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:

Target ssDNA activates the NADase activity of prokaryotic SPARTA immune system

Abstract

Argonaute proteins (Agos), which use small RNAs or DNAs as guides to recognize complementary nucleic acid targets, mediate RNA silencing in eukaryotes. In prokaryotes, Agos are involved in immunity: the short prokaryotic Ago/TIR–APAZ (SPARTA) immune system triggers cell death by degrading NAD+ in response to invading plasmids, but its molecular mechanisms remain unknown. Here we used cryo-electron microscopy to determine the structures of inactive monomeric and active tetrameric Crenotalea thermophila SPARTA complexes, revealing mechanisms underlying SPARTA assembly, RNA-guided recognition of target single-stranded DNA (ssDNA) and subsequent SPARTA tetramerization, as well as tetramerization-dependent NADase activation. The small RNA guides Ago to recognize its ssDNA target, inducing SPARTA tetramerization via both Ago- and TIR-mediated interactions and resulting in a two-stranded, parallel, head-to-tail TIR rearrangement primed for NAD+ hydrolysis. Our findings thus identify the molecular basis for target ssDNA-mediated SPARTA activation, which will facilitate the development of SPARTA-based biotechnological tools.

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: Overall architecture of monomeric C. thermophila SPARTAgRNA–target ssDNA complex.
Fig. 2: Recognition of the gRNA–ssDNA target duplex by the SPARTA complex.
Fig. 3: Overall architecture of the tetrameric SPARTAgRNA–ssDNA complex.
Fig. 4: Target ssDNA binding induces conformational changes in TIR domains.
Fig. 5: Formation of a SPARTA tetramer is mediated by pAgo and TIR domain interactions.
Fig. 6: Proposed mechanistic model of prokaryotic SPARTA complex activation in response to invading plasmid DNA.

Similar content being viewed by others

Data availability

The cryo-EM density maps have been deposited in the Electron Microscopy Data Bank (EMDB) under accession number EMD-35419 (monomeric SPARTAgRNA–ssDNA target complex); EMD-35420 (tetrameric SPARTAgRNA–ssDNA target complex in state 1); EMD-35421 (tetrameric SPARTAgRNA–ssDNA target complex in state 2) and EMD-36843 (SPARTAgRNA binary complex). The atomic coordinates have been deposited in the PDB with accession number PDB-8IFK (monomeric SPARTAgRNA–ssDNA target complex); PDB-8IFL (tetrameric SPARTAgRNA–ssDNA target complex in state 1); PDB-8IFM (tetrameric SPARTAgRNA–ssDNA target complex in state 2) and PDB-8K34 (SPARTAgRNA binary complex). Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request. Source data are provided with this paper.

References

  1. Koopal, B., Mutte, S. K. & Swarts, D. C. A long look at short prokaryotic Argonautes. Trends Cell Biol. 33, 605–618 (2023).

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

    Article  CAS  PubMed  Google Scholar 

  3. Kuhn, C. D. & Joshua-Tor, L. Eukaryotic Argonautes come into focus. Trends Biochem. Sci 38, 263–271 (2013).

    Article  CAS  PubMed  Google Scholar 

  4. Sheu-Gruttadauria, J. & MacRae, I. J. Structural foundations of RNA silencing by Argonaute. J. Mol. Biol. 429, 2619–2639 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. 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 

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

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Liu, J. et al. Argonaute2 is the catalytic engine of mammalian RNAi. Science 305, 1437–1441 (2004).

    Article  CAS  PubMed  Google Scholar 

  9. Kwak, P. B. & Tomari, Y. The N domain of Argonaute drives duplex unwinding during RISC assembly. Nat. Struct. Mol. Biol. 19, 145–151 (2012).

    Article  CAS  PubMed  Google Scholar 

  10. Dykxhoorn, D. M. & Lieberman, J. The silent revolution: RNA interference as basic biology, research tool, and therapeutic. Annu. Rev. Med. 56, 401–423 (2005).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  Google Scholar 

  12. Koopal, B. et al. Short prokaryotic Argonaute systems trigger cell death upon detection of invading DNA. Cell 185, 1471–1486.e1419 (2022).

