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:

Auto-inhibition and activation of a short Argonaute-associated TIR-APAZ defense system

Abstract

Short prokaryotic Ago accounts for most prokaryotic Argonaute proteins (pAgos) and is involved in defending bacteria against invading nucleic acids. Short pAgo associated with TIR-APAZ (SPARTA) has been shown to oligomerize and deplete NAD+ upon guide-mediated target DNA recognition. However, the molecular basis of SPARTA inhibition and activation remains unknown. In this study, we determined the cryogenic electron microscopy structures of Crenotalea thermophila SPARTA in its inhibited, transient and activated states. The SPARTA monomer is auto-inhibited by its acidic tail, which occupies the guide-target binding channel. Guide-mediated target binding expels this acidic tail and triggers substantial conformational changes to expose the Ago–Ago dimerization interface. As a result, SPARTA assembles into an active tetramer, where the four TIR domains are rearranged and packed to form NADase active sites. Together with biochemical evidence, our results provide a panoramic vision explaining SPARTA auto-inhibition and activation and expand understanding of pAgo-mediated bacterial defense systems.

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: Auto-inhibition of SPARTA by an acidic C-terminal tail.
Fig. 2: Guide-target duplex binding activates SPARTA.
Fig. 3: Ago–Ago interaction induced by guide-target binding is the prerequisite for SPARTA activation.
Fig. 4: TIR rearrangement and packing shapes NAD+ active sites in SPARTA tetramer.
Fig. 5: Proposal for the mechanism of SPARTA auto-inhibition and activation.

Similar content being viewed by others

Data availability

Atomic coordinates, maps and structure factors of the reported cryo-EM structures have been deposited in the PDB under accession codes 8J84 (SPARTA-apo), 8J9G (monomeric SPARTA bound with guide-target, state 1: Monomer-GT1), 8J8H (monomeric SPARTA bound with guide-target, state 2: Monomer-GT2), 8J9P (SPARTA dimer bound with guide-target) and 8JAY (SPARTA tetramer bound with guide-target) and in the Electron Microscopy Data Bank under accession codes EMD-36059 (SPARTA-apo), 36095 (monomeric SPARTA bound with guide-target, state 1: Monomer-GT1), 36070 (monomeric SPARTA bound with guide-target, state 2: Monomer-GT2), 36114 (SPARTA dimer bound with guide-target) and 36138 (SPARTA tetramer bound with guide-target). The structures used for comparison in this study are available under PDB accession codes 5GQ9, 5UZB and 6O0Q. Source data are provided with this paper.

References

  1. Hall, T. M. T. Structure and function of argonaute proteins. Structure 13, 1403–1408 (2005).

    Article  CAS  PubMed  Google Scholar 

  2. Meister, G. Argonaute proteins: functional insights and emerging roles. Nat. Rev. Genet. 14, 447–459 (2013).

    Article  CAS  PubMed  Google Scholar 

  3. Peters, L. & Meister, G. Argonaute proteins: mediators of RNA silencing. Mol. Cell 26, 611–623 (2007).

    Article  CAS  PubMed  Google Scholar 

  4. Makarova, K. S., Wolf, Y. I., Van der Oost, J. & Koonin, E. V. Prokaryotic homologs of Argonaute proteins are predicted to function as key components of a novel system of defense against mobile genetic elements. Biol. Direct 4, 29 (2009).

    Article  PubMed  PubMed Central  Google Scholar 

  5. Swarts, D. et al. The evolutionary journey of Argonaute proteins. Nat. Struct. Mol. Biol. 21, 743–753 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Ryazansky, S., Kulbachinskiy, A. & Aravin, A. A. The expanded universe of prokaryotic Argonaute proteins. MBio 9, e01935–18 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  7. Hegge, J., Swarts, D. & Van der Oost, J. Prokaryotic Argonaute proteins: novel genome-editing tools? Nat. Rev. Microbiol. 16, 5–11 (2018).

    Article  CAS  PubMed  Google Scholar 

  8. Willkomm, S., Makarova, K. S. & Grohmann, D. DNA silencing by prokaryotic Argonaute proteins adds a new layer of defense against invading nucleic acids. FEMS Microbiol. Rev. 42, 376–387 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Koopal, B., Mutte, S. K. & Swarts, D. C. A long look at short prokaryotic Argonautes. Trends Cell Biol. 33, 605–618 (2022).

    Article  PubMed  Google Scholar 

  10. Koopal, B. et al. Short prokaryotic Argonaute systems trigger cell death upon detection of invading DNA. Cell 185, 1471–1486.e19 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Garb, J. et al. Multiple phage resistance systems inhibit infection via SIR2-dependent NAD+ depletion. Nat. Microbiol. 7, 1849–1856 (2022).

    Article  CAS  PubMed  Google Scholar 

  12. Zaremba, M. et al. Short prokaryotic Argonautes provide defence against incoming mobile genetic elements through NAD+ depletion. Nat. Microbiol. 7, 1857–1869 (2022).

    Article  CAS  PubMed  Google Scholar 

  13. Wang, X. et al. Structural insights into mechanisms of Argonaute protein-associated NADase activation in bacterial immunity. Cell Res. 33, 699–711 (2023).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Medzhitov, P., Preston-Hurlburt, C. & Janeway, C. A. Jr. A human homologue of the Drosophila Toll protein signals activation of adaptive immunity. Nature 388, 394–397 (1997).

    Article  CAS  PubMed  Google Scholar 

  15. Van de Weyer, A. L. et al. A species-wide inventory of NLR genes and alleles in Arabidopsis thaliana. Cell 178, 1260–1272 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  16. Newman, R. M., Salunkhe, P., Godzik, A. & Reed, J. C. Identification and characterization of a novel bacterial virulence factor that shares homology with mammalian Toll/interleukin-1 receptor family proteins. Infect. Immun. 74, 594–601 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Essuman, K. et al. TIR domain proteins are an ancient family of NAD+-consuming enzymes. Curr. Biol. 28, 421–430 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Ofir, G. et al. Antiviral activity of bacterial TIR domains via immune signaling molecules. Nature 600, 116–120 (2021).

    Article  CAS  PubMed  Google Scholar 

  19. Doron, S. et al. Systematic discovery of antiphage defense systems in the microbial pangenome. Science 359, eaar4120 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  20. Essuman, K. et al. Shared TIR enzymatic functions regulate cell death and immunity across the tree of life. Science 377, eabo0001 (2022).

    Article  CAS  PubMed  Google Scholar 

  21. Morehouse, B. R. et al. STING cyclic dinucleotide sensing originated in bacteria. Nature 586, 429–433 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Morehouse, B. R. et al. Cryo-EM structure of an active bacterial TIR–STING filament complex. Nature 608, 803–807 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Tal, N. et al. Cyclic CMP and cyclic UMP mediate bacterial immunity against phages. Cell 184, 5728–5739 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Hogrel, G. et al. Cyclic nucleotide-induced helical structure activates a TIR immune effector. Nature 608, 808–812 (2022).

    Article  CAS  PubMed  Google Scholar 

  25. Martin, R. et al. Structure of the activated ROQ1 resistosome directly recognizing the pathogen effector XopQ. Science 370, eabd9993 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ma, S. et al. Direct pathogen-induced assembly of an NLR immune receptor complex to form a holoenzyme. Science 370, eabe3069 (2020).

    Article  CAS  PubMed  Google Scholar 

  27. Ve, T., Williams, S. J. & Kobe, B. Structure and function of Toll/interleukin-1 receptor/resistance protein (TIR) domains. Apoptosis 20, 250–261 (2015).

    Article  CAS  PubMed  Google Scholar 

  28. Nimma, S. et al. Structural evolution of TIR-domain signalosomes. Front. Immunol. 12, 784484 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Burroughs, A. M., Iyer, L. M. & Aravind, L. Two novel PIWI families: roles in inter-genomic conflicts in bacteria and mediator-dependent modulation of transcription in eukaryotes. Biol. Direct 8, 13 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  30. Burroughs, A. M., Ando, Y. & Aravind, L. New perspectives on the diversification of the RNA interference system: insights from comparative genomics and small RNA sequencing. RNA 5, 141–181 (2014).

    CAS  PubMed  Google Scholar 

  31. Willkomm, S. et al. Structural and mechanistic insights into an archaeal DNA-guided Argonaute protein. Nat. Microbiol. 2, 17035 (2017).

    Article  CAS  PubMed  Google Scholar 

  32. Tuschl, T. & Patel, D. J. Crystal structure of A. aeolicus Argonaute, a site-specific DNA-guided endoribonuclease, provides insights into RISC-mediated mRNA cleavage. Mol. Cell 19, 405–419 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  33. Wang, Y. et al. Structure of the guide-strand-containing argonaute silencing complex. Nature 456, 209–213 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. 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 

  35. Wang, J. et al. Reconstitution and structure of a plant NLR resistosome conferring immunity. Science 364, eaav5870 (2019).

    Article  CAS  PubMed  Google Scholar 

  36. Shi, Y. et al. Structural basis of SARM1 activation, substrate recognition, and inhibition by small molecules. Mol. Cell 82, 1643–1659 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Sporny, M. et al. Structural basis for SARM1 inhibition and activation under energetic stress. eLife 9, e62021 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Ve, T. et al. Structural basis of TIR-domain-assembly formation in MAL- and MyD88-dependent TLR4 signaling. Nat. Struct. Mol. Biol. 24, 743–751 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Sheng, G. et al. Structure-based cleavage mechanism of Thermus thermophilus Argonaute DNA guide strand-mediated DNA target cleavage. Proc. Natl Acad. Sci. USA 111, 652–657 (2014).

    Article  CAS  PubMed  Google Scholar 

  40. Ma, J. B. et al. Structural basis for 5′-end-specific recognition of guide RNA by the A. fulgidus Piwi protein. Nature 434, 666–670 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Manakova, E. et al. Structural basis for sequence-specific recognition of guide and target strands by the Archaeoglobus fulgidus Argonaute protein. Sci. Rep. 13, 6123 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Golovinas, E. et al. Prokaryotic Argonaute from Archaeoglobus fulgidus interacts with DNA as a homodimer. Sci. Rep. 11, 4518 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Zeng, Z. et al. A short prokaryotic Argonaute activates membrane effector to confer antiviral defense. Cell Host Microbe 30, 930–943 (2022).

    Article  CAS  PubMed  Google Scholar 

  44. Kim, S. Y., Jung, Y. & Lim, D. Argonaute system of Kordia jejudonensis is a heterodimeric nucleic acid-guided nuclease. Biochem. Biophys. Res. Commun. 525, 755–758 (2020).

    Article  CAS  PubMed  Google Scholar 

  45. Mastronarde, D. N. Automated electron microscope tomography using robust prediction of specimen movements. J. Struct. Biol. 152, 36–51 (2005).

    Article  PubMed  Google Scholar 

  46. Punjani, A., Rubinstein, J. L., Fleet, D. J. & Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 14, 290–296 (2017).

    Article  CAS  PubMed  Google Scholar 

  47. Zheng, S. Q. et al. MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Nat. Methods 14, 331–332 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Kucukelbir, A., Sigworth, F. J. & Tagare, H. D. Quantifying the local resolution of cryo-EM density maps. Nat. Methods 11, 63–65 (2014).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Pettersen, E. F. et al. UCSF ChimeraX: structure visualization for researchers, educators, and developers. Protein Sci. 30, 70–82 (2021).

    Article  CAS  PubMed  Google Scholar 

  51. Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 66, 486–501 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Liebschner, D. et al. Macromolecular structure determination using X-rays, neutrons and electrons: recent developments in Phenix. Acta Crystallogr. D 75, 861–877 (2019).

    Article  CAS  Google Scholar 

  53. Williams, C. J. et al. MolProbity: more and better reference data for improved all-atom structure validation. Protein Sci. 27, 293–315 (2018).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We would like to thank the Instrument Analysis Center at Shanghai Jiao Tong University for cryo-EM data collection. This work is supported by the National Key Research and Development Program of China (2018YFA0902000, Y.X.), STI2030-Major Projects (2021ZD0203400, Y.X.), the National Natural Science Foundation of China (32271330, M.C.; 31970547, Y.X.; 32000889, M.C.; 82321005, Y.X.), the Natural Science Foundation of Jiangsu Province (BK20190552, Y.X.), the Project Program of the State Key Laboratory of Natural Medicines, China Pharmaceutical University (SKLNMZZ202014, Y.X.) and the Fundamental Research Funds for the Central Universities (2632023GR17, Z.L.).

Author information

Authors and Affiliations

Authors

Contributions

L.G. and P.H. purified proteins and performed in vitro reconstitution of SPARTA complexes. Z.L. and M.C. determined cryo-EM structures. L.G. performed NAD+ hydrolysis assay, gel filtration assay and EMSA assay to biochemically characterize SPARTA. P.Y. prepared the SPARTA mutants. M.C., Y.X., L.G., P.H., Z.L., Y.C.S. and M.L. analyzed the data. P.H. prepared the figures. M.C. and Y.X. conceived experiments, supervised the work and wrote the manuscript. All authors discussed the results and contributed to the final manuscript.

Corresponding authors

Correspondence to Meirong Chen or Yibei Xiao.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks Andrey Kulbachinskiy, Dinshaw Patel and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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 The reconstitution of SPARTA complexes.

a. NADase activity at 37 °C and 50 °C. The experiments were repeated in triplicates, and the data represents mean values ± s.d. b. The size exclusion chromatography of reconstituted SPARTA complexes.

Source data

Extended Data Fig. 2 Single-particle cryo-EM data processing.

Data processing scheme of SPARTA-apo (a), monomeric SPARTA bound with guide-target (Monomer-GT) (b) and SPARTA tetramer sample (c).

Extended Data Fig. 3 Cryo-EM density maps of the SPARTA complexes.

Maps of SPARTA-apo (a), monomeric SPARTA bound with guide-target (Monomer-GT1) (b), monomeric SPARTA bound with guide-target (Monomer-GT2) (c), SPARTA dimer (d) and SPARTA tetramer (e) are colored by local resolution, and gold-standard FSC of 0.143 resolution graphs is indicated for each complex.

Extended Data Fig. 4 The structure of SPARTA apo form.

a. CrtTIR-APAZ and CrtAgo are shown side-by-side with each domain highlighted. b. The surface representation of APAZ and N domain connected by L1 linker from TtAgo (PDB:5GQ9). c. The overall structure comparison of CrtSPARTA with TtAgo (PDB:5GQ9).

Extended Data Fig. 5 The affinity of SPARTA to RNA guide.

a. The size exclusion chromatography of wild-type CrtSPARTA and C-tail truncated SPARTA (SPARTA_ΔCtail). The absorption at 260 nm and 280 nm is monitored and shown in red and blue, respectively. b. The binding affinity quantified by Electrophoretic Mobility Shift Assay (EMSA). 1 μM guide RNA was incubated with increasing amount of protein. The complex and 21nt free guide were resolved in Native-PAGE. The samples derived from the same experiment and gels were processed in parallel. The intensity of the free guide bands was measured, and the binding affinity was then analyzed using GraphPad Prism 9.3.0. The intensity of target band was quantified from single gel for Kd calculation.

Extended Data Fig. 6 Conformational changes induced by guide-target binding and propagation.

a. The cartoon representation of helixentry withdrawal upon guide-target binding. b. D-loop flipping upon the propagation of guide-target. c. Structure superimposition of SPARTA-apo, Monomer-GT1, Monomer-GT2 and dimer shows consecutive and unidirectional movement of the D-loop. d. The conformational changes likely coupling C-tail/TIR release with guide-target propagation. The structure of Monomer-GT2 (light blue) is superimposed onto Monomer-GT1 (grey), and the regions of substantial conformational changes in Monomer-GT2 are colored in yellow, orange, blue, and magenta.

Extended Data Fig. 7 The size exclusion chromatography of SPARTA mutants.

The mutant of D-loop truncation (a), and the Ago-Ago interface mutant SPARTA_ERDAAA (b) and SPARTA_YKAA (c), incubated with guide RNA and target DNA.

Extended Data Fig. 8 The Cryo-EM maps showing the flexibility of the TIR domain.

Comparison of TIR density in Monomer-GT1 (a), Monomer-GT2 (b), SPARTA dimer (c), and SPARTA tetramer (d).

Extended Data Fig. 9 The typical assemblies of TIR domains.

a. The structure of TIR domain in SPARTA. The secondary structures are indicated. Eukaryotic TIR domains are primarily arranged either as ‘scaffold assembly’ (b) or ‘enzymatic assembly’ (c), corresponding to parallel or anti-parallel two-stranded assembly via distinct interfaces, respectively. The involved interfaces are indicated by the frames.

Supplementary information

Supplementary Information

Supplementary Tables 1–3 and Supplementary Fig. 1.

Reporting Summary

Source data

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

Guo, L., Huang, P., Li, Z. et al. Auto-inhibition and activation of a short Argonaute-associated TIR-APAZ defense system. Nat Chem Biol 20, 512–520 (2024). https://doi.org/10.1038/s41589-023-01478-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-023-01478-0

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