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.

  • Article
  • Published:

poly(UG)-tailed RNAs in genome protection and epigenetic inheritance

An Author Correction to this article was published on 14 April 2021

This article has been updated

Abstract

Mobile genetic elements threaten genome integrity in all organisms. RDE-3 (also known as MUT-2) is a ribonucleotidyltransferase that is required for transposon silencing and RNA interference in Caenorhabditis elegans1,2,3,4. When tethered to RNAs in heterologous expression systems, RDE-3 can add long stretches of alternating non-templated uridine (U) and guanosine (G) ribonucleotides to the 3′ termini of these RNAs (designated poly(UG) or pUG tails)5. Here we show that, in its natural context in C. elegans, RDE-3 adds pUG tails to targets of RNA interference, as well as to transposon RNAs. RNA fragments attached to pUG tails with more than 16 perfectly alternating 3′ U and G nucleotides become gene-silencing agents. pUG tails promote gene silencing by recruiting RNA-dependent RNA polymerases, which use pUG-tailed RNAs (pUG RNAs) as templates to synthesize small interfering RNAs (siRNAs). Our results show that cycles of pUG RNA-templated siRNA synthesis and siRNA-directed pUG RNA biogenesis underlie double-stranded-RNA-directed transgenerational epigenetic inheritance in the C. elegans germline. We speculate that this pUG RNA–siRNA silencing loop enables parents to inoculate progeny against the expression of unwanted or parasitic genetic elements.

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: pUG tails are added to mRNA fragments in vivo.
Fig. 2: pUG tails convert inert RNA fragments into agents of gene silencing.
Fig. 3: Endogenous pUG RNAs exist and localize to germline Mutator foci.
Fig. 4: pUG RNAs are templates for RdRPs.
Fig. 5: pUG RNA–siRNA cycles drive heritable gene silencing.

Similar content being viewed by others

Data availability

All DNA and RNA sequencing data discussed in this publication have been deposited in NCBI’s Gene Expression Omnibus61 (GEO) under accession number GSE148134. Processed MiSeq and small-RNA sequencing reads are also provided in Supplementary Tables 1 and 4, respectively.

Code availability

Descriptions of custom scripts used to analyse MiSeq and RNA-seq data are provided in the Methods and the scripts are available upon request from the corresponding author. Custom Python scripts used to analyse small-RNA sequencing data have been deposited at https://github.com/Yuhan-Fei/pUG-analysis.

Change history

References

  1. Collins, J., Saari, B. & Anderson, P. Activation of a transposable element in the germ line but not the soma of Caenorhabditis elegans. Nature 328, 726–728 (1987).

    ADS  CAS  PubMed  Google Scholar 

  2. Ketting, R. F., Haverkamp, T. H. A., van Luenen, H. G. A. M. & Plasterk, R. H. A. mut-7 of C. elegans, required for transposon silencing and RNA interference, is a homolog of Werner syndrome helicase and RNaseD. Cell 99, 133–141 (1999).

    CAS  PubMed  Google Scholar 

  3. Tabara, H. et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell 99, 123–132 (1999).

    CAS  PubMed  Google Scholar 

  4. Chen, C.-C. G. et al. A member of the polymerase β nucleotidyltransferase superfamily is required for RNA interference in C. elegans. Curr. Biol. 15, 378–383 (2005).

    CAS  PubMed  Google Scholar 

  5. Preston, M. A. et al. Unbiased screen of RNA tailing activities reveals a poly(UG) polymerase. Nat. Methods 16, 437–445 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Fire, A. et al. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature 391, 806–811 (1998).

    ADS  CAS  PubMed  Google Scholar 

  7. Aravind, L. & Koonin, E. V. DNA polymerase β-like nucleotidyltransferase superfamily: identification of three new families, classification and evolutionary history. Nucleic Acids Res. 27, 1609–1618 (1999).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Martin, G. & Keller, W. RNA-specific ribonucleotidyl transferases. RNA 13, 1834–1849 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  9. Detwiler, M. R., Reuben, M., Li, X., Rogers, E. & Lin, R. Two zinc finger proteins, OMA-1 and OMA-2, are redundantly required for oocyte maturation in C. elegans. Dev. Cell 1, 187–199 (2001).

    CAS  PubMed  Google Scholar 

  10. Parrish, S. & Fire, A. Distinct roles for RDE-1 and RDE-4 during RNA interference in Caenorhabditis elegans. RNA 7, 1397–1402 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Tabara, H., Yigit, E., Siomi, H. & Mello, C. C. The dsRNA binding protein RDE-4 interacts with RDE-1, DCR-1, and a DExH-box helicase to direct RNAi in C. elegans. Cell 109, 861–871 (2002).

