scholar 

  22. Rowe, H. M. & Trono, D. Dynamic control of endogenous retroviruses during development. Virology 411, 273–287 (2011).

    Article  CAS  PubMed  Google Scholar 

  23. Loebel, D. A., Tsoi, B., Wong, N., O’Rourke, M. P. & Tam, P. P. Restricted expression of ETn-related sequences during post-implantation mouse development. Gene Expr. Patterns 4, 467–471 (2004).

    Article  CAS  PubMed  Google Scholar 

  24. Macfarlan, T. S. et al. Embryonic stem cell potency fluctuates with endogenous retrovirus activity. Nature 487, 57–63 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Boroviak, T. et al. Single cell transcriptome analysis of human, marmoset and mouse embryos reveals common and divergent features of preimplantation development. Development 145, dev167833 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Wolf, G. et al. KRAB-zinc finger protein gene expansion in response to active retrotransposons in the murine lineage. eLife 9, e56337 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Grow, E. J. et al. Intrinsic retroviral reactivation in human preimplantation embryos and pluripotent cells. Nature 522, 221–225 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Liu, D. et al. Single-cell RNA-sequencing reveals the existence of naive and primed pluripotency in pre-implantation rhesus monkey embryos. Genome Res. 28, 1481–1493 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Johnson, W. E. & Coffin, J. M. Constructing primate phylogenies from ancient retrovirus sequences. Proc. Natl Acad. Sci. USA 96, 10254–10260 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Sundaram, V. et al. Widespread contribution of transposable elements to the innovation of gene regulatory networks. Genome Res. 24, 1963–1976 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Goke, J. et al. Dynamic transcription of distinct classes of endogenous retroviral elements marks specific populations of early human embryonic cells. Cell Stem Cell 16, 135–141 (2015).

    Article  CAS  PubMed  Google Scholar 

  32. Colombo, A. R., Triche, T. Jr. & Ramsingh, G. Transposable element expression in acute myeloid leukemia transcriptome and prognosis. Sci. Rep. 8, 16449 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Kong, Y. et al. Transposable element expression in tumors is associated with immune infiltration and increased antigenicity. Nat. Commun. 10, 5228 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Tam, O. H., Ostrow, L. W. & Gale Hammell, M. Diseases of the nERVous system: retrotransposon activity in neurodegenerative disease. Mob. DNA 10, 32 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Babaian, A. & Mager, D. L. Endogenous retroviral promoter exaptation in human cancer. Mob. DNA 7, 24 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  36. Pasquesi, G. I. M. et al. Vertebrate lineages exhibit diverse patterns of transposable element regulation and expression across tissues. Genome Biol. Evol. 12, 506–521 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Goerner-Potvin, P. & Bourque, G. Computational tools to unmask transposable elements. Nat. Rev. Genet. 19, 688–704 (2018).

    Article  CAS  PubMed  Google Scholar 

  38. Todd, C. D., Deniz, O., Taylor, D. & Branco, M. R. Functional evaluation of transposable elements as enhancers in mouse embryonic and trophoblast stem cells. eLife 8, e44344 (2019). This study uses individual genetic deletions of candidate ERV-derived regulatory elements in mouse embryonic and TSCs, and a broader CRISPR interference screen that results in minimal changes in the expression of nearby genes, to demonstrate that many TE-derived enhancers do not influence nearby gene expression.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Fuentes, D. R., Swigut, T. & Wysocka, J. Systematic perturbation of retroviral LTRs reveals widespread long-range effects on human gene regulation. eLife 7, e35989 (2018). This paper uses an elegant novel CRISPR activation/interference screen in human EC cells to reveal significant contributions of hundreds of ape-specific LH5HS elements to the regulation of nearby genes.

    Article  PubMed  PubMed Central  Google Scholar 

  40. McDole, K. et al. In toto imaging and reconstruction of post-implantation mouse development at the single-cell level. Cell Level. Cell 175, 859–876.e33 (2018).

    CAS  PubMed  Google Scholar 

  41. Pijuan-Sala, B., Guibentif, C. & Gottgens, B. Single-cell transcriptional profiling: a window into embryonic cell-type specification. Nat. Rev. Mol. Cell Biol. 19, 399–412 (2018).

    Article  CAS  PubMed  Google Scholar 

  42. Frank, J. A. & Feschotte, C. Co-option of endogenous viral sequences for host cell function. Curr. Opin. Virol. 25, 81–89 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Hancks, D. C. & Kazazian, H. H. Jr. Roles for retrotransposon insertions in human disease. Mob. DNA 7, 9 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Payer, L. M. & Burns, K. H. Transposable elements in human genetic disease. Nat. Rev. Genet. 20, 760–772 (2019).

    Article  CAS  PubMed  Google Scholar 

  45. Gilbert, N., Lutz-Prigge, S. & Moran, J. V. Genomic deletions created upon LINE-1 retrotransposition. Cell 110, 315–325 (2002).

    Article  CAS  PubMed  Google Scholar 

  46. Rowe, H. M. et al. KAP1 controls endogenous retroviruses in embryonic stem cells. Nature 463, 237–240 (2010). This work first identifies the critical role of the KRAB zinc finger co-repressor KAP1 in repressing IAP elements and many other TEs in mouse embryos and ESCs.

    Article  CAS  PubMed  Google Scholar 

  47. Helleboid, P. Y. et al. The interactome of KRAB zinc finger proteins reveals the evolutionary history of their functional diversification. EMBO J. 38, e101220 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Stoll, G. A. et al. Structure of KAP1 tripartite motif identifies molecular interfaces required for retroelement silencing. Proc. Natl Acad. Sci. USA 116, 15042–15051 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Dodge, J. E., Kang, Y. K., Beppu, H., Lei, H. & Li, E. Histone H3-K9 methyltransferase ESET is essential for early development. Mol. Cell Biol. 24, 2478–2486 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Schultz, D. C., Ayyanathan, K., Negorev, D., Maul, G. G. & Rauscher, F. J. 3rd SETDB1: a novel KAP-1-associated histone H3, lysine 9-specific methyltransferase that contributes to HP1-mediated silencing of euchromatic genes by KRAB zinc-finger proteins. Genes Dev. 16, 919–932 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Quenneville, S. et al. The KRAB-ZFP/KAP1 system contributes to the early embryonic establishment of site-specific DNA methylation patterns maintained during development. Cell Rep. 2, 766–773 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Walsh, C. P., Chaillet, J. R. & Bestor, T. H. Transcription of IAP endogenous retroviruses is constrained by cytosine methylation. Nat. Genet. 20, 116–117 (1998). This paper first describes the role of DNA methylation via the maintenance DNA methyltransferase DNMT1 in the repression of TEs (IAP elements) in mouse embryos.

    Article  CAS  PubMed  Google Scholar 

  53. Smith, Z. D. & Meissner, A. DNA methylation: roles in mammalian development. Nat. Rev. Genet. 14, 204–220 (2013).

    Article  CAS  PubMed  Google Scholar 

  54. Jönsson, M. E. et al. Activation of neuronal genes via LINE-1 elements upon global DNA demethylation in human neural progenitors. Nat. Commun. 10, 3182 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  55. Bestor, T. H., Edwards, J. R. & Boulard, M. Notes on the role of dynamic DNA methylation in mammalian development. Proc. Natl Acad. Sci. USA 112, 6796–6799 (2015).

    Article  CAS  PubMed  Google Scholar 

  56. Deaton, A. M. & Bird, A. CpG islands and the regulation of transcription. Genes Dev. 25, 1010–1022 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Rosic, S. et al. Evolutionary analysis indicates that DNA alkylation damage is a byproduct of cytosine DNA methyltransferase activity. Nat. Genet. 50, 452–459 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Quenneville, S. et al. In embryonic stem cells, ZFP57/KAP1 recognize a methylated hexanucleotide to affect chromatin and DNA methylation of imprinting control regions. Mol. Cell 44, 361–372 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Boulard, M., Rucli, S., Edwards, J. R. & Bestor, T. H. Methylation-directed glycosylation of chromatin factors represses retrotransposon promoters. Proc. Natl Acad. Sci. USA 117, 14292–14298 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Liu, H., Chang, L. H., Sun, Y., Lu, X. & Stubbs, L. Deep vertebrate roots for mammalian zinc finger transcription factor subfamilies. Genome Biol. Evol. 6, 510–525 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Shi, H. et al. ZFP57 regulation of transposable elements and gene expression within and beyond imprinted domains. Epigenetics Chromatin 12, 49 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Garcia-Garcia, M. J., Shibata, M. & Anderson, K. V. Chato, a KRAB zinc-finger protein, regulates convergent extension in the mouse embryo. Development 135, 3053–3062 (2008).