  13. Miyoshi, T., Ito, K., Murakami, R. & Uchiumi, T. Structural basis for the recognition of guide RNA and target DNA heteroduplex by Argonaute. Nat. Commun. 7, 11846 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. 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 

  15. 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. Wiley Interdiscip. Rev. RNA 5, 141–181 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. 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 

  17. Guo, L. et al. Structural basis for auto-inhibition and activation of a short prokaryotic Argonaute associated TIR–APAZ defense system. Preprint at bioRxiv https://doi.org/10.1101/2023.07.12.548734 (2023).

  18. Clabbers, M. T. B. et al. MyD88 TIR domain higher-order assembly interactions revealed by microcrystal electron diffraction and serial femtosecond crystallography. Nat. Commun. 12, 2578 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Toshchakov, V. Y. & Neuwald, A. F. A survey of TIR domain sequence and structure divergence. Immunogenetics 72, 181–203 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Xu, Y. et al. Structural basis for signal transduction by the Toll/interleukin-1 receptor domains. Nature 408, 111–115 (2000).

    Article  CAS  PubMed  Google Scholar 

  21. Essuman, K., Milbrandt, J., Dangl, J. L. & Nishimura, M. T. Shared TIR enzymatic functions regulate cell death and immunity across the tree of life. Science 377, eabo0001 (2022).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. 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 

  24. Shi, Y. et al. Structural basis of SARM1 activation, substrate recognition, and inhibition by small molecules. Mol. Cell 82, 1643–1659 e1610 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Manik, M. K. et al. Cyclic ADP ribose isomers: production, chemical structures, and immune signaling. Science 377, eadc8969 (2022).

    Article  CAS  PubMed  Google Scholar 

  26. 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 

  27. Burdett, H., Hu, X., Rank, M., Maruta, N. & Kobe, B. Self-association configures the NAD+-binding site of plant NLR TIR domains. Preprint at bioRxiv https://doi.org/10.1101/2021.10.02.462850 (2022).

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

    Article  CAS  PubMed  Google Scholar 

  29. Horsefield, S. et al. NAD+ cleavage activity by animal and plant TIR domains in cell death pathways. Science 365, 793–799 (2019).

    Article  CAS  PubMed  Google Scholar 

  30. 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 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. 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 

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

    Article  CAS  PubMed  Google Scholar 

  34. Patel, D. J., Yu, Y. & Jia, N. Bacterial origins of cyclic nucleotide-activated antiviral immune signaling. Mol. Cell 82, 4591–4610 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Tal, N. et al. Cyclic CMP and cyclic UMP mediate bacterial immunity against phages. Cell 184, 5728–5739 e5716 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

  37. Scheres, S. H. RELION: implementation of a Bayesian approach to cryo-EM structure determination. J. Struct. Biol. 180, 519–530 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. 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 

  39. 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 

  40. Rohou, A. & Grigorieff, N. CTFFIND4: fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 192, 216–221 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  41. Rosenthal, P. B. & Henderson, R. Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy. J. Mol. Biol. 333, 721–745 (2003).

    Article  CAS  PubMed  Google Scholar 

  42. Buchan, D. W., Minneci, F., Nugent, T. C., Bryson, K. & Jones, D. T. Scalable web services for the PSIPRED Protein Analysis Workbench. Nucleic Acids Res. 41, W349–W357 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. 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 

Download references

Acknowledgements

We thank the staff at Southern University of Science and Technology (SUSTech) Cryo-EM Center for assistance in data collection on the SUSTech Titan KRIOS cryo-electron microscope. This work was supported by the National Natural Science Foundation of China (grant no. 32270050 to N.J.), Guangdong and Shenzhen Natural Science Foundation (grant no. 2023A1515012420 and JCYJ20220530114409022 to N.J.), Key Project of Shenzhen Science and Technology Innovation Commission (grant no. JCYJ20220818100616034) and the Guangdong Provincial Science and Technology Innovation Council Grant (2017B030301018).