    CAS  PubMed  Google Scholar 

  12. Tsai, H.-Y. et al. A ribonuclease coordinates siRNA amplification and mRNA cleavage during RNAi. Cell 160, 407–419 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Ko, F. C. F. & Chow, K. L. A novel thioredoxin-like protein encoded by the C. elegans dpy-11 gene is required for body and sensory organ morphogenesis. Development 129, 1185–1194 (2002).

    CAS  PubMed  Google Scholar 

  14. Lin, R. A gain-of-function mutation in oma-1, a C. elegans gene required for oocyte maturation, results in delayed degradation of maternal proteins and embryonic lethality. Dev. Biol. 258, 226–239 (2003).

    CAS  PubMed  Google Scholar 

  15. Fischer, S. E. J., Wienholds, E. & Plasterk, R. H. A. Continuous exchange of sequence information between dispersed Tc1 transposons in the Caenorhabditis elegans genome. Genetics 164, 127–134 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Voronina, E., Seydoux, G., Sassone-Corsi, P. & Nagamori, I. RNA granules in germ cells. Cold Spring Harb. Perspect. Biol. 3, a002774 (2011).

    PubMed  PubMed Central  Google Scholar 

  17. Phillips, C. M., Montgomery, T. A., Breen, P. C. & Ruvkun, G. MUT-16 promotes formation of perinuclear Mutator foci required for RNA silencing in the C. elegans germline. Genes Dev. 26, 1433–1444 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Austin, J. & Kimble, J. glp-1 is required in the germ line for regulation of the decision between mitosis and meiosis in C. elegans. Cell 51, 589–599 (1987).

    CAS  PubMed  Google Scholar 

  19. Jose, A. M., Garcia, G. A. & Hunter, C. P. Two classes of silencing RNAs move between Caenorhabditis elegans tissues. Nat. Struct. Mol. Biol. 18, 1184–1188 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Buratti, E. & Baralle, F. E. Characterization and functional implications of the RNA binding properties of nuclear factor TDP-43, a novel splicing regulator of CFTR exon 9. J. Biol. Chem. 276, 36337–36343 (2001).

    CAS  PubMed  Google Scholar 

  21. Kuo, P.-H., Doudeva, L. G., Wang, Y.-T., Shen, C.-K. J. & Yuan, H. S. Structural insights into TDP-43 in nucleic-acid binding and domain interactions. Nucleic Acids Res. 37, 1799–1808 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Sijen, T. et al. On the role of RNA amplification in dsRNA-triggered gene silencing. Cell 107, 465–476 (2001).

    CAS  PubMed  Google Scholar 

  23. Ambros, V., Lee, R. C., Lavanway, A., Williams, P. T. & Jewell, D. MicroRNAs and other tiny endogenous RNAs in C. elegans. Curr. Biol. 13, 807–818 (2003).

    CAS  PubMed  Google Scholar 

  24. Sijen, T., Steiner, F. A., Thijssen, K. L. & Plasterk, R. H. A. Secondary siRNAs result from unprimed RNA synthesis and form a distinct class. Science 315, 244–247 (2007).

    ADS  CAS  PubMed  Google Scholar 

  25. Pak, J. & Fire, A. Distinct populations of primary and secondary effectors during RNAi in C. elegans. Science 315, 241–244 (2007).

    ADS  CAS  PubMed  Google Scholar 

  26. Gu, W. et al. Distinct Argonaute-mediated 22G-RNA pathways direct genome surveillance in the C. elegans germline. Mol. Cell 36, 231–244 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Billi, A. C., Fischer, S. E. J. & Kim, J. K. Endogenous RNAi pathways in C. elegans. WormBook https://doi.org/10.1895/wormbook.1.170.1 (2014). 

  28. Vastenhouw, N. L. et al. Gene expression: long-term gene silencing by RNAi. Nature 442, 882 (2006).

    ADS  CAS  PubMed  Google Scholar 

  29. Buckley, B. A. et al. A nuclear Argonaute promotes multigenerational epigenetic inheritance and germline immortality. Nature 489, 447–451 (2012).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  30. Ashe, A. et al. piRNAs can trigger a multigenerational epigenetic memory in the germline of C. elegans. Cell 150, 88–99 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Shirayama, M. et al. piRNAs initiate an epigenetic memory of nonself RNA in the C. elegans germline. Cell 150, 65–77 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Luteijn, M. J. et al. Extremely stable Piwi-induced gene silencing in Caenorhabditis elegans. EMBO J. 31, 3422–3430 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Sapetschnig, A., Sarkies, P., Lehrbach, N. J. & Miska, E. A. Tertiary siRNAs mediate paramutation in C. elegans. PLoS Genet. 11, e1005078 (2015).

    PubMed  PubMed Central  Google Scholar 

  34. Yigit, E. et al. Analysis of the C. elegans Argonaute family reveals that distinct Argonautes act sequentially during RNAi. Cell 127, 747–757 (2006).