    Article  CAS  PubMed  Google Scholar 

  63. Li, X. et al. A maternal-zygotic effect gene, Zfp57, maintains both maternal and paternal imprints. Dev. Cell 15, 547–557 (2008). This study first identifies the KRAB-ZFP ZFP57 as essential for maintaining the mono-allelic expression of many imprinted genes in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Casademunt, E. et al. The zinc finger protein NRIF interacts with the neurotrophin receptor p75(NTR) and participates in programmed cell death. EMBO J. 18, 6050–6061 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Eszter Posfai, J. P. S. Defining totipotency using criteria of increasing stringency. biorxiv https://doi.org/10.1101/2020.03.02.972893 (2020).

  66. Torres-Padilla, M. E. On transposons and totipotency. Philos. Trans. R. Soc. Lond. B Biol. Sci. 375, 20190339 (2020). This review highlights and critically assesses the diverse roles of species-specific TEs in totipotency (part of a great collection of research and review articles on TEs based on the Royal Society discussion meeting ‘Crossroads between Transposons and Gene Rregulation’ held in May 2019).

    Article  PubMed  PubMed Central  Google Scholar 

  67. De Iaco, A. et al. DUX-family transcription factors regulate zygotic genome activation in placental mammals. Nat. Genet. 49, 941–945 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. Leidenroth, A. et al. Evolution of DUX gene macrosatellites in placental mammals. Chromosoma 121, 489–497 (2012).

    Article  CAS  PubMed  Google Scholar 

  69. De Iaco, A., Verp, S., Offner, S., Grun, D. & Trono, D. DUX is a non-essential synchronizer of zygotic genome activation. Development 147, dev177725 (2020).

    PubMed  Google Scholar 

  70. Jachowicz, J. W. et al. LINE-1 activation after fertilization regulates global chromatin accessibility in the early mouse embryo. Nat. Genet. 49, 1502–1510 (2017). Using TALE-based transcriptional silencing to modulate L1 transcription, this study finds that modulating L1 in zygotes (but not at the late two-cell stage) affects genome-wide chromatin accessibility and reduces progression to the blastocyst stage, suggesting an important role of L1 in mammalian pre-implantation development.

    Article  CAS  PubMed  Google Scholar 

  71. Hendrickson, P. G. et al. Conserved roles of mouse DUX and human DUX4 in activating cleavage-stage genes and MERVL/HERVL retrotransposons. Nat. Genet. 49, 925–934 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Andrew Modzelewski, W. S. A species-specific retrotransposon drives a conserved Cdk2ap1 isoform essential for preimplantation development. biorxiv https://doi.org/10.1101/2021.03.24.436683 (2021).

  73. Percharde, M. et al. A LINE1–nucleolin partnership regulates early development and ESC identity. Cell 174, 391–405.e19 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Stojic, L. et al. Specificity of RNAi, LNA and CRISPRi as loss-of-function methods in transcriptional analysis. Nucleic Acids Res. 46, 5950–5966 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Bassett, A. R. et al. Considerations when investigating lncRNA function in vivo. eLife 3, e03058 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Korotkevich, E. et al. The apical domain is required and sufficient for the first lineage segregation in the mouse embryo. Dev. Cell 40, 235–247.e7 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Nishioka, N. et al. The Hippo signaling pathway components Lats and Yap pattern Tead4 activity to distinguish mouse trophectoderm from inner cell mass. Dev. Cell 16, 398–410 (2009).

    Article  CAS  PubMed  Google Scholar 

  78. Posfai, E. et al. Position- and Hippo signaling-dependent plasticity during lineage segregation in the early mouse embryo. eLife 6, e22906 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  79. Nelson, A. C., Mould, A. W., Bikoff, E. K. & Robertson, E. J. Mapping the chromatin landscape and Blimp1 transcriptional targets that regulate trophoblast differentiation. Sci. Rep. 7, 6793 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  80. Sun, M. A. et al. Deciphering the evolution of the transcriptional and regulatory landscape in human placenta. bioRxiv https://doi.org/10.1101/2020.09.11.289686 (2020).

  81. Furukawa, S., Kuroda, Y. & Sugiyama, A. A comparison of the histological structure of the placenta in experimental animals. J. Toxicol. Pathol. 27, 11–18 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  82. Haig, D. Retroviruses and the placenta. Curr. Biol. 22, R609–R613 (2012). This review suggests that retroviruses evolved placental expression to facilitate retroviral transmission from mother to offspring and from offspring to mother, and that retroviruses and host defence systems contributed to the remarkable diversity among mammalian placentae.

    Article  CAS  PubMed  Google Scholar 

  83. Chazaud, C., Yamanaka, Y., Pawson, T. & Rossant, J. Early lineage segregation between epiblast and primitive endoderm in mouse blastocysts through the Grb2–MAPK pathway. Dev. Cell 10, 615–624 (2006).

    Article  CAS  PubMed  Google Scholar 

  84. Nichols, J. et al. Formation of pluripotent stem cells in the mammalian embryo depends on the POU transcription factor Oct4. Cell 95, 379–391 (1998).

    Article  CAS  PubMed  Google Scholar 

  85. Niwa, H., Miyazaki, J. & Smith, A. G. Quantitative expression of Oct-3/4 defines differentiation, dedifferentiation or self-renewal of ES cells. Nat. Genet. 24, 372–376 (2000).

    Article  CAS  PubMed  Google Scholar 

  86. Kunarso, G. et al. Transposable elements have rewired the core regulatory network of human embryonic stem cells. Nat. Genet. 42, 631–634 (2010).

    Article  CAS  PubMed  Google Scholar 

  87. Bourque, G. et al. Evolution of the mammalian transcription factor binding repertoire via transposable elements. Genome Res. 18, 1752–1762 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Marnetto, D. et al. Evolutionary rewiring of human regulatory networks by waves of genome expansion. Am. J. Hum. Genet. 102, 207–218 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Hermant, C. & Torres-Padilla, M. E. TFs for TEs: the transcription factor repertoire of mammalian transposable elements. Genes Dev. 35, 22–39 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Ganesh, S. & Svoboda, P. Retrotransposon-associated long non-coding RNAs in mice and men. Pflug. Arch. 468, 1049–1060 (2016).

    Article  CAS  Google Scholar 

  91. Koyanagi-Aoi, M. et al. Differentiation-defective phenotypes revealed by large-scale analyses of human pluripotent stem cells. Proc. Natl Acad. Sci. USA 110, 20569–20574 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Lu, X. et al. The retrovirus HERVH is a long noncoding RNA required for human embryonic stem cell identity. Nat. Struct. Mol. Biol. 21, 423–425 (2014).

    Article  CAS  PubMed  Google Scholar 

  93. Wang, J. et al. Primate-specific endogenous retrovirus-driven transcription defines naive-like stem cells. Nature 516, 405–409 (2014).

    Article  CAS  PubMed  Google Scholar 

  94. Takahashi, K. et al. The pluripotent stem cell-specific transcript ESRG is dispensable for human pluripotency. PLoS Genet. 17, e1009587 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Black, S. G. et al. Viral particles of endogenous betaretroviruses are released in the sheep uterus and infect the conceptus trophectoderm in a transspecies embryo transfer model. J. Virol. 84, 9078–9085 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Lewis, S. H., Reynolds-Kohler, C., Fox, H. E. & Nelson, J. A. HIV-1 in trophoblastic and villous Hofbauer cells, and haematological precursors in eight-week fetuses. Lancet 335, 565–568 (1990).

    Article  CAS  PubMed  Google Scholar 

  97. Bianchi, D. W. The inadvertent discovery of human fetal cell microchimerism. Clin. Chem. 64, 1400–1401 (2018).

    Article  CAS  PubMed  Google Scholar 

  98. Haig, D. Transposable elements: self-seekers of the germline, team-players of the soma. Bioessays 38, 1158–1166 (2016).

    Article  CAS  PubMed  Google Scholar 

  99. Smith, Z. D. et al. Epigenetic restriction of extraembryonic lineages mirrors the somatic transition to cancer. Nature 549, 543–547 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  100. Corsinotti, A. et al. Global and stage specific patterns of Krüppel-associated-box zinc finger protein gene expression in murine early embryonic cells. PLoS ONE 8, e56721 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Rowe, H. M. et al. TRIM28 repression of retrotransposon-based enhancers is necessary to preserve transcriptional dynamics in embryonic stem cells. Genome Res. 23, 452–461 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Pontis, J. et al. Hominoid-specific transposable elements and KZFPs facilitate human embryonic genome activation and control transcription in naive human ESCs. Cell Stem Cell 24, 724–735.e5 (2019). Based on the stage-specific repression of SVA-derived neuronal enhancers by hominoid-specific KRAB-ZFPs, this study suggests that KRAB-ZFPs may allow the tissue-specific expression of TE-derived regulatory elements, which may facilitate their co-option by repressing immediate early, potentially toxic effects of TEs early in development.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Roussigne, M. et al. The THAP domain: a novel protein motif with similarity to the DNA-binding domain of P element transposase. Trends Biochem. Sci. 28, 66–69 (2003).

