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:

Mammalian PIWI–piRNA–target complexes reveal features for broad and efficient target silencing

Abstract

The PIWI-interacting RNA (piRNA) pathway is an adaptive defense system wherein piRNAs guide PIWI family Argonaute proteins to recognize and silence ever-evolving selfish genetic elements and ensure genome integrity. Driven by this intensive host–pathogen arms race, the piRNA pathway and its targeted transposons have coevolved rapidly in a species-specific manner, but how the piRNA pathway adapts specifically to target silencing in mammals remains elusive. Here, we show that mouse MILI and human HILI piRNA-induced silencing complexes (piRISCs) bind and cleave targets more efficiently than their invertebrate counterparts from the sponge Ephydatia fluviatilis. The inherent functional differences comport with structural features identified by cryo-EM studies of piRISCs. In the absence of target, MILI and HILI piRISCs adopt a wider nucleic-acid-binding channel and display an extended prearranged piRNA seed as compared with EfPiwi piRISC, consistent with their ability to capture targets more efficiently than EfPiwi piRISC. In the presence of target, the seed gate—which enforces seed–target fidelity in microRNA RISC—adopts a relaxed state in mammalian piRISC, revealing how MILI and HILI tolerate seed–target mismatches to broaden the target spectrum. A vertebrate-specific lysine distorts the piRNA seed, shifting the trajectory of the piRNA–target duplex out of the central cleft and toward the PAZ lobe. Functional analyses reveal that this lysine promotes target binding and cleavage. Our study therefore provides a molecular basis for the piRNA targeting mechanism in mice and humans, and suggests that mammalian piRNA machinery can achieve broad target silencing using a limited supply of piRNA species.

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: piRNA–target binding is more efficient in mammals than in invertebrates.
Fig. 2: Structure features of mammalian PIWI proteins.
Fig. 3: Structural basis for piRNA targeting.
Fig. 4: Biochemical characterization of target cleavage by MILI and HILI.
Fig. 5: Mutation analysis of mammalian-specific amino acid residue.

Similar content being viewed by others

Data availability

The cryo-EM density maps for HILI–piRNA, MILI–piRNA, MILI–piRNA–target (15 nt), EfPiwiN959K–piRNA and EfPiwiN959K–piRNA–target (16 nt) have been deposited in the Electron Microscopy Data Bank under accession codes EMD-33800, EMD-33817, EMD-33801, EMD-33809 and EMD-33797, respectively. Corresponding atomic models have been deposited in the Protein Data Bank under accession IDs 7YFX, 7YGN, 7YFY, 7YG6 and 7YFQ, respectively. The raw sequence reads for cleavage products by MILI, HILI and EfPiwi are available from NCBI under session PRJNA1018221. The following publicly available datasets are used in this study: piRNA reference GSE57420 and GSE122199, PDB 7KX7, 7KX9 and 5GUH. Source data are provided with this paper.

Code availability

The pipeline analyzing degradome sequencing data was deposited in GitHub (https://github.com/CMACH508/GuidewithTarget)54.

References

  1. Bourque, G. et al. Ten things you should know about transposable elements. Genome Biol. 19, 199 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Khurana, J. S. et al. Adaptation to P element transposon invasion in Drosophila melanogaster. Cell 147, 1551–1563 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Aravin, A. A., Hannon, G. J. & Brennecke, J. The Piwi-piRNA pathway provides an adaptive defense in the transposon arms race. Science 318, 761–764 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Parhad, S. S. & Theurkauf, W. E. Rapid evolution and conserved function of the piRNA pathway. Open Biol. 9, 180181 (2019).

    Article  CAS  PubMed  Google Scholar 

  5. Malone, C. D. & Hannon, G. J. Small RNAs as guardians of the genome. Cell 136, 656–668 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Jangam, D., Feschotte, C. & Betran, E. Transposable element domestication as an adaptation to evolutionary conflicts. Trends Genet. 33, 817–831 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Levin, H. L. & Moran, J. V. Dynamic interactions between transposable elements and their hosts. Nat. Rev. Genet. 12, 615–627 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Sinzelle, L., Izsvak, Z. & Ivics, Z. Molecular domestication of transposable elements: from detrimental parasites to useful host genes. Cell. Mol. Life Sci. 66, 1073–1093 (2009).