    CAS  PubMed  Google Scholar 

  35. Talsky, K. B. & Collins, K. Initiation by a eukaryotic RNA-dependent RNA polymerase requires looping of the template end and is influenced by the template-tailing activity of an associated uridyltransferase. J. Biol. Chem. 285, 27614–27623 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Czech, B. et al. piRNA-guided genome defense: from biogenesis to silencing. Annu. Rev. Genet. 52, 131–157 (2018).

    CAS  PubMed  Google Scholar 

  37. Brenner, S. The genetics of Caenorhabditis elegans. Genetics 77, 71–94 (1974).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Smith, T., Heger, A. & Sudbery, I. UMI-tools: modeling sequencing errors in Unique Molecular Identifiers to improve quantification accuracy. Genome Res. 27, 491–499 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Martin, M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.journal 17, 10–12 (2011).

    Google Scholar 

  40. Dobin, A. et al. STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15–21 (2013).

    CAS  PubMed  Google Scholar 

  41. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  42. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Phanstiel, D. H., Boyle, A. P., Araya, C. L. & Snyder, M. P. Sushi.R: flexible, quantitative and integrative genomic visualizations for publication-quality multi-panel figures. Bioinformatics 30, 2808–2810 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Wagih, O. ggseqlogo: a versatile R package for drawing sequence logos. Bioinformatics 33, 3645–3647 (2017).

    CAS  PubMed  Google Scholar 

  45. Alcazar, R. M., Lin, R. & Fire, A. Z. Transmission dynamics of heritable silencing induced by double-stranded RNA in Caenorhabditis elegans. Genetics 180, 1275–1288 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    CAS  PubMed  Google Scholar 

  47. Jin, Y., Tam, O. H., Paniagua, E. & Hammell, M. TEtranscripts: a package for including transposable elements in differential expression analysis of RNA-seq datasets. Bioinformatics 31, 3593–3599 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. Cunningham, F. et al. Ensembl 2019. Nucleic Acids Res. 47, D745–D751 (2019).

    CAS  PubMed  Google Scholar 

  49. Claycomb, J. M. et al. The Argonaute CSR-1 and its 22G-RNA cofactors are required for holocentric chromosome segregation. Cell 139, 123–134 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Wu, W.-S. et al. piRTarBase: a database of piRNA targeting sites and their roles in gene regulation. Nucleic Acids Res. 47, D181–D187 (2019).

    CAS  PubMed  Google Scholar 

  51. Wan, G. et al. Spatiotemporal regulation of liquid-like condensates in epigenetic inheritance. Nature 557, 679–683 (2018).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  52. Schwartz, M. L. & Jorgensen, E. M. SapTrap, a toolkit for high-throughput CRISPR/Cas9 gene modification in Caenorhabditis elegans. Genetics 202, 1277–1288 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Dickinson, D. J., Slabodnick, M. M., Chen, A. H. & Goldstein, B. SapTrap assembly of repair templates for Cas9-triggered homologous recombination with a self-excising cassette. MicroPublication Biol. https://doi.org/10.17912/W2KT0N (2018).

  54. Dickinson, D. J., Pani, A. M., Heppert, J. K., Higgins, C. D. & Goldstein, B. Streamlined genome engineering with a self-excising drug selection cassette. Genetics 200, 1035–1049 (2015).

    PubMed  PubMed Central  Google Scholar 

  55. Frøkjaer-Jensen, C. et al. Single-copy insertion of transgenes in Caenorhabditis elegans. Nat. Genet. 40, 1375–1383 (2008).

    PubMed  PubMed Central  Google Scholar 

  56. Haeussler, M. et al. Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection tool CRISPOR. Genome Biol. 17, 148 (2016).

    PubMed  PubMed Central  Google Scholar 

  57. Dodson, A. E. & Kennedy, S. Germ granules coordinate RNA-based epigenetic inheritance pathways. Dev. Cell 50, 704–715 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Trapnell, C., Pachter, L. & Salzberg, S. L. TopHat: discovering splice junctions with RNA-seq. Bioinformatics 25, 1105–1111 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Ramírez, F., Dündar, F., Diehl, S., Grüning, B. A. & Manke, T. deepTools: a flexible platform for exploring deep-sequencing data. Nucleic Acids Res. 42, W187–W191 (2014).

    PubMed  PubMed Central  Google Scholar 

  60. Liao, Y., Smyth, G. K. & Shi, W. featureCounts: an efficient general purpose program for assigning sequence reads to genomic features. Bioinformatics 30, 923–930 (2014).

    CAS  PubMed  Google Scholar 

  61. Edgar, R., Domrachev, M. & Lash, A. E. Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res. 30, 207–210 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Xiong, Y. & Steitz, T. A. Mechanism of transfer RNA maturation by CCA-adding enzyme without using an oligonucleotide template. Nature 430, 640–645 (2004).