    Article  CAS  PubMed  Google Scholar 

  104. Clouaire, T. et al. The THAP domain of THAP1 is a large C2CH module with zinc-dependent sequence-specific DNA-binding activity. Proc. Natl Acad. Sci. USA 102, 6907–6912 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Parker, J. B., Yin, H., Vinckevicius, A. & Chakravarti, D. Host cell factor-1 recruitment to E2F-bound and cell-cycle-control genes is mediated by THAP11 and ZNF143. Cell Rep. 9, 967–982 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Parker, J. B., Palchaudhuri, S., Yin, H., Wei, J. & Chakravarti, D. A transcriptional regulatory role of the THAP11–HCF-1 complex in colon cancer cell function. Mol. Cell Biol. 32, 1654–1670 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Campagne, S., Saurel, O., Gervais, V. & Milon, A. Structural determinants of specific DNA-recognition by the THAP zinc finger. Nucleic Acids Res. 38, 3466–3476 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Bessière, D. et al. Structure-function analysis of the THAP zinc finger of THAP1, a large C2CH DNA-binding module linked to Rb/E2F pathways. J. Biol. Chem. 283, 4352–4363 (2008).

    Article  PubMed  Google Scholar 

  109. Dejosez, M. et al. Ronin is essential for embryogenesis and the pluripotency of mouse embryonic stem cells. Cell 133, 1162–1174 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Durruthy-Durruthy, J. et al. Spatiotemporal reconstruction of the human blastocyst by single-cell gene-expression analysis informs induction of naive pluripotency. Dev. Cell 38, 100–115 (2016).

    Article  CAS  PubMed  Google Scholar 

  111. Dejosez, M. et al. Ronin/Hcf-1 binds to a hyperconserved enhancer element and regulates genes involved in the growth of embryonic stem cells. Genes Dev. 24, 1479–1484 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Michaud, J. et al. HCFC1 is a common component of active human CpG-island promoters and coincides with ZNF143, THAP11, YY1, and GABP transcription factor occupancy. Genome Res. 23, 907–916 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Mazars, R. et al. The THAP-zinc finger protein THAP1 associates with coactivator HCF-1 and O-GlcNAc transferase: a link between DYT6 and DYT3 dystonias. J. Biol. Chem. 285, 13364–13371 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Cosby, R. L. et al. Recurrent evolution of vertebrate transcription factors by transposase capture. Science 371, eabc6405 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Arnold, S. J. & Robertson, E. J. Making a commitment: cell lineage allocation and axis patterning in the early mouse embryo. Nat. Rev. Mol. Cell Biol. 10, 91–103 (2009).

    Article  CAS  PubMed  Google Scholar 

  116. Bardot, E. S. & Hadjantonakis, A. K. Mouse gastrulation: coordination of tissue patterning, specification and diversification of cell fate. Mech. Dev. 163, 103617 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Svoboda, P. et al. RNAi and expression of retrotransposons MuERV-L and IAP in preimplantation mouse embryos. Dev. Biol. 269, 276–285 (2004).

    Article  CAS  PubMed  Google Scholar 

  118. Piko, L., Hammons, M. D. & Taylor, K. D. Amounts, synthesis, and some properties of intracisternal A particle-related RNA in early mouse embryos. Proc. Natl Acad. Sci. USA 81, 488–492 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Poznanski, A. A. & Calarco, P. G. The expression of intracisternal A particle genes in the preimplantation mouse embryo. Dev. Biol. 143, 271–281 (1991).

    Article  CAS  PubMed  Google Scholar 

  120. Auclair, G., Guibert, S., Bender, A. & Weber, M. Ontogeny of CpG island methylation and specificity of DNMT3 methyltransferases during embryonic development in the mouse. Genome Biol. 15, 545 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Chapman, V., Forrester, L., Sanford, J., Hastie, N. & Rossant, J. Cell lineage-specific undermethylation of mouse repetitive DNA. Nature 307, 284–286 (1984).

    Article  CAS  PubMed  Google Scholar 

  122. Brulet, P., Kaghad, M., Xu, Y. S., Croissant, O. & Jacob, F. Early differential tissue expression of transposon-like repetitive DNA sequences of the mouse. Proc. Natl Acad. Sci. USA 80, 5641–5645 (1983).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Maksakova, I. A. & Mager, D. L. Transcriptional regulation of early transposon elements, an active family of mouse long terminal repeat retrotransposons. J. Virol. 79, 13865–13874 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. O’Donnell, K. A., An, W., Schrum, C. T., Wheelan, S. J. & Boeke, J. D. Controlled insertional mutagenesis using a LINE-1 (ORFeus) gene-trap mouse model. Proc. Natl Acad. Sci. USA 110, E2706–E2713 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  125. Wallace, N. A., Belancio, V. P. & Deininger, P. L. L1 mobile element expression causes multiple types of toxicity. Gene 419, 75–81 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Cammas, F. et al. Mice lacking the transcriptional corepressor TIF1β are defective in early postimplantation development. Development 127, 2955–2963 (2000).

    Article  CAS  PubMed  Google Scholar 

  127. Matsui, T. et al. Proviral silencing in embryonic stem cells requires the histone methyltransferase ESET. Nature 464, 927–931 (2010).

    Article  CAS  PubMed  Google Scholar 

  128. Dai, H. Q. et al. TET-mediated DNA demethylation controls gastrulation by regulating Lefty–Nodal signalling. Nature 538, 528–532 (2016).

    Article  PubMed  CAS  Google Scholar 

  129. Harten, S. K. et al. The first mouse mutants of D14Abb1e (Fam208a) show that it is critical for early development. Mamm. Genome 25, 293–303 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Soehn, A. S. et al. Periphilin is strongly expressed in the murine nervous system and is indispensable for murine development. Genesis 47, 697–707 (2009).

    Article  CAS  PubMed  Google Scholar 

  131. Zhu, Y., Wang, G. Z., Cingöz, O. & Goff, S. P. NP220 mediates silencing of unintegrated retroviral DNA. Nature 564, 278–282 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Robbez-Masson, L. et al. The HUSH complex cooperates with TRIM28 to repress young retrotransposons and new genes. Genome Res. 28, 836–845 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Rossant, J., Sanford, J. P., Chapman, V. M. & Andrews, G. K. Undermethylation of structural gene sequences in extraembryonic lineages of the mouse. Dev. Biol. 117, 567–573 (1986).

    Article  CAS  PubMed  Google Scholar 

  134. Yuan, P. et al. Eset partners with Oct4 to restrict extraembryonic trophoblast lineage potential in embryonic stem cells. Genes Dev. 23, 2507–2520 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Hemberger, M. Epigenetic landscape required for placental development. Cell Mol. Life Sci. 64, 2422–2436 (2007).

    Article  CAS  PubMed  Google Scholar 

  136. Hemberger, M. Genetic–epigenetic intersection in trophoblast differentiation: implications for extraembryonic tissue function. Epigenetics 5, 24–29 (2010).

    Article  CAS  PubMed  Google Scholar 

  137. Yang, X. et al. Nuclear reprogramming of cloned embryos and its implications for therapeutic cloning. Nat. Genet. 39, 295–302 (2007).

    Article  CAS  PubMed  Google Scholar 

  138. Tanaka, S. et al. Placentomegaly in cloned mouse concepti caused by expansion of the spongiotrophoblast layer. Biol. Reprod. 65, 1813–1821 (2001).

    Article  CAS  PubMed  Google Scholar 

  139. Ng, R. K. et al. Epigenetic restriction of embryonic cell lineage fate by methylation of Elf5. Nat. Cell Biol. 10, 1280–1290 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Liu, S. et al. Setdb1 is required for germline development and silencing of H3K9me3-marked endogenous retroviruses in primordial germ cells. Genes Dev. 28, 2041–2055 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Shibata, M., Blauvelt, K. E., Liem, K. F. Jr. & Garcia-Garcia, M. J. TRIM28 is required by the mouse KRAB domain protein ZFP568 to control convergent extension and morphogenesis of extra-embryonic tissues. Development 138, 5333–5343 (2011). This study, carefully characterizing conditional and knockout mouse embryos for Trim28, reveals that the essential role of ZFP568 during mouse gastrulation is partially phenocopied in conditional Trim28 mutants, also establishing that KAP1 (encoded by Trim28) has crucial extra-embryonic roles.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Shibata, M. & Garcia-Garcia, M. J. The mouse KRAB zinc-finger protein CHATO is required in embryonic-derived tissues to control yolk sac and placenta morphogenesis. Dev. Biol. 349, 331–341 (2011).

    Article  CAS  PubMed  Google Scholar 

  143. Yang, P. et al. A placental growth factor is silenced in mouse embryos by the zinc finger protein ZFP568. Science 356, 757–759 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. DeChiara, T. M., Efstratiadis, A. & Robertson, E. J. A growth-deficiency phenotype in heterozygous mice carrying an insulin-like growth factor II gene disrupted by targeting. Nature 345, 78–80 (1990).

    Article  CAS  PubMed  Google Scholar 

  145. Sun, F. L., Dean, W. L., Kelsey, G., Allen, N. D. & Reik, W. Transactivation of Igf2 in a mouse model of Beckwith–Wiedemann syndrome. Nature 389, 809–815 (1997).