    Article  CAS  PubMed  Google Scholar 

  9. Ozata, D. M., Gainetdinov, I., Zoch, A., O’Carroll, D. & Zamore, P. D. PIWI-interacting RNAs: small RNAs with big functions. Nat. Rev. Genet. 20, 89–108 (2019).

    Article  CAS  PubMed  Google Scholar 

  10. Aravin, A. et al. A novel class of small RNAs bind to MILI protein in mouse testes. Nature 442, 203–207 (2006).

    Article  CAS  PubMed  Google Scholar 

  11. Girard, A., Sachidanandam, R., Hannon, G. J. & Carmell, M. A. A germline-specific class of small RNAs binds mammalian Piwi proteins. Nature 442, 199–202 (2006).

    Article  PubMed  Google Scholar 

  12. Grivna, S. T., Beyret, E., Wang, Z. & Lin, H. A novel class of small RNAs in mouse spermatogenic cells. Genes Dev. 20, 1709–1714 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Aravin, A. A., Sachidanandam, R., Girard, A., Fejes-Toth, K. & Hannon, G. J. Developmentally regulated piRNA clusters implicate MILI in transposon control. Science 316, 744–747 (2007).

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

    Article  CAS  PubMed  Google Scholar 

  16. Kofler, R. piRNA clusters need a minimum size to control transposable element invasions. Genome Biol. Evol. 12, 736–749 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  18. Brennecke, J. et al. Discrete small RNA-generating loci as master regulators of transposon activity in. Cell 128, 1089–1103 (2007).

    Article  CAS  PubMed  Google Scholar 

  19. Yu, T. et al. The piRNA response to retroviral invasion of the koala genome. Cell 179, 632–643 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Rosenkranz, D., Zischler, H. & Gebert, D. piRNAclusterDB 2.0: update and expansion of the piRNA cluster database. Nucleic Acids Res. 50, D259–D264 (2022).

    Article  CAS  PubMed  Google Scholar 

  21. Matsumoto, N. et al. Crystal structure of silkworm PIWI-clade Argonaute Siwi bound to piRNA. Cell 167, 484–497 (2016).

    Article  CAS  PubMed  Google Scholar 

  22. Yamaguchi, S. et al. Crystal structure of Drosophila Piwi. Nat. Commun. 11, 858 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Anzelon, T. A. et al. Structural basis for piRNA targeting. Nature 597, 285–289 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Flores-Jasso, C. F., Salomon, W. E. & Zamore, P. D. Rapid and specific purification of Argonaute-small RNA complexes from crude cell lysates. RNA 19, 271–279 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Schirle, N. T. & MacRae, I. J. The crystal structure of human Argonaute2. Science 336, 1037–1040 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Bartel, D. P. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell 116, 281–297 (2004).

    Article  CAS  PubMed  Google Scholar 

  27. Lewis, B. P., Burge, C. B. & Bartel, D. P. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 120, 15–20 (2005).

    Article  CAS  PubMed  Google Scholar 

  28. Schirle, N. T., Sheu-Gruttadauria, J. & MacRae, I. J. Structural basis for microRNA targeting. Science 346, 608–613 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Wang, Y. et al. Structure of an argonaute silencing complex with a seed-containing guide DNA and target RNA duplex. Nature 456, 921–926 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Wang, Y. et al. Nucleation, propagation and cleavage of target RNAs in Ago silencing complexes. Nature 461, 754–761 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Arif, A. et al. GTSF1 accelerates target RNA cleavage by PIWI-clade Argonaute proteins. Nature 608, 618–625 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Wee, L. M., Flores-Jasso, C. F., Salomon, W. E. & Zamore, P. D. Argonaute divides its RNA guide into domains with distinct functions and RNA-binding properties. Cell 151, 1055–1067 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Gainetdinov, I. et al. Relaxed targeting rules help PIWI proteins silence transposons. Nature 619, 394–402 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Wang, W. et al. The initial uridine of primary piRNAs does not create the tenth adenine that is the hallmark of secondary piRNAs. Mol. Cell 56, 708–716 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Wang, W. et al. Slicing and binding by Ago3 or Aub trigger Piwi-bound piRNA production by distinct mechanisms. Mol. Cell 59, 819–830 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Klum, S. M., Chandradoss, S. D., Schirle, N. T., Joo, C. & MacRae, I. J. Helix-7 in Argonaute2 shapes the microRNA seed region for rapid target recognition. EMBO J. 37, 75–88 (2018).