    ADS  CAS  PubMed  Google Scholar 

  63. Guang, S. et al. An Argonaute transports siRNAs from the cytoplasm to the nucleus. Science 321, 537–541 (2008).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  64. Guang, S. et al. Small regulatory RNAs inhibit RNA polymerase II during the elongation phase of transcription. Nature 465, 1097–1101 (2010).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  65. Zhang, C. et al. mut-16 and other mutator class genes modulate 22G and 26G siRNA pathways in Caenorhabditis elegans. Proc. Natl Acad. Sci. USA 108, 1201–1208 (2011).

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  66. Remy, J.-J. Stable inheritance of an acquired behavior in Caenorhabditis elegans. Curr. Biol. 20, R877–R878 (2010).

    CAS  PubMed  Google Scholar 

  67. Rechavi, O. et al. Starvation-induced transgenerational inheritance of small RNAs in C. elegans. Cell 158, 277–287 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Schott, D., Yanai, I. & Hunter, C. P. Natural RNA interference directs a heritable response to the environment. Sci. Rep. 4, 7387 (2015).

    Google Scholar 

  69. Jobson, M. A. et al. Transgenerational effects of early life starvation on growth, reproduction, and stress resistance in Caenorhabditis elegans. Genetics 201, 201–212 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Moore, R. S., Kaletsky, R. & Murphy, C. T. Piwi/PRG-1 Argonaute and TGF-β mediate transgenerational learned pathogenic avoidance. Cell 177, 1827–1841 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Posner, R. et al. Neuronal small RNAs control behavior transgenerationally. Cell 177, 1814–1826 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank past and present members of the Kennedy, S. E. Butcher and Wickens laboratories for helpful discussions; the Biopolymers Facility at Harvard Medical School (HMS) for Illumina sequencing; the Dana-Farber/Harvard Cancer Center DNA Resource Core for Sanger sequencing; and the Taplin Mass Spectrometry Facility in the Cell Biology Department at HMS for performing LC–MS/MS. Some strains were provided by the Caenorhabditis Genetics Center (CGC), which is funded by the NIH Office of Research Infrastructure Programs (P40 OD010440). Some strains were provided by the S. Mitani laboratory through the National BioResource Project (Tokyo, Japan), which is part of the International C. elegans Gene Knockout Consortium. A.S. and J.Y. were supported by the Ruth L. Kirschstein T32 Predoctoral NRSA (T32GM096911) and NSF Graduate Research Fellowships (A.S., DGE1144152, DGE1745303; J.Y., DGE1745303). A.E.D. is a Damon Runyon Fellow supported by the Damon Runyon Cancer Research Foundation (DRG-2304-17). D.J.P. was supported by a Ruth L. Kirschstein National Research Service Award (1F32GM125345-01). S.K. was supported by NIH grants GM088289 and GM132286., and M.W. by NIH grant GM50942.

Author information

Authors and Affiliations

Authors

Contributions

A.S. contributed to Figs. 1a–d, 3a–e, 5a, 5c–e, Extended Data Figs. 1a, b, 2a, b, 3a, 5a–d, 6a–d, 8a, c, d, f, 9, 10a–c, Supplementary Tables 1, 2. J.Y. contributed to Figs. 2a–c, 3f, 4a–d, 5b, Extended Data Figs. 2c, d, 3b, 4a, b, 7a–c, 8b, e, Supplementary Tables 3, 4. D.J.P., J.G. and J.G.S. contributed to Supplementary Table 2. A.E.D. contributed to Fig. 4a, Extended Data Figs. 1a, b, 5a, 10b, Supplementary Tables 1, 3. Y.F. contributed to Fig. 4d, Extended Data Figs. 7a–c, 8e, Supplementary Table 4. A.S., M.W. and S.K. conceived the project. S.K. supervised the project. A.S. and S.K. wrote the manuscript.

Corresponding author

Correspondence to Scott Kennedy.

Ethics declarations

Competing interests

M.W. has a patent (US20160145666A1) through Wisconsin Alumni Research Foundation (Madison, WI) for methods, kits and compositions of matter relating to poly(UG) polymerases.

Additional information

Peer review information Nature thanks Rene Ketting, Taiowa Montgomery 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.