    Article  CAS  PubMed  Google Scholar 

  146. Patel, A. et al. DNA conformation induces adaptable binding by tandem zinc finger proteins. Cell 173, 221–233.e12 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Chien, H. C. et al. Targeted disruption in mice of a neural stem cell-maintaining, KRAB-Zn finger-encoding gene that has rapidly evolved in the human lineage. PLoS ONE 7, e47481 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Ben-Haim, N. et al. The nodal precursor acting via activin receptors induces mesoderm by maintaining a source of its convertases and BMP4. Dev. Cell 11, 313–323 (2006).

    Article  CAS  PubMed  Google Scholar 

  149. Robertson, E. J. Dose-dependent Nodal/Smad signals pattern the early mouse embryo. Semin. Cell Dev. Biol. 32, 73–79 (2014).

    Article  CAS  PubMed  Google Scholar 

  150. Beddington, R. S. & Robertson, E. J. Axis development and early asymmetry in mammals. Cell 96, 195–209 (1999).

    Article  CAS  PubMed  Google Scholar 

  151. Hemberger, M., Hanna, C. W. & Dean, W. Mechanisms of early placental development in mouse and humans. Nat. Rev. Genet. 21, 27–43 (2020).

    Article  CAS  PubMed  Google Scholar 

  152. Dupressoir, A. et al. Syncytin-A knockout mice demonstrate the critical role in placentation of a fusogenic, endogenous retrovirus-derived, envelope gene. Proc. Natl Acad. Sci. USA 106, 12127–12132 (2009). Following the initial identification of syncytins in humans and mice in 2000 and 2005, this study first shows that loss of the endogenous retroviral env gene Syncytin-A in mice disables syncytiotrophoblast fusion, leading to embryonic lethality due to inability to establish the maternal–fetal interface.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Dupressoir, A. et al. A pair of co-opted retroviral envelope syncytin genes is required for formation of the two-layered murine placental syncytiotrophoblast. Proc. Natl Acad. Sci. USA 108, E1164–E1173 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Lavialle, C. et al. Paleovirology of ‘syncytins’, retroviral env genes exapted for a role in placentation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 368, 20120507 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  155. Cornelis, G. et al. Retroviral envelope syncytin capture in an ancestrally diverged mammalian clade for placentation in the primitive Afrotherian tenrecs. Proc. Natl Acad. Sci. USA 111, E4332–E4341 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Herve, C. A., Forrest, G., Lower, R., Griffiths, D. J. & Venables, P. J. Conservation and loss of the ERV3 open reading frame in primates. Genomics 83, 940–943 (2004).

    Article  CAS  PubMed  Google Scholar 

  157. de Parseval, N. et al. Comprehensive search for intra- and inter-specific sequence polymorphisms among coding envelope genes of retroviral origin found in the human genome: genes and pseudogenes. BMC Genomics 6, 117 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  158. Mallet, F. et al. The endogenous retroviral locus ERVWE1 is a bona fide gene involved in hominoid placental physiology. Proc. Natl Acad. Sci. USA 101, 1731–1736 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. de Parseval, N. & Heidmann, T. Physiological knockout of the envelope gene of the single-copy ERV-3 human endogenous retrovirus in a fraction of the Caucasian population. J. Virol. 72, 3442–3445 (1998).

    Article  PubMed  PubMed Central  Google Scholar 

  160. Kim, H. S. et al. Human endogenous retrovirus (HERV)-R family in primates: chromosomal location, gene expression, and evolution. Gene 370, 34–42 (2006).

    Article  CAS  PubMed  Google Scholar 

  161. Mangeney, M. et al. Placental syncytins: genetic disjunction between the fusogenic and immunosuppressive activity of retroviral envelope proteins. Proc. Natl Acad. Sci. USA 104, 20534–20539 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Sugimoto, J. et al. Suppressyn localization and dynamic expression patterns in primary human tissues support a physiologic role in human placentation. Sci. Rep. 9, 19502 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Frank, J. A. Antiviral activity of a human placental protein of retroviral origin. biorxiv https://doi.org/10.1101/2020.08.23.263665 (2020).

  164. Sugimoto, J., Sugimoto, M., Bernstein, H., Jinno, Y. & Schust, D. A novel human endogenous retroviral protein inhibits cell–cell fusion. Sci. Rep. 3, 1462 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  165. Buchrieser, J. et al. IFITM proteins inhibit placental syncytiotrophoblast formation and promote fetal demise. Science 365, 176–180 (2019).

    Article  CAS  PubMed  Google Scholar 

  166. Ono, R. et al. Deletion of Peg10, an imprinted gene acquired from a retrotransposon, causes early embryonic lethality. Nat. Genet. 38, 101–106 (2006). This study first demonstrates that the PEG10 gene was co-opted from gag and pol proteins derived from ERVs to perform an essential placental function in mice.

    Article  CAS  PubMed  Google Scholar 

  167. Kaneko-Ishino, T. & Ishino, F. Mammalian-specific genomic functions: newly acquired traits generated by genomic imprinting and LTR retrotransposon-derived genes in mammals. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 91, 511–538 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Youngson, N. A., Kocialkowski, S., Peel, N. & Ferguson-Smith, A. C. A small family of sushi-class retrotransposon-derived genes in mammals and their relation to genomic imprinting. J. Mol. Evol. 61, 481–490 (2005).

    Article  CAS  PubMed  Google Scholar 

  169. Henke, C. et al. Selective expression of sense and antisense transcripts of the sushi-ichi-related retrotransposon-derived family during mouse placentogenesis. Retrovirology 12, 9 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  170. Abed, M. et al. The Gag protein PEG10 binds to RNA and regulates trophoblast stem cell lineage specification. PLoS ONE 14, e0214110 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Koppes, E., Himes, K. P. & Chaillet, J. R. Partial loss of genomic imprinting reveals important roles for Kcnq1 and Peg10 imprinted domains in placental development. PLoS ONE 10, e0135202 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  172. Suzuki, S. et al. Retrotransposon silencing by DNA methylation can drive mammalian genomic imprinting. PLoS Genet. 3, e55 (2007).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  173. Ono, R. et al. A retrotransposon-derived gene, PEG10, is a novel imprinted gene located on human chromosome 7q21. Genomics 73, 232–237 (2001).

    Article  CAS  PubMed  Google Scholar 

  174. Hantak, M. P., Einstein, J., Kearns, R. B. & Shepherd, J. D. Intercellular communication in the nervous system goes viral. Trends Neurosci. 44, 248–259 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Kitazawa, M., Tamura, M., Kaneko-Ishino, T. & Ishino, F. Severe damage to the placental fetal capillary network causes mid- to late fetal lethality and reduction in placental size in Peg11/Rtl1 KO mice. Genes. Cell 22, 174–188 (2017).

    Article  CAS  Google Scholar 

  176. Sekita, Y. et al. Role of retrotransposon-derived imprinted gene, Rtl1, in the feto-maternal interface of mouse placenta. Nat. Genet. 40, 243–248 (2008).

    Article  CAS  PubMed  Google Scholar 

  177. Edwards, C. A. et al. The evolution of the DLK1–DIO3 imprinted domain in mammals. PLoS Biol. 6, e135 (2008).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  178. Haig, D. Maternal–fetal conflict, genomic imprinting and mammalian vulnerabilities to cancer. Philos. Trans. R. Soc. Lond. B Biol. Sci. 370, 20140178 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  179. Haig, D. Going retro: transposable elements, embryonic stem cells, and the mammalian placenta (retrospective on DOI 10.1002/bies.201300059). Bioessays 37, 1154 (2015).

    Article  PubMed  Google Scholar 

  180. McGrath, J. & Solter, D. Completion of mouse embryogenesis requires both the maternal and paternal genomes. Cell 37, 179–183 (1984).

    Article  CAS  PubMed  Google Scholar 

  181. Surani, M. A., Barton, S. C. & Norris, M. L. Nuclear transplantation in the mouse: heritable differences between parental genomes after activation of the embryonic genome. Cell 45, 127–136 (1986).

    Article  CAS  PubMed  Google Scholar 

  182. Kono, T. et al. Birth of parthenogenetic mice that can develop to adulthood. Nature 428, 860–864 (2004).

    Article  CAS  PubMed  Google Scholar 

  183. Reik, W. & Lewis, A. Co-evolution of X-chromosome inactivation and imprinting in mammals. Nat. Rev. Genet. 6, 403–410 (2005).

    Article  CAS  PubMed  Google Scholar 

  184. Hanna, C. W. et al. Endogenous retroviral insertions drive non-canonical imprinting in extra-embryonic tissues. Genome Biol. 20, 225 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  185. Andergassen, D., Smith, Z. D., Rinn, J. L. & Meissner, A. Diverse mechanisms for epigenetic imprinting in mammals. bioRxiv https://doi.org/10.1101/2021.04.30.442087f (2021).