Author information

Authors and Affiliations

Authors

Contributions

J.-T.Z. and X.-Y.W. undertook biochemical studies, from sample preparation and purification, and biochemical assays. J.-T.Z also performed cryo-EM data collection data processing and structure refinement. N.C. contributed to the cryo-EM sample preparation, and R.T. participated in the helpful discussions regarding this project. N.J. directed the research. N.J. and J.-T.Z. wrote the manuscript with input from other authors.

Corresponding author

Correspondence to Ning Jia.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature Chemical Biology thanks 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 In vitro assembly of the SPARTAgRNA-ssDNA target complex.

(a) Size exclusion chromatography and SDS-PAGE profiles of the purified SPARTA complex. Data are representative of three independent experiments. (b) Size exclusion chromatography profile of the SPARTAgRNA and SPARTAgRNA-ssDNA complexes.

Source data

Extended Data Fig. 2 CryoEM reconstruction of monomeric and tetrameric SPARTA complexes in two different assembly states.

(a, d and g) Fourier Shell Correlation curve of the monomeric SPARTA complex (a) and the tetrameric SPARTA complex in assembly state 1(d) and state 2 (g). (b, e and h) Direction distribution plot of the monomeric SPARTA complex (b) and the tetrameric SPARTA complex in assembly state 1 (e) and state 2 (h). (c, f and i) Final three-dimensional reconstructed map of the monomeric SPARTA complex (c) and the tetrameric SPARTA complex in assembly state 1 (f) and state 2 (i), colored according to local resolution.

Extended Data Fig. 3 CryoEM densities of each domain and the gRNA- ssDNA target duplex in the monomeric SPARTAgRNA-target ssDNA complex structure.

Densities for indicated regions are shown in the context of the atomic model.

Extended Data Fig. 4 Structure and sequence comparisons of the SPARTA complex with related proteins.

(a) Structural comparison of the prokaryotic R. sphaeroides RsAgo complex (PDB 5AWH, in grey) and monomeric SPARTAgRNA-ssDNA complex (in color). (b) Multiple sequence alignment of the conserved catalytic tetrad motif in the PIWI domain of C. thermophila SPARTA, NP_036286.2 (Homo sapiens AGO2), WP_011011654.1 (Pyrococcus furiosus Argonaute) and WP_011174533.1 (Thermus thermophilus Argonaute). The catalytic motif DEDX is indicated with red font. (c) Structural comparison of the catalytic tetrad between T. thermophilus TtAgo (grey, PDB 3F73) and C. thermophila SPARTA (dark green). (d) Structural alignment of the R. sphaeroides RsAgo N domain, L1, L2 (PDB 5AWH) and C. thermophila SPARTA APAZ domain. (e) The C. thermophila SPARTA TIR domain (magenta) structurally resembles the H. sapiens MyD88 TIR domain (grey, PDB 7BEQ). (f) Ribbon representation of the SPARTA TIR domain. The conserved structural elements (αA–αE, βA–βE) and the BB loop critical for NADase activity are labeled.

Extended Data Fig. 5 TIR domain assembly patterns.

(a) Ribbon (left) and surface (right) comparision of tetrameric SPARTA complex in two different states. Tetrameric SPARTA in the second state can transfer to the first state via 180° rotation horizontally. (b) Ribbon and schematic representations of TIR domain assembly patterns in H. sapiens MyD88 (PDB 7BEQ), H. sapiens SARM1 (PDB 6O0Q) and AbTIR (PDB 7UXU).

Extended Data Fig. 6 Overall architecture of the SPARTAgRNA binary complex.

(a and b) Direction distribution plot (a) and Fourier Shell Correlation curve (b) of the SPARTAgRNA complex. (c and d) Surface (c) and ribbon (d) representations of the SPARTAgRNA complex. (e) CryoEM densities of the guide RNA in the SPARTAgRNA complex. (f) CryoEM densities of the sensor loop and sensor helix in the SPARTAgRNA complex. (g) Structural comparison of the SPARTAgRNA complex with the protomer2 of tetrameric SPARTA complex. Vector length correlates with the domain movement scale. (h) Structural comparison between the SPARTAgRNA binary and the protomer 2 of the tetrameric SPARTAgRNA-ssDNA complex. The C-tail region of SPARTAgRNA binary is indicated.