Extended data figures and tables

Extended Data Fig. 1 Analysis of oma-1 pUGylation sites.

a, Illumina MiSeq was performed (n = 1 biological experiment) on oma-1 pUG PCR products derived from WT and rde-3(-) worms, with or without oma-1 dsRNA. The number of sequenced pUG RNAs (y-axis) mapping to each pUGylation site (x-axis) is shown. Inset, total number of sequenced oma-1 pUG RNAs from indicated samples and total number of these sequenced pUG RNAs in which the oma-1 sequence was spliced. b, MiSeq-sequenced oma-1 pUG RNAs were sorted into four groups on the basis of the nucleotide at the last templated position (−1) of the oma-1 mRNA. The percentage of oma-1 pUG RNAs (MiSeq reads) with each nucleotide in the −1 position is shown. Logo analysis was then performed on each of the four groups to determine the probability of finding each nucleotide at the first position of the pUG tail (+1), as well as at the second-to-last templated nucleotide of oma-1 (−2). This analysis showed that if the last templated nucleotide of the oma-1 mRNA fragment was an A or a C, then RDE-3 was equally likely to add a U or a G as the first nucleotide of an elongating pUG tail. If, however, the last templated nucleotide was a U or G, then RDE-3 preferentially added a G or U, respectively, as the first nucleotide in an elongating pUG tail. Note, to perform the analyses in this figure, we assumed that if a U or G could have been genomically encoded, then it was. If, instead, RDE-3 added the U or G shown in the −1 position as the first nucleotide of the pUG tail, then these data show that the second nucleotide that RDE-3 prefers to add is a G after a U or a U after a G. CCA-adding rNT enzymes modify the 3′ termini of transfer RNAs (tRNAs) with non-templated CCA nucleotides. The mechanism by which these enzymes add non-templated nonhomopolymeric stretches of nucleotides is thought to involve allosteric regulation of the nucleotide-binding pocket by the 3′ nucleotide of a substrate tRNA62. A similar mechanism may explain how RDE-3 can add pUG tails to its mRNA substrates. For instance, when the 3′ nucleotide of an RDE-3 substrate is a U, the rNTP binding pocket of RDE-3 might adopt a structure that preferentially binds G and vice versa when the 3′ nucleotide of an RDE-3 substrate is a G. Such a model could explain how a single rNT enzyme adds perfectly alternating U and G nucleotides to RNA substrates. There are also alternative models for how RDE-3 might add pUG tails to an RNA. These include: (1) the existence of a poly(AC) nucleic acid template used by RDE-3 during pUG tail synthesis, (2) the existence of one or more rNTs that cooperate with RDE-3 to produce pUG tails, or (3) the possibility that RDE-3 binds and incorporates UG or GU dinucleotides. We disfavour the first two possibilities, as these models are difficult to reconcile with the observation that RDE-3 adds UG repeats to tethered RNAs in yeast or in Xenopus oocytes5. The third proposed model may be true, but because our sequencing shows that pUG tails can initiate with either a U or a G (this figure, Supplementary Table 1), then RDE-3 would need to be able to bind both UG and GU dinucleotides. Determining the mechanism by which RDE-3 adds pUG tails will probably involve structural studies and/or in vitro pUGylation assays using recombinant RDE-3 protein.

Extended Data Fig. 2 RNAi-triggered pUGylation and pUG RNA-directed gene silencing are general and sequence-specific.

a, gfp::h2b, rde-3(-); gfp::h2b and WT (no gfp::h2b) worms were fed E. coli expressing either empty vector control or gfp dsRNA. b, WT and rde-3(-) worms were fed E. coli expressing empty vector control and either oma-1 or dpy-11 dsRNA. In a, b, gfp, dpy-11 and oma-1 pUG RNAs were detected using the assay outlined in Fig. 1a. Data are representative of three biologically independent experiments. c, rde-1(ne219); oma-1(zu405ts) worms were injected with either an oma-1 (n = 6) or a gfp (n = 10) pUG RNA. n = 3 for no injection. The percentage of embryos hatched was scored for the progeny of injected worms. Panels below the x-axis show RNAs run on 2% agarose gel to assess RNA integrity. Data are mean ± s.d. d, rde-1(ne219); gfp::h2b worms were injected with either an oma-1 or a gfp pUG RNA (n = 10 for both, 3 for no injection). Data are mean ± s.d. of percentage of progeny with gfp::h2b silenced. In c, d, all pUG tails were 36 nt in length.

Extended Data Fig. 3 RDE-3-mediated pUGylation is necessary for RNAi.

a, Worms of the indicated genotypes (all harbouring the oma-1(zu405ts) mutation) were treated with or without oma-1 dsRNA. For each experiment, the percentage of embryos hatched was scored at 20 °C and averaged for six individual worms per treatment for each genotype. rde-1(ne219) mutants, which cannot respond to dsRNA3, serve as a control for this experiment. Data are mean ± s.d of three biologically independent experiments. b, Control or rde-3(ne298) worms (all rde-1(ne219); oma-1(zu405ts) background) were injected with oma-1 pUG RNAs and the percentage of embryos hatched was scored at 20 °C. n = 10 noninjected and 16 injected worms for control. n = 8 noninjected and 14 injected worms for rde-3(ne298). Data are mean ± s.d.