  186. Pask, A. J. et al. Analysis of the platypus genome suggests a transposon origin for mammalian imprinting. Genome Biol. 10, R1 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  187. Juan, A. M. & Bartolomei, M. S. Evolving imprinting control regions: KRAB zinc fingers hold the key. Genes Dev. 33, 1–3 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Takahashi, N. et al. ZNF445 is a primary regulator of genomic imprinting. Genes Dev. 33, 49–54 (2019). This study establishes that the essential KRAB-ZFP ZFP57 pairs with ZNF445 to maintain all but one (Peg10) ICRs in mice.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Alexander, K. A., Wang, X., Shibata, M., Clark, A. G. & García-García, M. J. TRIM28 controls genomic imprinting through distinct mechanisms during and after early genome-wide reprogramming. Cell Rep. 13, 1194–1205 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  190. Mugford, J. W., Yee, D. & Magnuson, T. Failure of extra-embryonic progenitor maintenance in the absence of dosage compensation. Development 139, 2130–2138 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Marahrens, Y., Panning, B., Dausman, J., Strauss, W. & Jaenisch, R. Xist-deficient mice are defective in dosage compensation but not spermatogenesis. Genes Dev. 11, 156–166 (1997).

    Article  CAS  PubMed  Google Scholar 

  192. Lyon, M. F. Gene action in the X-chromosome of the mouse (Mus musculus L.). Nature 190, 372–373 (1961). In this paper, Mary Lyon — who discovered the epigenetic process of X-chromosome inactivation — suggests that LINE elements, which are enriched on mammalian X chromosomes, are important for this process.

    Article  CAS  PubMed  Google Scholar 

  193. Lyon, M. F. X-chromosome inactivation: a repeat hypothesis. Cytogenet. Cell Genet. 80, 133–137 (1998).

    Article  CAS  PubMed  Google Scholar 

  194. Woods, L. et al. Decidualisation and placentation defects are a major cause of age-related reproductive decline. Nat. Commun. 8, 352 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  195. Erkenbrack, E. M. et al. The mammalian decidual cell evolved from a cellular stress response. PLoS Biol. 16, e2005594 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  196. Samstein, R. M., Josefowicz, S. Z., Arvey, A., Treuting, P. M. & Rudensky, A. Y. Extrathymic generation of regulatory T cells in placental mammals mitigates maternal-fetal conflict. Cell 150, 29–38 (2012). This article demonstrates using sophisticated allogenic and syngeneic mouse models that the lack of the eutherian-specific MIR element-derived Foxp3 enhancer plays an essential role in mitigating maternal–fetal conflict, thereby providing a compelling example that a single TE-derived regulatory sequence can be essential for mammalian development.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Ferreira, L. M. et al. A distant trophoblast-specific enhancer controls HLA-G expression at the maternal-fetal interface. Proc. Natl Acad. Sci. USA 113, 5364–5369 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  198. Emera, D. et al. Convergent evolution of endometrial prolactin expression in primates, mice, and elephants through the independent recruitment of transposable elements. Mol. Biol. Evol. 29, 239–247 (2012).

    Article  CAS  PubMed  Google Scholar 

  199. Emera, D. & Wagner, G. P. Transformation of a transposon into a derived prolactin promoter with function during human pregnancy. Proc. Natl Acad. Sci. USA 109, 11246–11251 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  200. McLean, M. et al. A placental clock controlling the length of human pregnancy. Nat. Med. 1, 460–463 (1995).

    Article  CAS  PubMed  Google Scholar 

  201. Dunn-Fletcher, C. E. et al. Anthropoid primate-specific retroviral element THE1B controls expression of CRH in placenta and alters gestation length. PLoS Biol. 16, e2006337 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  202. van de Lagemaat, L. N., Landry, J. R., Mager, D. L. & Medstrand, P. Transposable elements in mammals promote regulatory variation and diversification of genes with specialized functions. Trends Genet. 19, 530–536 (2003).

    Article  PubMed  CAS  Google Scholar 

  203. Nishihara, H. et al. Coordinately co-opted multiple transposable elements constitute an enhancer for wnt5a expression in the mammalian secondary palate. PLoS Genet. 12, e1006380 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  204. Nishihara, H., Smit, A. F. & Okada, N. Functional noncoding sequences derived from SINEs in the mammalian genome. Genome Res. 16, 864–874 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Hirakawa, M., Nishihara, H., Kanehisa, M. & Okada, N. Characterization and evolutionary landscape of AmnSINE1 in amniota genomes. Gene 441, 100–110 (2009).

    Article  CAS  PubMed  Google Scholar 

  206. Griffith, O. W. & Wagner, G. P. The placenta as a model for understanding the origin and evolution of vertebrate organs. Nat. Ecol. Evol. 1, 72 (2017).

    Article  PubMed  Google Scholar 

  207. Sundaram, V. & Wysocka, J. Transposable elements as a potent source of diverse cis-regulatory sequences in mammalian genomes. Philos. Trans. R. Soc. Lond. B Biol. Sci. 375, 20190347 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Dunlap, K. A. et al. Endogenous retroviruses regulate periimplantation placental growth and differentiation. Proc. Natl Acad. Sci. USA 103, 14390–14395 (2006). This study shows that knock down of the sheep env-derived enJSRVs leads to failure to differentiate binucleate trophoblast giant cells and lethality, providing compelling evidence that ERVs in mammals other than mice have co-opted TEs for essential developmental functions.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  209. Perez-Garcia, V. et al. Placentation defects are highly prevalent in embryonic lethal mouse mutants. Nature 555, 463–468 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. D’Souza, A. W. & Wagner, G. P. Malignant cancer and invasive placentation: a case for positive pleiotropy between endometrial and malignancy phenotypes. Evol. Med. Public Health 2014, 136–145 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  211. Kshitiz et al. Evolution of placental invasion and cancer metastasis are causally linked. Nat. Ecol. Evol. 3, 1743–1753 (2019).

    Article  CAS  PubMed  Google Scholar 

  212. Treger, R. S. et al. The lupus susceptibility locus Sgp3 encodes the suppressor of endogenous retrovirus expression SNERV. Immunity 50, 334–347.e9 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Natri, H., Garcia, A. R., Buetow, K. H., Trumble, B. C. & Wilson, M. A. The pregnancy pickle: evolved immune compensation due to pregnancy underlies sex differences in human diseases. Trends Genet. 35, 478–488 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Greenbaum, S. & Greenbaum, G. Evolution of immune sexual dimorphism in response to placental invasiveness: a response to Natri et al. Trends Genet. 36, 3–5 (2019).

    Article  PubMed  CAS  Google Scholar 

  215. Roth, O. et al. Evolution of male pregnancy associated with remodeling of canonical vertebrate immunity in seahorses and pipefishes. Proc. Natl Acad. Sci. USA 117, 9431–9439 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Cornelis, G. et al. An endogenous retroviral envelope syncytin and its cognate receptor identified in the viviparous placental Mabuya lizard. Proc. Natl Acad. Sci. USA 114, E10991–E11000 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Gehring, M. & Satyaki, P. R. Endosperm and imprinting, inextricably linked. Plant. Physiol. 173, 143–154 (2017).

    Article  CAS  PubMed  Google Scholar 

  218. Pignatta, D., Novitzky, K., Satyaki, P. R. V. & Gehring, M. A variably imprinted epiallele impacts seed development. PLoS Genet. 14, e1007469 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  219. Turco, M. Y. et al. Trophoblast organoids as a model for maternal–fetal interactions during human placentation. Nature 564, 263–267 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  220. Turco, M. Y. et al. Long-term, hormone-responsive organoid cultures of human endometrium in a chemically defined medium. Nat. Cell Biol. 19, 568–577 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Mirna Marinic, V. J. L. Derivation of endometrial gland organoids from term post-partum placenta. biorxiv https://doi.org/10.1101/753780v1 (2019).

  222. Beccari, L. et al. Multi-axial self-organization properties of mouse embryonic stem cells into gastruloids. Nature 562, 272–276 (2018).

    Article  CAS  PubMed  Google Scholar 

  223. Yoney, A. et al. WNT signaling memory is required for ACTIVIN to function as a morphogen in human gastruloids. eLife 7, e38279 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  224. Hayashi, K. et al. Offspring from oocytes derived from in vitro primordial germ cell-like cells in mice. Science 338, 971–975 (2012).

    Article  CAS  PubMed  Google Scholar 

  225. Hayashi, K., Ohta, H., Kurimoto, K., Aramaki, S. & Saitou, M. Reconstitution of the mouse germ cell specification pathway in culture by pluripotent stem cells. Cell 146, 519–532 (2011).

    Article  CAS  PubMed  Google Scholar 

  226. Yamashiro, C. et al. Generation of human oogonia from induced pluripotent stem cells in vitro. Science 362, 356–360 (2018).

    Article  CAS  PubMed  Google Scholar 

  227. Okae, H. et al. Derivation of human trophoblast stem cells. Cell Stem Cell 22, 50–63.e6 (2018).

    Article  CAS  PubMed  Google Scholar 

  228. Haider, S. et al. Self-renewing trophoblast organoids recapitulate the developmental program of the early human placenta. Stem Cell Rep. 11, 537–551 (2018).