    Article  CAS  PubMed  Google Scholar 

  37. Dai, S. et al. A family of C. elegans VASA homologs control Argonaute pathway specificity and promote transgenerational silencing. Cell Rep. 40, 111265 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. De Fazio, S. et al. The endonuclease activity of Mili fuels piRNA amplification that silences LINE1 elements. Nature 480, 259–263 (2011).

    Article  PubMed  Google Scholar 

  39. Dowling, M. et al. In vivo PIWI slicing in mouse testes deviates from rules established in vitro. RNA 29, 308–316 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Vourekas, A. et al. Mili and Miwi target RNA repertoire reveals piRNA biogenesis and function of Miwi in spermiogenesis. Nat. Struct. Mol. Biol. 19, 773–781 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Gou, L. T. et al. Pachytene piRNAs instruct massive mRNA elimination during late spermiogenesis. Cell Res 24, 680–700 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Goh, W. S. et al. piRNA-directed cleavage of meiotic transcripts regulates spermatogenesis. Genes Dev. 29, 1032–1044 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Wu, P. H. et al. The evolutionarily conserved piRNA-producing locus pi6 is required for male mouse fertility. Nat. Genet. 52, 728–739 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Choi, H., Wang, Z. & Dean, J. Sperm acrosome overgrowth and infertility in mice lacking chromosome 18 pachytene piRNA. PLoS Genet. 17, e1009485 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Gou, L. T. et al. Ubiquitination-deficient mutations in human Piwi cause male infertility by impairing histone-to-protamine exchange during spermiogenesis. Cell 169, 1090–1104 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Wu, X. et al. The biogenesis and functions of piRNAs in human diseases. Mol. Ther. Nucleic Acids 21, 108–120 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Lei, J. & Frank, J. Automated acquisition of cryo-electron micrographs for single particle reconstruction on an FEI Tecnai electron microscope. J. Struct. Biol. 150, 69–80 (2005).

    Article  PubMed  Google Scholar 

  48. Jumper, J. et al. Highly accurate protein structure prediction with AlphaFold. Nature 596, 583–589 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Emsley, P. & Cowtan, K. Coot: model-building tools for molecular graphics. Acta Crystallogr D. Biol. Crystallogr 60, 2126–2132 (2004).

    Article  PubMed  Google Scholar 

  50. Adams, P. D. et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D. Biol. Crystallogr 66, 213–221 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Schrödinger, L., & DeLano, W. PyMOL http://www.pymol.org/pymol (2020).

  52. Goddard, T. D. et al. UCSF ChimeraX: meeting modern challenges in visualization and analysis. Protein Sci. 27, 14–25 (2018).

    Article  CAS  PubMed  Google Scholar 

  53. Lu, W. P. & Fei, L. A logarithmic approximation to initial rates of enzyme reactions. Anal. Biochem. 316, 58–65 (2003).

    Article  CAS  PubMed  Google Scholar 

  54. Zeng, L. GuidewithTarget (Degradome Sequencing Bioinformatics Pipeline). GitHub https://github.com/CMACH508/GuidewithTarget (2023).

  55. Piuco, R. & Galante, P. A. piRNAdb: a piwi-interacting RNA database. Preprint at BioRxiv https://doi.org/10.1101/2021.09.21.461238 (2021).