Extended Data Fig. 4 pUG tails must be appended to sense RNAs of >50 nt for functionality.

a, b, rde-1(ne219); oma-1(zu405ts) worms were injected with: an oma-1 pUG RNA consisting of the sense or antisense strand of the same 541-nt-long oma-1 mRNA fragment (beginning at the aug) with a 36-nt 3′ pUG tail (a; n = 9 for both; n = 3 for no injection); oma-1 pUG RNAs consisting of oma-1 mRNA fragments of varying lengths (with position 1 starting at the aug of the oma-1 mRNA sequence) all appended to a 36-nt pUG tail (b; n = 6 (no injection), 10 (1–50), 17 (1–100), 8 (1–270), 9 (271–540) and 15 (1–540)). In a, b, percentage of embryonic arrest was scored at 20 °C. Data are mean ± s.d.

Extended Data Fig. 5 Endogenous targets of pUGylation in C. elegans.

a, mRNAs upregulated in rde-3(-) mutants (Supplementary Table 2) were compared to published lists of: (1) RNAs targeted by CSR-1-bound endo-siRNAs49, (2) piRNA-targeted mRNAs (based on predictive and experimental approaches)50, and (3) WAGO-class mRNAs26. P values were generated using a one-sided Fisher’s exact test. This analysis showed statistically significant overlap between the mRNAs upregulated in rde-3(-) mutants and both piRNA targets and WAGO-class mRNAs. bd, Total RNA was extracted from WT or rde-3(-) worms. The assay outlined in Fig. 1a was used to detect pUG RNAs for two DNA transposons (Tc4v and Tc5) and a retrotransposon (Cer3) that were significantly upregulated in rde-3(-) worms (b); predicted protein-coding mRNAs that were significantly upregulated in rde-3(-) worms (c); and two randomly selected mRNAs whose expression does not change in rde-3(-) mutants (d). Results are representative of three biologically independent experiments. The same reverse-transcribed samples were used for c and d and, therefore, the gsa-1 loading control is the same for both panels.

Extended Data Fig. 6 Mutator foci probably coordinate pUG RNA biogenesis in germ cells.

a, dpy-11 and oma-1 pUG PCR (Fig. 1a) were performed on total RNA from glp-1(q224/ts) worms grown at 15 °C (germ cells present) or 25 °C (approximately 99% of germ cells absent), with or without oma-1 and dpy-11 dsRNA. Data are representative of two biologically independent experiments. The samples in a are the same as those used in Fig. 3e and, therefore, the gsa-1 loading control is the same. b, oma-1 pUG PCR was performed on total RNA extracted from wild-type, rde-3(-) and mut-16(pk710) worms, with or without oma-1 dsRNA. Data are representative of four biologically independent experiments. c, RT–qPCR was used to quantify levels of oma-1 pUG RNAs in wild-type, rde-3(-) and mut-16(pk710) worms, with or without oma-1 dsRNA. Data are represented as fold change in the levels of oma-1 pUG RNAs with or without oma-1 dsRNA (y-axis) for each strain (x-axis). n = 3 biologically independent samples per treatment for each strain. Data are mean ± s.d. d, RT–qPCR was used to quantify levels of Tc1 pUG RNAs in wild-type, rde-3(-) and mut-16(pk710) worms. The RNA samples used for d are the same as those used in c, except that the data for the samples with and without oma-1 dsRNA were pooled for each strain. n = 6 biologically independent samples for each strain. Data are mean ± s.d. The analyses in c and d showed that mut-16 mutants produced more oma-1, but fewer Tc1, pUG RNAs than wild-type worms. The increased levels of oma-1 pUG RNAs in mut-16(pk710) worms was also suggested by the results in b. Together, these data suggest that Mutator foci probably have an important role in coordinating pUG RNA biogenesis in germ cells, as pUG RNA levels become misregulated in mut-16(pk710) mutants.

Extended Data Fig. 7 pUG RNAs are templates for RdRPs.

a, A biological replicate of the experiment shown in Fig. 4d was performed. oma-1(SNP) pUG or pGC RNAs were injected into rde-1(ne219); oma-1(zu405ts) germlines. SNP location is indicated with the dotted line. Injected worms were collected 1–4 h after injection, total RNA was isolated and small RNAs (20–30 nt) were sequenced. The distribution of 22G siRNAs mapping antisense to oma-1 is shown, with 22G siRNA reads normalized to reads per million total reads. oma-1 pUG (but not pGC) RNA injection triggered 22G siRNA production near the site of the pUG tail (pUG-specific 22G siRNAs). For unknown reasons, both pUG and pGC RNA injections triggered production of small RNAs around 400 bp 5′ of either tail. b, The length distribution of small-RNA reads mapping antisense to oma-1 is shown for small RNAs sequenced after oma-1(SNP) pUG RNA injections (Fig. 4d and a). c, The proportion of 22-nt small RNAs mapping antisense to oma-1 containing 5′ A, U, G or C is shown.