    Article  CAS  Google Scholar 

  229. Marinic, M., Rana, S. & Lynch, V. J. Derivation of endometrial gland organoids from term placenta. Placenta 101, 75–79 (2020).

    Article  PubMed  CAS  Google Scholar 

  230. Wolf, D. & Goff, S. P. Embryonic stem cells use ZFP809 to silence retroviral DNAs. Nature 458, 1201–1204 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  231. Loebel, D. A., O’Rourke, M. P., Steiner, K. A., Banyer, J. & Tam, P. P. Isolation of differentially expressed genes from wild-type and Twist mutant mouse limb buds. Genesis 33, 103–113 (2002).

    Article  CAS  PubMed  Google Scholar 

  232. Chen, Z. F. & Behringer, R. R. Twist is required in head mesenchyme for cranial neural tube morphogenesis. Genes Dev. 9, 686–699 (1995).

    Article  CAS  PubMed  Google Scholar 

  233. Gray, P. A. et al. Mouse brain organization revealed through direct genome-scale TF expression analysis. Science 306, 2255–2257 (2004).

    Article  CAS  PubMed  Google Scholar 

  234. Lu, X. et al. Twist1 is involved in trophoblast syncytialization by regulating GCM1. Placenta 39, 45–54 (2016).

    Article  CAS  PubMed  Google Scholar 

  235. Kruger, I. et al. Sp1/Sp3 compound heterozygous mice are not viable: impaired erythropoiesis and severe placental defects. Dev. Dyn. 236, 2235–2244 (2007).

    Article  PubMed  CAS  Google Scholar 

  236. Sawyer, M. H., Nachlas, N. E. Jr. & Panem, S. C-type viral antigen expression in human placenta. Nature 275, 62–64 (1978).

    Article  CAS  PubMed  Google Scholar 

  237. Tsumura, A. et al. Maintenance of self-renewal ability of mouse embryonic stem cells in the absence of DNA methyltransferases Dnmt1, Dnmt3a and Dnmt3b. Genes Cell 11, 805–814 (2006).

    Article  CAS  Google Scholar 

  238. Argelaguet, R. et al. Multi-omics profiling of mouse gastrulation at single-cell resolution. Nature 576, 487–491 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  239. Mochizuki, K. et al. SETDB1 is essential for mouse primordial germ cell fate determination by ensuring BMP signaling. Development 145, dev164160 (2018).

    Article  PubMed  CAS  Google Scholar 

  240. Zhu, D. et al. Lysine-specific demethylase 1 regulates differentiation onset and migration of trophoblast stem cells. Nat. Commun. 5, 3174 (2014).

    Article  PubMed  CAS  Google Scholar 

  241. Herzog, M. et al. TIF1β association with HP1 is essential for post-gastrulation development, but not for Sertoli cell functions during spermatogenesis. Dev. Biol. 350, 548–558 (2011).

    Article  CAS  PubMed  Google Scholar 

  242. McClintock, C. B. The origin and behavior of mutable loci in maize. Proc. Natl Acad. Sci. USA 36, 344–355 (1950).

    Article  Google Scholar 

  243. Comfort, N. C. From controlling elements to transposons: Barbara McClintock and the Nobel Prize. Trends Biochem. Sci. 26, 454–457 (2001). This paper is a brief historical perspective on Barbara McClintock’s discovery of TEs in maize.

    Article  CAS  PubMed  Google Scholar 

  244. Feschotte, C. & Pritham, E. J. DNA transposons and the evolution of eukaryotic genomes. Annu. Rev. Genet. 41, 331–368 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  245. Zimmerly, S. & Semper, C. Evolution of group II introns. Mob. DNA 6, 7 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  246. Koonin, E. V. Viruses and mobile elements as drivers of evolutionary transitions. Philos. Trans. R Soc. Lond. B Biol. Sci. 371,, 20150442 (2016).

    Article  CAS  Google Scholar 

  247. Novikova, O. & Belfort, M. Mobile group II introns as ancestral eukaryotic elements. Trends Genet. 33, 773–783 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Kramerov, D. A. & Vassetzky, N. S. Short retroposons in eukaryotic genomes. Int. Rev. Cytol. 247, 165–221 (2005).

    Article  CAS  PubMed  Google Scholar 

  249. Malik, H. S., Burke, W. D. & Eickbush, T. H. The age and evolution of non-LTR retrotransposable elements. Mol. Biol. Evol. 16, 793–805 (1999).

    Article  CAS  PubMed  Google Scholar 

  250. Krupovic, M., Dolja, V. V. & Koonin, E. V. Origin of viruses: primordial replicators recruiting capsids from hosts. Nat. Rev. Microbiol. 17, 449–458 (2019).

    Article  CAS  PubMed  Google Scholar 

  251. Dewannieux, M., Esnault, C. & Heidmann, T. LINE-mediated retrotransposition of marked Alu sequences. Nat. Genet. 35, 41–48 (2003).

    Article  CAS  PubMed  Google Scholar 

  252. Muñoz-Lopez, M., et al. LINE-1 retrotransposition impacts the genome of human pre-implantation embryos and extraembryonic tissues. biorxiv https://www.biorxiv.org/content/10.1101/522623v1 (2019).

  253. Gagnier, L., Belancio, V. P. & Mager, D. L. Mouse germ line mutations due to retrotransposon insertions. Mob. DNA 10, 15 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  254. Richardson, S. R. et al. Heritable L1 retrotransposition in the mouse primordial germline and early embryo. Genome Res. 27, 1395–1405 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  255. Nellaker, C. et al. The genomic landscape shaped by selection on transposable elements across 18 mouse strains. Genome Biol. 13, R45 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  256. Gilbert, C. & Feschotte, C. Horizontal acquisition of transposable elements and viral sequences: patterns and consequences. Curr. Opin. Genet. Dev. 49, 15–24 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  257. Tarlinton, R. E., Meers, J. & Young, P. R. Retroviral invasion of the koala genome. Nature 442, 79–81 (2006).

    Article  CAS  PubMed  Google Scholar 

  258. Birtle, Z. & Ponting, C. P. Meisetz and the birth of the KRAB motif. Bioinformatics 22, 2841–2845 (2006). This study traces the origin of the KRAB motif that characterizes KRAB zinc finger genes to sea urchins and presents evidence that a KRAB motif progenitor was present in the last common ancestor of animals, plants and fungi.

    Article  CAS  PubMed  Google Scholar 

  259. Najafabadi, H. S. et al. C2H2 zinc finger proteins greatly expand the human regulatory lexicon. Nat. Biotechnol. 33, 555–562 (2015).

    Article  CAS  PubMed  Google Scholar 

  260. Jumpei Ito, I. K. et al. Endogenous retroviruses drive KRAB zinc-finger family protein expression for tumor suppression. biorxiv https://doi.org/10.1101/2020.02.02.931501 (2020).

  261. Ivanov, D., Stone, J. R., Maki, J. L., Collins, T. & Wagner, G. Mammalian SCAN domain dimer is a domain-swapped homolog of the HIV capsid C-terminal domain. Mol. Cell 17, 137–143 (2005).

    Article  CAS  PubMed  Google Scholar 

  262. Williams, A. J., Blacklow, S. C. & Collins, T. The zinc finger-associated SCAN box is a conserved oligomerization domain. Mol. Cell Biol. 19, 8526–8535 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  263. Al Chiblak, M., Steinbeck, F., Thiesen, H. J. & Lorenz, P. DUF3669, a “domain of unknown function” within ZNF746 and ZNF777, oligomerizes and contributes to transcriptional repression. BMC Mol. Cell Biol. 20, 60 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  264. Emerson, R. O. & Thomas, J. H. Gypsy and the birth of the SCAN domain. J. Virol. 85, 12043–12052 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  265. Jorgensen, P. Yolk. Curr. Biol. 18, R103–R104 (2008).

    Article  CAS  PubMed  Google Scholar 

  266. Arendt, D. & Nubler-Jung, K. Rearranging gastrulation in the name of yolk: evolution of gastrulation in yolk-rich amniote eggs. Mech. Dev. 81, 3–22 (1999).

    Article  CAS  PubMed  Google Scholar 

  267. Griffith, O. W. et al. Embryo implantation evolved from an ancestral inflammatory attachment reaction. Proc. Natl Acad. Sci. USA 114, E6566–E6575 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  268. Ross, C. & Boroviak, T. E. Origin and function of the yolk sac in primate embryogenesis. Nat. Commun. 11, 3760 (2020).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  269. Pereira, P. N. et al. Amnion formation in the mouse embryo: the single amniochorionic fold model. BMC Dev. Biol. 11, 48 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  270. Renfree, M. B. Review: Marsupials: placental mammals with a difference. Placenta 31 (Suppl.), S21–S26 (2010).

  271. Rodriguez, T. A., Srinivas, S., Clements, M. P., Smith, J. C. & Beddington, R. S. Induction and migration of the anterior visceral endoderm is regulated by the extra-embryonic ectoderm. Development 132, 2513–2520 (2005).