  56. Langmead, B., Trapnell, C., Pop, M. & Salzberg, S. L. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 10, R25 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Chen, S., Zhou, Y., Chen, Y. & Gu, J. fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34, i884–i890 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  58. 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 

  59. Camacho, C. et al. BLAST+: architecture and applications. BMC Bioinf. 10, 421 (2009).

    Article  Google Scholar 

  60. Danecek, P. et al. Twelve years of SAMtools and BCFtools. Gigascience 10, giab008 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank C. C. Mello, P. D. Zamore, D. Li and H. Gao for insightful suggestions; H. Yu (Westlake University) for critical proofreading of the manuscript; I. Gainetdinov and P. D. Zamore for guidance in analyzing Degradome Sequencing data, members of Shen and Wu laboratories for discussions; the Cryo-EM Facility of Westlake University for providing support on cryo-EM data collection; the Key Laboratory of Zhejiang Province for Aptamers and Theranostics for their technical support in the SPR assay and the Mass Spectrometry and Metabolomics Core Facility of Westlake University for protein analysis. This work was supported by Westlake Education Foundation, Zhejiang Provincial Foundation of China (2021R01013) and National Natural Science Foundation of China (NSFC32070628) to E.-Z.S., National Natural Science Foundation of China (32271261), Zhejiang Provincial Natural Science Foundation of China (LR22C050003), Westlake University (1011103860222B1) and Institutional Startup Grant from the Westlake Education Foundation (101486021901) to J.W. and the National Natural Science Foundation of China (NO. 92068103) and Westlake Laboratory of Life Science to C.-Q.S.

Author information

Authors and Affiliations

Authors

Contributions

E.-Z.S. conceptualized the study. Zhiqing Li, Zhenzhen Li, Y.Z., L. Zhou, Q.X., L.L., L. Zeng, J.X., H.N., J.Z., Q.Y., D.L., M.G., Y.H., S.T., Z.Z., C.-Q.S., J.W. and E.-Z.S. performed the investigations. E.-Z.S. wrote the original draft. Z.Z., C.-Q.S., J.W. and E.-Z.S. reviewed and edited the manuscript.C.-Q.S., J.W. and E.-Z.S. supervised the project.

Corresponding authors

Correspondence to Chun-Qing Song, Jianping Wu or En-Zhi Shen.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Structural & Molecular Biology thanks Claus Kuhn and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available. Sara Osman was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

Additional information

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

Extended data

Extended Data Fig. 1 Retrotransposon and piRNA cluster in invertebrates and mammals.

a,b, Ratio of retrotransposon (a) and piRNA cluster (b) of invertebrate and mammalian genomes.

Extended Data Fig. 2 PIWI family tree and preparation of Piwi-piRNA complexes.

a, Phylogram of the PIWI proteins shows that MILI, HILI, and EfPiwi proteins belong to the same branch. Branch lengths are shown above the line. b,c, Size-exclusion chromatography profile of the purified MILI-piRNA complex (arrow) (b) and HILI-piRNA complex (arrow) (c) using a modified oligo capture approach. d,e, SDS-PAGE analysis of the purified MILI-piRNA complex (d) and HILI-piRNA complex (e). Gels were stained with Coomassie blue (representative of protein purifications performed at least ten times independently). f,g, Guide RNA-mediated cleavage assay with MILI (f) or HILI (g). MILI or HILI loaded with or without synthetic guide RNAs were incubated with 5′-FAM-labeled fully matched (FM) target RNAs or non-complementary (NC) target RNAs for 1 h at 37 °C. Products were resolved on a denaturing urea-PAGE gel alongside synthetic 22-nt marker and 1-nt RNA ladder. Similar results are seen in three experiments. h, piRNA and target RNA sequences used in f, g. The fluorescein amidite (FAM) modification in the target strand is highlighted in green, and the cleavage site is indicated with a red arrow. The cleaved 5′ fragment with FAM label is 24 nucleotides.

Source data

Extended Data Fig. 3 EMSA of piRISC binding to target RNAs and kinetics of Ago2-miRNA target binding.

a, Representative PAGE of the protein-guide RNA complexes binding to another fully matched (FM) target RNA. Non-complementary (NC) target is a negative control. Similar results are seen in two experiments. b, The guide and target RNAs used in this experiment. c, Equilibrium dissociation constant (KD) of g2-g8 complementary target for protein-guide RNA complexes. Data are mean ± s.d. from independent triplicates. d, SPR traces showing binding kinetics of g2-g8 complementary target for Ago2-miRNA binary complex. e, SPR sensorgrams of non-complementary target RNA binding to Ago2-miRNA complex.