Extended Data Fig. 8 De novo pUGylation events in progeny are required for TEI.

a, oma-1(zu405ts) worms were fed bacteria expressing empty vector control or oma-1 dsRNA and the percentage of embryos hatched at 20 °C was scored for six generations. Data are mean ± s.d. of three biologically independent experiments. For each experiment, the percentage of embryos hatched at 20 °C was averaged for six individual worms per treatment for each genotype. b, rde-1(ne219); oma-1(zu405ts) worms were injected with co-injection marker alone (n = 12) or co-injection marker + oma-1 pUG RNA (n = 19) and the percentage of embryos hatched at 20 °C was scored for four generations in lineages of worms established from injected parents (see Methods for details of experimental set-up). Data are mean ± s.d. P values were generated using two-tailed unpaired Student’s t-test. c, c38d9.2 and Tc1 pUG RNA expression quantified by RT–qPCR in embryos collected from wild-type, rde-3(-) or MAGO12 worms. Fold change is normalized to rde-3(-). Each point (n) represents a biologically independent replicate, n = 3 independent replicates per strain. Data are mean ± s.d. d, Same experiment as Fig. 5d. rde-1(ne219); oma-1(zu405ts) worms were injected with an oma-1(SNP) pUG RNA or with co-injection marker only. Co-injection marker-expressing F1 progeny were picked and allowed to lay their F2 broods. oma-1 pUG PCR was performed on total RNA from F2 progeny. Shown are data from three biological replicates. e, Two biological replicates of small RNAs sequenced from the progeny of rde-1(ne219); oma-1(zu405ts) worms injected with oma-1(SNP) pUG or pGC RNAs are shown. Dotted line indicates the location of the SNP incorporated into oma-1. The distribution of 22G siRNAs mapping antisense to oma-1 is shown, with 22G siRNA reads normalized to reads per million total reads. In Fig. 4d and Extended Data Fig. 7a, small RNAs were sequenced 1–4 h after injection and 100% of 22G siRNAs antisense to the region of the engineered SNP in oma-1 were found to encode the complement of the SNP. Shown here, less than 1% of 22G siRNAs from the progeny of injected worms encoded the SNP complement. siRNAs mapping near the pUG tail were observed only after oma-1(SNP) pUG RNA injection (pUG-specific siRNAs). For unknown reasons, both oma-1(SNP) pUG and pGC RNAs triggered production of small RNAs 5′ of the pUG-specific siRNAs. It is possible that these siRNAs were triggered by systems that respond to foreign RNAs, such as the piRNA system. Further work will be needed to determine the aetiology of these siRNAs. f, Same experiment as Fig. 5e. oma-1(zu405ts) hermaphrodites were fed oma-1 dsRNA and crossed to rde-3(ne298); oma-1(zu405ts) males. F2 progeny from this cross were genotyped for rde-3(ne298). WT and rde-3(ne298) homozygous F3 progeny were phenotyped for the percentage of embryos hatched at 20 °C. Three biologically independent crosses (P0 1–3) were performed. Data are mean ± s.d. P values were generated using two-tailed unpaired Student’s t-tests.

Extended Data Fig. 9 Working model for pUG RNA–siRNA cycling during RNAi.

Initiation: exogenous and constitutive (that is, genomically encoded such as dsRNA or piRNAs) triggers direct RDE-3 to pUGylate RNAs previously fragmented by factors in the RNAi pathway. Maintenance: pUG RNAs are templates for secondary siRNA synthesis by RdRPs. Argonaute proteins (termed WAGOs) bind secondary siRNAs and: (1) target homologous RNAs for transcriptional and translational silencing29,34,63,64, as well as (2) direct the cleavage and de novo pUGylation of additional mRNAs. In this way, cycles of pUG RNA-based siRNA production and siRNA-directed mRNA pUGylation maintain silencing over time and across generations. This model shows germline perinuclear condensates termed Mutator foci as the likely sites of pUG RNA biogenesis in germ cells for several reasons. RDE-3 localizes to Mutator foci17 and we show, in Fig. 3d, that endogenous pUG RNAs localize to Mutator foci. The fact that enzyme and enzyme product both localize to Mutator foci suggests that Mutator foci may be sites of RNA pUGylation. In addition, although pUG RNAs are still made in mut-16 mutants (Extended Data Fig. 6b–d), which lack Mutator foci, the levels of both dsRNA-triggered and endogenous pUG RNAs are misregulated. Thus, while RDE-3 still has enzymatic activity in the absence of Mutator foci, these perinuclear condensates are probably coordinating target recognition and pUGylation in wild-type worms. Indeed, both the endonuclease RDE-8, which cleaves mRNAs targeted by dsRNA12, and the RdRP RRF-117 also localize to Mutator foci, further suggesting that pUG RNA–siRNA cycling occurs in Mutator foci. Previous studies have shown that animals lacking RDE-3 still produce some 22G endo-siRNAs, including 22G siRNAs that associate with the Argonaute CSR-1 and whose biogenesis depends upon the RdRP EGO-126,65. Thus, EGO-1 may also produce some 22G siRNAs via a pUG RNA-independent mechanism. A previous study showed that, in rrf-1 mutants that lack germlines, sel-1 RNAi causes a small fraction of sel-1 mRNA fragments to be uridylated in a largely RDE-3-dependent manner in the soma12. These results suggest that, in somatic tissues, RDE-3 may add non-templated Us to the 3′ termini of mRNA fragments generated during RNAi. It was proposed that this uridylation may be important for turnover or decay of RNAi targets12. Our work, combined with this earlier data about RDE-3-dependent uridylation12, suggests two models. First, RDE-3 may possess two distinct catalytic activities: uridylation and pUGylation. According to this model, RDE-3 might add Us or UGs depending on context (for example, cell or tissue type, or developmental timing). Alternatively, the mRNA uridylation observed in the soma could depend on RDE-3 and the pUGylation system, but may be mediated by another, currently unknown, poly(U) polymerase.

Extended Data Fig. 10 pUG RNA shortening may act as a brake on TEI.

a, The gel shown is the same as in Fig. 5a, except that oma-1 pUG RNAs from the P0 generation are included for WT and rde-3(-) worms. Data are representative of three biologically independent experiments. b, oma-1 pUG RNA reads from MiSeq (n = 1 biological experiment) were mapped to oma-1 and the length of the oma-1 mRNA portion of each pUG RNA read was determined (y-axis). In the box plot, the box represents the interquartile range (IQR) and the centre line shows the median of lengths at the indicated generations after dsRNA treatment. The y-axis starts at the aug of the oma-1 mRNA. The whiskers extend to values below and above 1.5 × IQR from the first and third quartiles, respectively. Data beyond the end of the whiskers are outliers and plotted as points. These data support the results in a, showing that the oma-1 mRNA fragments appended to get pUG tails, and thus oma-1 pUG RNAs, get shorter in each generation during RNAi-triggered TEI. c, A ratchet model to explain pUG RNA shortening. pUG RNA shortening may be due to the 3′→5′ directionality of RdRPs, which, during the maintenance phase of pUG–siRNA cycling (see model in Extended Data Fig. 9), causes each turn of the pUG–siRNA cycle to trigger cleavage and pUGylation of target mRNAs at sites more 5′ than in the previous cycle. Eventually, pUG RNAs are too short to act as RdRP templates, cycling cannot be maintained and silencing ends. Additional support for the ratchet model comes from Fig. 5c, which shows that RNAi-triggered pUG RNAs are longer in MAGO12 mutant worms than in wild-type worms. Note: the P0 worms in Fig. 5c were exposed to dsRNA continuously from embryos to adulthood, when they were collected. These longer pUG RNAs are probably due to continued initiation of pUGylation triggered by the exogenously provided dsRNA without downstream pUG–siRNA cycling. In the absence of this cycling, pUG RNA shortening does not occur. Finally, a number of recent studies in C. elegans have reported transgenerational inheritance of acquired traits, which lasts three to four generations66,67,68,69,70,71. As shown in a, the expression of oma-1 RNAi–directed pUG RNAs also endures for three to four generations. These shared generational timescales of inheritance hint that the inheritance of acquired traits in C. elegans may be mediated by pUG RNAs whose generational ‘half-life’ is limited to three to four generations owing to the built-in brake on TEI provided by pUG RNA shortening.

Supplementary information

Supplementary Figure 1

This figure contains the uncropped gels.

Reporting Summary

Supplementary Table 1

pUG RNA sequencing data. This table contains oma-1 pUG RNA reads from sequencing performed on an Illumina MiSeq, our calculations of the accuracy of pUG tails (based on Sanger and Illumina sequencing) and Tc1 pUG RNAs sequenced using Sanger sequencing.

Supplementary Table 2

RNAs upregulated in rde-3(-) mutants. List of RNAs upregulated in rde-3(-) mutants (adjusted P value <0.05 and log2(fold change) >1.5).

Supplementary Table 3

Normalized peptide counts for LC-MS/MS. Peptide counts from LC-MS/MS were normalized to total peptides for each pull-down sample. Resulting normalized peptide counts are provided.

Supplementary Table 4

Small RNA reads mapping to oma-1. oma-1 small RNAs sequenced after oma-1(SNP) pUG and pGC RNA injections (with a no injection control) in either the injected generation or from the progeny of injected animals are provided.

Supplementary Table 5

Oligos, C. elegans strains and pUG RNAs used in this study.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Shukla, A., Yan, J., Pagano, D.J. et al. poly(UG)-tailed RNAs in genome protection and epigenetic inheritance. Nature 582, 283–288 (2020). https://doi.org/10.1038/s41586-020-2323-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-020-2323-8

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

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