    Article  CAS  PubMed  Google Scholar 

  272. Waldrip, W. R., Bikoff, E. K., Hoodless, P. A., Wrana, J. L. & Robertson, E. J. Smad2 signaling in extraembryonic tissues determines anterior–posterior polarity of the early mouse embryo. Cell 92, 797–808 (1998).

    Article  CAS  PubMed  Google Scholar 

  273. Lawson, K. A. et al. Bmp4 is required for the generation of primordial germ cells in the mouse embryo. Genes Dev. 13, 424–436 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  274. Martin, J. F. & Wagner, G. P. The origin of platelets enabled the evolution of eutherian placentation. Biol. Lett. 15, 20190374 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Roberts, R. M., Green, J. A. & Schulz, L. C. The evolution of the placenta. Reproduction 152, R179–189 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The Macfarlan laboratory is funded by the National Institutes of Health (NIH) (DIR 1ZIAHD008933) with additional funding through the Eunice Kennedy Shriver National Institute of Child Health and Human Development (NICHD) human placenta project. The authors thank L. Robertson, P. Kurbel, C. Simon, M. Almeida and J. Frank, and members of the Macfarlan laboratory as well as the NICHD Genetics and Epigenetics of Development Group, for discussions and critical comments on the manuscript. Special thanks to K. Pfeifer, P. Rocha, J. Thompson, A. Ivanoff, R. Cosby and M. Gauchier as well as the reviewers for their critical feedback. They apologize to those authors not cited due to space constraints.

Author information

Authors and Affiliations

Authors

Contributions

Both authors contributed equally to all aspects of the article.

Corresponding authors

Correspondence to Anna D. Senft or Todd S. Macfarlan.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Genetics thanks M. Branco, J. Rossant and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

Dfam TE Classification: https://www.dfam.org/classification/

Dfam TE database: https://www.dfam.org/home

Endoderm Explorer (GUI mouse scRNA-seq): https://endoderm-explorer.com/

gnomAD human sequencing data (including pLi): https://gnomad.broadinstitute.org/

Gypsy TE Database (GyDB): http://gydb.org/index.php/Main_Page

Human Gastrulation (GUI human scRNA-seq): http://human-gastrula.net/

Maternal–Fetal interface (GUI human scRNA-seq): https://maternal-fetal-interface.cellgeni.sanger.ac.uk/

Mouse Genome Informatics (MGI): http://www.informatics.jax.org/

Mouse gastrulation (GUI mouse scRNA-seq): https://marionilab.cruk.cam.ac.uk/MouseGastrulation2018/

Pfam protein domain database: https://pfam.xfam.org/

RepeatMasker Genomic Datasets: http://repeatmasker.org/genomicDatasets/RMGenomicDatasets.html

TEToolkit: https://github.com/mhammell-laboratory/TEtranscriptsTEhub (collaborative platform and reference site): https://tehub.org/

Supplementary information

Glossary

Transposable elements

(TEs). Mobile DNA segments present in virtually all domains of life whose remains make up nearly 50% of most mammalian genomes.

Transpose

Pertaining to transposition, which is the process by which transposable elements (TEs) change their genomic position, leading to the description of TEs as ‘jumping genes’.

Krüppel-associated box zinc finger proteins

(KRAB-ZFPs). KRAB domain containing C2H2-type zinc finger proteins that predominantly act as transcriptional repressors of transposable elements (TEs).

C2H2-type zinc finger

One of the most common domains found in the transcription factors of higher eukaryotes consisting of two cysteines and two histidines that coordinate the binding of a zinc ion, most frequently found in an array of repeated fingers encoded by a single exon.

Purifying selection

Preservation of a nucleotide or amino acid sequence by selective removal of mutations that reduce fitness (for example, visible by an excess of synonymous mutations, which conserve protein structure, relative to non-synonymous mutations).

Long terminal repeat

(LTR). A direct repeat that marks the boundaries of LTR retrotransposons that contain regulatory elements and polyadenylation signals that regulate and enable viral transcription. Initially identical LTRs can recombine, giving rise to so-called solo LTRs.

Endogenous retroviruses

(ERVs). Proviral remnants of ancient retroviral infections of the germ line. ERVs encode viral proteins such as envelopes (env), polymerases (pol) and structural proteins (gag) that, in some cases, allow them to exit as viral particles.

Germ line

Cells in an organism that pass on genetic material to the progeny. The germ line comprises egg and sperm cells (germ cells) and their precursors (primordial germ cells).

Amniotes

A clade of tetrapod vertebrates that comprises reptiles, birds and mammals that evolved ~320 million years ago and relies on internal fertilization and novel extra-embryonic tissues to develop on land.

Extra-embryonic tissues

Fetus-derived tissues that support embryonic development but (for the most part) do not give rise to embryonic structures.

Placenta

A temporary, shared organ of embryonic, extra-embryonic and (in eutheria) also maternal origin that allows development inside the mother by mediating nutrient and waste exchange.

Eutheria

Mammals that develop a placenta from three tissues: the maternal decidua as well as the fetal allantois and chorion.

Therian

Live-born mammals that include marsupials and eutheria.

Invasive placenta

A placenta is invasive when it makes direct contact with maternal blood. This is the case in both mouse and humans, as trophectoderm-derived trophoblast giant cells replace maternal endothelial cells to line maternal blood vessels.

Long interspersed elements

(LINEs). Autonomous non-LTR retrotransposons that comprise the largest transposable element (TE) class in human and mice (~22% of total genomic content).

Short interspersed elements

(SINEs). Non-autonomous non-LTR retrotransposons that rely on the long interspersed element (LINE) machinery to mobilize and are abundant in mammalian genomes (~7.5% mouse, ~13.5% human).

SVA

(SINE variable-number tandem-repeat Alu). Hominid-specific chimeric non-autonomous (long interspersed element 1 (L1)-dependent) retroelements with variable copy number (~2,700 copies in humans, with some of them active) that are a composite of short interspersed element (SINE)-R, variable-number tandem-repeat and Alu elements.

Implantation

The process by which eutherian embryos attach to the uterine wall to establish a sustained maternal–fetal interface that maintains pregnancy. Implantation induces an inflammatory response and can be followed by invasion.

Gastrulation

The embryonic process, lasting for ~72 h from days 6 to 8 in mouse development, that specifies somatic cell fates and a subset of extra-embryonic cell types as well as the germ line, after which twins can no longer form.

Co-option

A process by which natural selection repurposes transposable elements (TEs) for host functions.

Long interspersed element 1

(L1; also known as LINE-1). An active mammalian LINE family in mouse and human (representing 17–20% of their genomes).

DNA methylation

Covalent chemical modification in the form of methylation of cytosines at the fifth position (5-methyl-cytosine (5mC)) that in mammals occurs genome wide predominantly at CpG dinucleotides.

DUF3669

A domain of unknown function found in a few ancient Krüppel-associated box zinc finger proteins (KRAB-ZFPs) that may aid in KRAB-ZFP oligomerization.

SCAN

A gag capsid protein-derived domain found in selected ancient Krüppel-associated box zinc finger proteins (KRAB-ZFPs) that mediates protein interactions.

Zinc fingerprint amino acids

Amino acids at positions −1, +2, +3, and +6 of the C2H2 α-helix that are responsible for direct interactions with specific DNA bases, thus determining the specificity of the zinc fingers.

Zygotic genome activation

(ZGA). The process during which transcription from maternal and paternal genomes in the embryo takes over from maternal control. It occurs at the two-cell or four-cell stage in mice and humans, respectively.

MuERV-L

A murine-specific family of long terminal repeat (LTR) retrotransposons that are highly transcribed at zygotic genome activation (ZGA), present in ~1,500 copies. These elements are related to the endogenous retrovirus L (ERV-L) family of retrotransposons found throughout eutherian mammals that include human ERV-L (HERV-L) elements found in ~2,000 copies.

Long non-coding RNAs

(lncRNAs). RNA molecules >200 nucleotides long that do not encode proteins.

Trophectoderm

A mammalian-specific extra-embryonic cell lineage that gives rise to invasive trophoblast giant cells that mediate implantation in eutheria and to the chorion that contributes to all mammalian placentae.

Blastocyst

A thin-walled hollow embryonic structure containing a eutherian-specific inner cell mass, the future embryo and — dependent on its maturation — one or two extra-embryonic cell lineages.

Trophoblast stem cells

(TSCs). Undifferentiated extra-embryonic stem cells isolated from the trophectoderm in the blastocyst or TSCs shortly after implantation that can functionally contribute to placental development when reintroduced into blastocysts.

Enhancers

Non-coding short (typically 100–1,000 bp in length) DNA sequences that act to drive transcription independent of their relative distance, location or orientation to their cognate promoter. Putative enhancers bear histone modifications such as acetylated histone H3 lysine 27 (H3K27ac) and methylated H3 lysine 4 (H3K4me1).

Embryonic stem cells

(ESCs). Pluripotent stem cells derived from the inner cell mass of the blastocyst that can functionally contribute to the germ line and somatic tissues when reintroduced into blastocysts.