Extended Data Fig. 7 Conformational changes of the SPARTAgRNA complex upon target ssDNA binding.

(a) Structural comparison between the SPARTAgRNA binary and the protomer 2 of the tetrameric SPARTAgRNA-ssDNA complex. The sensor loop and sensor helix regions are indicated. (b and d) Effect of sensor loop and sensor helix mutants together with the C-tail deletion mutant on NADase activity. Data are representative of three independent experiments. Data are presented as mean values ± s.d. (c) SDS-PAGE profiles of SPARTA mutants. ΔC-tail indicates deletion of the C-tail in the APAZ domain. KHR and KE mutants indicate mutation of residues K354, H358, R361 in the APAZ domain and K325, E327 in the PIWI domain to alanines, respectively. Loop GS linker indicates the replacement of Y321PIWI to Y328PIWI with a GS linker. A MBP tag is fused to the N terminal of TIR-APAZ subunit in all SPARTA mutants. Data are representative of three independent experiments.

Source data

Extended Data Fig. 8 Structural comparison of the MID and TIR domains of SPARTA and other proteins.

(a) Structural comparison of the MID domains of the monomeric SPARTA complex (grey) and the tetrameric SPARTA complex (yellow). (b) Multiple sequence alignment of the loop in the SPARTA MID domain and other Agos. The aligned sequences are from WP_109649955.1 (Maribacter polysiphoniae Argonaute, MapAgo), ABP72561.1 (Rhodobacter sphaeroides Argonaute, RsAgo), WP_011011654.1 (Pyrococcus furiosus Argonaute, PfAgo), NP_036286.2 (human AGO2), WP_010880937.1 (Aquifex aeolicus Argonaute, AaAgo), WP_011174533.1 (Thermus thermophilus Argonaute, TtAgo), WP_014295921.1 (Marinitoga piezophila Argonaute, MpAgo). (c) Structural alignment of the loop in the SPARTA MID domain with T. thermophilus TtAgo (grey, PDB 3HVR). (d) Docking of NAD+ into the TIR domains of tetrameric SPARTA, and superimposition of SPARTA TIRs with AbTIR (PDB 7UXU). (e) The putative NAD+ binding pocket in the SPARTA TIR domains superimposed with AbTIR.

Extended Data Fig. 9 Size exclusion chromatography analysis of SPARTA mutants.

(a) Size exclusion chromatography (left) and SDS-PAGE profiles (right) of SPARTA mutants. E133-D137 (GSGSG) indicates the replacement of loop (133-EERVD-137) with a GSGSG linker in the MID domain. G42P and E50A indicate substitution of residues G42 and E50 in the TIR domain into proline and alanine, respectively. A MBP tag is fused to the N terminal of TIR-APAZ subunit in all SPARTA mutants. Red asterisk indicates the individual MBP tag. Data are representative of three independent experiments. (b) The left panel indicates size exclusion chromatography analysis of SPARTA E77A mutant. Then the tetrameric SPARTAE77A-gRNA-ssDNA complex was reloaded for a second round of size exclusion chromatography analysis (right panel).

Source data

Supplementary information

Supplementary Information

Supplementary Fig. 1 and Tables 1 and 2.

Reporting Summary

Source data

Source Data Figs. 2c, 3a and 5c,e and Extended Data Fig. 7b,d

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed gels of Extended Data Fig. 1a.

Source Data Extended Data Fig. 7

Unprocessed gels of Extended Data Fig. 7c.

Source Data Extended Data Fig. 9

Unprocessed gels of Extended Data Fig. 9a.

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

Zhang, JT., Wei, XY., Cui, N. et al. Target ssDNA activates the NADase activity of prokaryotic SPARTA immune system. Nat Chem Biol 20, 503–511 (2024). https://doi.org/10.1038/s41589-023-01479-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41589-023-01479-z

This article is cited by

Search

Quick links

Nature Briefing Microbiology

Sign up for the Nature Briefing: Microbiology newsletter — what matters in microbiology research, free to your inbox weekly.

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