Extended Data Fig. 4 Cryo-EM analysis of MILI-piRNA, MILI-piRNA-target (15 nt).

a,b, Data processing workflow of cryo-EM analysis of MILI-piRNA (a) and MILI-piRNA-target (15 nt) (b). Representative motion-corrected dose-weighted micrographs were selected out of 4,602 and 8,107 micrographs for the two datasets, respectively. Scale bar of the micrographs: 50 nm. Box size of 2D class averages: 209 Å. Local resolution, angular distribution, gold-standard FSC curves of the final maps (resolution cut-off at FSC = 0.143), and model vs. map curves of the final maps against the final models (resolution cut-off at FSC = 0.5) were presented. Please note that an open state was also observed during 2D and 3D classifications in the MILI-piRNA-target (15 nt) dataset.

Extended Data Fig. 5 Cryo-EM analysis of HILI-piRNA.

Data processing workflow for HILI-piRNA cryo-EM analysis. Representative motion-corrected dose-weighted micrograph was selected out of 2,136 micrographs. Scale bar of the micrograph: 50 nm. Box size of 2D class averages: 209 Å. Local resolution, angular distribution, gold-standard FSC curves of the final maps (resolution cut-off at FSC = 0.143), and model vs. map curves of the final map against the final model (resolution cut-off at FSC = 0.5) were presented. The density of the bound piRNA in the selected 3D class (with a dotted box) was indicated by a red arrow.

Extended Data Fig. 6 Structural comparison of individual domains of MILI, HILI, EfPiwi and Siwi, and electron density map as well as structural basis for piRNA targeting in the non-seed region.

a, Superposition of the individual domains of MILI (pink), HILI (orange), EfPiwi (PDB: 7KX7) (aquamarine), and Siwi (PDB: 5GUH) (chartreuse). b, The density map (grey mesh) for the entire bound piRNA in MILI–piRISC is shown in the left panel. A close-up view of the EM density map (grey mesh) for the piRNA 5′ segment is shown in the right panel. Modeled guide RNAs are shown as red sticks. c, The density map (grey mesh) for the entire bound piRNA in HILI-piRISC is shown in the left panel. Two close-up views of the EM density map (grey mesh) for the piRNA 5′ segment and 3′ half are shown in the middle and right panels. Modeled guide RNAs are shown as red sticks. d,e, Non-seed interactions between piRNA-target RNA duplex and EfPiwi protein (d) and MILI protein (e). f, Relative target-binding ratios for MILI, HILI, and EfPiwi. The relative ratio here is mismatched target binding % over fully matched target binding %. Data are from independent duplicates.

Source data

Extended Data Fig. 7 Effect of GTSF1 on slicer activity of PIWI proteins and biochemical determination of target binding, target cleavage and mismatch tolerance of Siwi.

a, Coomassie blue-stained SDS-PAGE of purified recombinant Flag-tagged GTSF1 proteins from different species (representative of protein purifications performed three times independently). b, piRISC target cleavage assays in the presence of increasing amounts of GTSF1. Em, Ephydatia muelleri; Bm, Bombyx mori; Mm, Mus musculus; Hs, Homo sapiens. Note: The Ephydatia fluviatilis genome sequence is unavailable, so EfGFST1 remains unidentified. GTSF1 from the related sponge Ephydatia muelleri was therefore incubated with EfPiwi. Similar results are seen in three experiments. c, Representative native PAGE analysis of protein-guide RNA complexes binding to fully matched (FM) target RNA. NC, non-complementary (NC) target negative control. Similar results are seen in three experiments. d, Representative in vitro cleavage reactions in the presence of equimolar substrate (target RNA) and enzyme (Piwi-piRNA complexes). Similar results are seen in three experiments. e, Guide-target-pairing schematic for targets with 3-nt mismatch regions to guide RNA used in (f, g). f, Representative urea-PAGE showing the ability of EfPiwi, Siwi, MILI, and HILI to cleave targets with three consecutive mismatches between g2 and g22. Similar results are seen in two experiments. g, Quantification of the in vitro cleavage assays (right) showing changes in cleavage activity for three consecutive mismatches between g2 and g22.

Source data

Extended Data Fig. 8 Sequence alignment of vertebrate Ago3-like PIWI proteins.

The secondary structure of MmMILI is indicated above the sequences, color coded by domains as indicated in Fig. 2. Key residues are marked by triangles. Species abbreviations: Mm, Mus musculus (mouse); Hs, Homo sapiens (human); Mm, Macaca mulatta (macaque); Ec, Equus caballus (horse); Dr, Danio rerio (zebrafish); Xt, Xenopus tropicalis (frog); Ef, Ephydatia fluviatilis (sponge); Hv, Hydra vulgaris (hydra); Bm, Bombyx mori (silkworm); Dm, Drosophila melanogaster (fruit fly); Ba, Bicyclus anynana (wood nymph); So, Sitophilus oryzae (rice weevil).

Extended Data Fig. 9 Cryo-EM analysis of EfPiwiN959K-piRNA, EfPiwiN959K-piRNA-target (16 nt).

a,b, Data processing workflow of cryo-EM analysis of EfPiwiN959K-piRNA (a) and EfPiwiN959K-piRNA-target (16 nt) (b). Representative motion-corrected dose-weighted micrographs were selected out of 2,279 and 3,983 micrographs for the two datasets, respectively. Scale bar of the micrographs: 50 nm. Box size of 2D class averages: 209 Å. Local resolution, angular distribution, gold-standard FSC curves of the final maps (resolution cut-off at FSC = 0.143), and model vs. map curves of the final maps against the final models (resolution cut-off at FSC = 0.5) were presented.

Extended Data Fig. 10 Target binding and cleavage data by EfPiwi, MILI, HILI and Ago2 in vitro and analysis of Drosophila Ago3-catalyzed splicing of transcripts in vivo.

a, Guide-target pairing schematic for targets with increasing complementarity for target binding analysis in (b). b, Equilibrium dissociation constants (KD) of Argonaute-guide complexes binding target RNAs with increasing complementarity. Data points from biological duplicates are shown. c, Guide-target pairing schematic for targets with increasing complementarity for cleavage analysis in (d, e). d, Quantification of reaction products shown as fraction of target cleaved. Data are from two independent experiments. e, Representative urea-PAGE showing EfPiwi-, MILI-, HILI-, and Ago2-mediated cleavage of targets with increasing guide complementarity. Similar results are seen in two experiments. f, Schema of data preparation and preprocessing to identify Ago3-mediated cleavage products in Drosophila ovaries. Ago3-bound piRNAs (Ago3_IP.fa) and non-associated piRNA (Ago3_bg.fa) were identified and compared to potential 3′ cleavage products whose abundance decreased in mutant ago3 degradome sequencing libraries (mutago3; SRR1926158) compared to wild-type ago3 degradome sequencing libraries (wtago3; SRR1926188). See Methods for details. g, piRNA queries (Ago3_IP.fa/Ago3_bg.fa) were aligned to candidate targets whose cleaved 5′ ends are complementary to g2–g10 of a piRNA. Targets with extended pairing were identified from g2–g25. For each pairing arrangement g2–gX, the fraction of cleaved targets was calculated as the fraction of cleaved targets explained by Ago3-associated piRNAs minus the fraction explained by Ago3-non-associated piRNAs. h, Plot of the fraction of cleaved targets in fly ovaries for contiguous pairing from nucleotide g2–gX, as indicated. Ago3-non-associated piRNA were sampled 10 times.

Source data

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2, Figs. 1–3 and Methods.

Reporting Summary

Peer Review File

Source data

Source Data Fig. 1

Unprocessed gel images.

Source Data Fig. 3

Unprocessed gel images.

Source Data Fig. 4

Unprocessed gel images.

Source Data Fig. 5

Unprocessed gel images.

Source Data Extended Data Fig. 2

Unprocessed gel images.

Source Data Extended Data Fig. 6

Unprocessed gel images.

Source Data Extended Data Fig. 7

Unprocessed gel images.

Source Data Extended Data Fig. 10

Unprocessed gel images.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Z., Li, Z., Zhang, Y. et al. Mammalian PIWI–piRNA–target complexes reveal features for broad and efficient target silencing. Nat Struct Mol Biol (2024). https://doi.org/10.1038/s41594-024-01287-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1038/s41594-024-01287-6

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