Trophoblast giant cells

Heterogeneous, large, polyploid trophectoderm-derived cells that mediate implantation and decidualization but also line maternal blood vessels and produce hormones in the placenta.

RLTR13

Mouse-specific endogenous retroviruses (ERVs) of the ERV-K family (K standing for the lysine tRNA primer used during reverse transcription) with many subfamilies, for example RLTR13D5 (~685 copies) or RLTR13D6 (~790 copies).

Amnion

Extra-embryonic tissue in amniotes that fills with fluid to form a sac that protects the embryo and provides a watery environment for developing on land.

Allantois

An extra-embryonic mesoderm-derived vascularized sac-like structure that allows gas exchange and disposes of liquid waste. It forms part of the eutherian placenta and becomes the umbilical cord.

Chorion

An extra-embryonic tissue that exchanges gases with the environment and originated in amniotes.

Yolk sac

The oldest extra-embryonic tissue in amniotes that originally absorbed nutrients deposited in yolk but has — despite the lack of yolk in eutherian mammals — remained essential for embryonic development, for example for embryonic patterning and blood development.

Extravillous trophoblast cells

Invasive, differentiated human trophoblast cells (equivalent to mouse trophoblast giant cells) that invade deeply into the uterine stroma, where they line and remodel maternal spiral arteries to supply the placenta with maternal blood.

Extra-embryonic endoderm

The precursor tissue of parts of the mammalian yolk sacs and substantial parts of the developing mammalian gut tube and associated endodermal organs. Extra-embryonic endoderm is also important for embryonic patterning and the specification of embryonic cell types, for example the future brain.

Pluripotency

The ability to functionally contribute to all somatic cells and the germ line of an organism that is maintained after implantation until specific cell types and the germ line are specified through gastrulation.

ERV-K

A subfamily of endogenous retroviruses (ERVs) that uses a lysine (K) tRNA that binds to the viral primer binding site to prime reverse transcription. ERV-K elements in mice contain many mobile elements (for example, intracisternal A-type particle (IAP), early transposon (ETn)).

Embryonic carcinoma cells

(EC cells). Malignant pluripotent cells derived from germ cell tumours called teratocarcinomas. Their culture preceded and paved the way for establishing embryonic stem cells (ESCs).

CRISPR activation/interference

Sequence-specific activation or repression of gene expression on the transcriptional level by exploiting the bacterial genetic immune system clustered regularly interspaced short palindromic repeats (CRISPR) pathway with nuclease-dead Cas9 (dCas9) fused to an activation or repression domain.

LTR5HS

Long terminal repeat (LTR) of one of the youngest families of human endogenous retrovirus-K (HERV-K) elements present in ~700 copies in the human genome.

Epigenetic reprogramming

Erasure and remodelling of epigenetic marks, such as DNA methylation and histone modifications, particularly during early development and in the germ line.

Intracisternal A-type particles

(IAPs). A rodent-specific and still actively retrotransposing class of transposable elements (TEs). They are among the most mutagenic long terminal repeat (LTR) retrotransposons in mammals and are present at ~1,000 full-length copies per haploid genome in mice.

Early transposon

(ETn). A type of non-autonomous mouse-specific long terminal repeat (LTR) retrotransposon that is mobilized by related MusD elements. They are called ‘early transposons’ because they are undetectable in differentiated cell lines. They still actively retrotranspose.

P element

A eukaryotic DNA transposon discovered as one of the first transposable elements (TEs) in Drosophila melanogaster.

Host–transposase fusions

(HTFs). Fusion proteins combining DNA transposase and host-derived domains that often originate via alternative splicing.

Somatic cell nuclear transfer

Transfer of a nucleus of an adult somatic cell to an enucleated egg cell. Pioneered in frogs, nuclear transfer can be used to clone animals as it leads to epigenetic reprogramming towards pluripotency.

Genetic drift

Random sampling from parental genomes in the next generation that generates changes in the frequency of existing genetic traits in a population.

Maternal–fetal interface

Sites where the genetically distinct mother and fetus come into indirect contact. It can refer both to the contact of mother and fetus in the functional placenta and to the implantation site.

Chorioallantoic fusion

Fusion between the allantois and the chorion at day 8–9 of mouse development that subsequently allows the vasculature of the allantois to grow into chorion-derived villi to form the labyrinth layer of the placenta.

Syncytiotrophoblast

Multinucleated layers (two in mice, one in humans) in the placenta that are in indirect contact with maternal blood and act as a transport surface.

env

Retroviral gene encoding a cell surface protein that mediates viral–host cell fusion.

gag

Retroviral genes encoding structural proteins of retroviruses that contribute to capsid formation and are essential for infectivity.

pol

Retroviral genes encoding proteins with enzymatic activities required for viral replication (for example, reverse transcriptase), integration and protein cleavage.

sushi-ichi

Long terminal repeat (LTR) retrotransposons of the Ty3/Gypsy class that encode chromodomain integrases, a primer binding site, two open reading frames for gag and pol genes and a polypurine tract.

Labyrinth

The innermost layer of the mouse placenta that is the main site of nutrient and gas exchange and consists of the syncytiotrophoblast and the vascular network.

Spongiotrophoblast

The hormone-secreting, trophectoderm-derived middle layer of the mouse placenta sandwiched between the outer secondary trophoblast giant cells and the inner labyrinth layer with poorly understood, yet essential, functions in pregnancy.

Genomic imprinting

The parent of origin-specific expression of genes that is established during germ-line development and depends on differential DNA methylation of so-called imprinting control regions (ICRs). Canonical imprints solely depend on DNA methylation, whereas non-canonical imprints (prevalent in the placenta) initially depend on maternally inherited repressive histone marks that are replaced by DNA methylation after implantation.

Gynogenetic

Pertaining to gynogenesis, which is development whereby the embryo contains only maternal chromosomes. Biparental, that is diploid, gynogenetic embryos are produced by transplanting pronuclei between one-cell stage embryos to yield two female pronuclei.

Androgenetic

Pertaining to androgenesis, which is development whereby the embryo contains only paternal chromosomes (see gynogenetic).

Parthenogenetically

Pertaining to parthenogenesis, which is reproduction without fertilization (‘virgin birth’). Parthenogenesis is common in invertebrates and plants.

Mammalian apparent LTR retrotransposons

(MaLRs). Non-autonomous mouse-specific endogenous retroviruses (ERVs) that are the most common retroviral elements in the mouse, making up 4.8% of the genome, and are related to mouse and human ERV-L and human long terminal repeat (LTR) transposon-like human element 1 (THE1; an LTR-containing retrotransposon) elements.

AmnSINE1

A short interspersed nuclear element (SINE) family that amplified in the common amniote ancestor ~320 million years ago that has been conserved in ~200–700 copies in mouse and human genomes.

X-chromosome inactivation

The mammalian-specific process of almost complete silencing of one of the two X chromosomes to equalize the sex chromosome dose in female theria.

Decidualization

Changes of the maternal endometrium in preparation for, and during, mammalian pregnancy. It occurs during menstruation in humans, but is induced by pregnancy in mice (where it can also be triggered by uterine injection with mineral oil drops).

Peripheral regulatory T cells

(pTreg cells). A FOXP3-expressing, specialized immunosuppressive T cell lineage generated extrathymically (peripheral). Regulatory T cells are also generated in the thymus (tTreg cells). Regulatory T cells have essential functions in preventing fatal autoimmunity and inflammation and additional roles in metabolism and tissue repair.

Mammalian-wide interspersed repeats

(MIRs). A widespread subfamily of short interspersed nuclear elements (SINEs) that is often conserved in mammals.

Allogeneic

Genetically dissimilar, and hence immunologically incompatible (within a species); for example, major histocompatibility complex (MHC)-mismatched mouse strains. The opposite of syngeneic, which means genetically similar or identical.

Medium reiteration frequency interspersed repeats

(MERs). A loose grouping of diverse types of DNA transposons and retrotransposons. For example, whereas MER20 is a non-autonomous eutherian-specific subfamily of hobo Activator Tam3 (hAT) Charlie DNA transposons (~4,800 mouse and ~16,000 human copies), MER77 (~400 copies in mouse and 1,200 copies in human) and MER39 (~2600 human copies) are endogenous retroviruses (ERVs).

Parturition

The act or process of giving birth.

LTR transposon-like human element 1B

(THE1B). A long terminal repeat (LTR)-containing retrotransposon that colonized anthropoid primates ~50 million years ago and is present in ~20,000 copies in the human genome.

Epistatic

Pertaining to epistasis, which is the genetic phenomenon whereby the effect of a mutation depends on the presence or absence of other mutations.

MT2B2

A mouse-specific long terminal repeat (LTR) of a class 3 endogenous retrovirus (ERV) present in ~15,000 copies.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Senft, A.D., Macfarlan, T.S. Transposable elements shape the evolution of mammalian development. Nat Rev Genet 22, 691–711 (2021). https://doi.org/10.1038/s41576-021-00385-1

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41576-021-00385-1

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing