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:

Cyclic nucleotide-induced helical structure activates a TIR immune effector

An Author Correction to this article was published on 11 January 2023

This article has been updated

Abstract

Cyclic nucleotide signalling is a key component of antiviral defence in all domains of life. Viral detection activates a nucleotide cyclase to generate a second messenger, resulting in activation of effector proteins. This is exemplified by the metazoan cGAS–STING innate immunity pathway1, which originated in bacteria2. These defence systems require a sensor domain to bind the cyclic nucleotide and are often coupled with an effector domain that, when activated, causes cell death by destroying essential biomolecules3. One example is the Toll/interleukin-1 receptor (TIR) domain, which degrades the essential cofactor NAD+ when activated in response to infection in plants and bacteria2,4,5 or during programmed nerve cell death6. Here we show that a bacterial antiviral defence system generates a cyclic tri-adenylate that binds to a TIR–SAVED effector, acting as the ‘glue’ to allow assembly of an extended superhelical solenoid structure. Adjacent TIR subunits interact to organize and complete a composite active site, allowing NAD+ degradation. Activation requires extended filament formation, both in vitro and in vivo. Our study highlights an example of large-scale molecular assembly controlled by cyclic nucleotides and reveals key details of the mechanism of TIR enzyme activation.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: M. ketosireducens CBASS generates a cA3 second messenger to activate the TIR–SAVED NADase effector.
Fig. 2: Structure of the activated TIR–SAVED/cA3 assembly.
Fig. 3: TIR–SAVED oligomerization is required for NADase activity.
Fig. 4: Multimerization of TIR–SAVED generates a conserved composite TIR active site.

Similar content being viewed by others

Data availability

The electron microscopy data deposition D_1292120121 has been assigned the following accession codes: PDB ID 7QQK and EMD-14122. Source data are provided with this paper. All other data are provided in the Supplementary Information.

Change history

References

  1. Cai, X., Chiu, Y. H. & Chen, Z. J. The cGAS-cGAMP-STING pathway of cytosolic DNA sensing and signaling. Mol. Cell 54, 289–296 (2014).

    Article  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  3. Millman, A., Melamed, S., Amitai, G. & Sorek, R. Diversity and classification of cyclic-oligonucleotide-based anti-phage signalling systems. Nat. Microbiol. 5, 1608–1615 (2020).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  5. Wan, L. et al. TIR domains of plant immune receptors are NAD+-cleaving enzymes that promote cell death. Science 365, 799–803 (2019).

    Article  ADS  CAS  Google Scholar 

  6. Summers, D. W., Gibson, D. A., DiAntonio, A. & Milbrandt, J. SARM1-specific motifs in the TIR domain enable NAD+ loss and regulate injury-induced SARM1 activation. Proc. Natl Acad. Sci. USA 113, E6271–E6280 (2016).

    Article  ADS  CAS  Google Scholar 

  7. Kazlauskiene, M., Kostiuk, G., Venclovas, C., Tamulaitis, G. & Siksnys, V. A cyclic oligonucleotide signaling pathway in type III CRISPR-Cas systems. Science 357, 605–609 (2017).

    Article  ADS  CAS  Google Scholar 

  8. Niewoehner, O. et al. Type III CRISPR–Cas systems produce cyclic oligoadenylate second messengers. Nature 548, 543–548 (2017).

    Article  ADS  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  10. Athukoralage, J. S. & White, M. F. Cyclic oligoadenylate signalling and regulation by ring nucleases during type III CRISPR defence. RNA 27, 855–867 (2021).

    Article  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  13. Ka, D., Oh, H., Park, E., Kim, J. H. & Bae, E. Structural and functional evidence of bacterial antiphage protection by Thoeris defense system via NAD+ degradation. Nat. Commun. 11, 2816 (2020).

    Article  ADS  CAS  Google Scholar 

  14. Burroughs, A. M., Zhang, D., Schaffer, D. E., Iyer, L. M. & Aravind, L. Comparative genomic analyses reveal a vast, novel network of nucleotide-centric systems in biological conflicts, immunity and signaling. Nucl. Acids Res. 43, 10633–10654 (2015).

    Article  CAS  Google Scholar 

  15. Makarova, K. S. et al. Evolutionary and functional classification of the CARF domain superfamily, key sensors in prokaryotic antivirus defense. Nucl. Acids Res. 48, 8828–8847 (2020).

    Article  CAS  Google Scholar 

  16. Takeuchi, M. & Hatano, K. Proposal of six new species in the genus Microbacterium and transfer of Flavobacterium marinotypicum ZoBell and Upham to the genus Microbacterium as Microbacterium maritypicum comb. nov. Int. J. Syst. Bacteriol. 48, 973–982 (1998).

    Article  CAS  Google Scholar 

  17. Canto, C., Menzies, K. J. & Auwerx, J. NAD+ metabolism and the control of energy homeostasis: a balancing act between mitochondria and the nucleus. Cell Metab. 22, 31–53 (2015).

    Article  CAS  Google Scholar 

  18. Zhou, Y. et al. Determining the extremes of the cellular NAD(H) level by using an Escherichia coli NAD+-auxotrophic mutant. Appl. Environ. Microbiol. 77, 6133–6140 (2011).

    Article  ADS  CAS  Google Scholar 

  19. Lowey, B. & Kranzusch, P. J. CD-NTases and nucleotide second messenger signaling. Curr. Biol. 30, R1106–R1108 (2020).

    Article  CAS  Google Scholar 

  20. Grüschow, S., Athukoralage, J. S., Graham, S., Hoogeboom, T. & White, M. F. Cyclic oligoadenylate signalling mediates Mycobacterium tuberculosis CRISPR defence. Nucl. Acids Res. 47, 9259–9270 (2019).

    Article  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  22. Mirdita, M., Ovchinnikov, S. & Steinegger, M. ColabFold - making protein folding accessible to all. Nat. Methods 19, 679–682 (2022).

    Article  CAS  Google Scholar 

  23. Lowey, B. et al. CBASS immunity uses CARF-related effectors to sense 3'-5'- and 2'-5'-linked cyclic oligonucleotide signals and protect bacteria from phage infection. Cell 182, 38–49 (2020).

    Article  CAS  Google Scholar 

  24. Fatma, S., Chakravarti, A., Zeng, X. & Huang, R. H. Molecular mechanisms of the CdnG-Cap5 antiphage defense system employing 3',2'-cGAMP as the second messenger. Nature Commun. 12, 6381 (2021).

    Article  ADS  CAS  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  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  Google Scholar 

  27. 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  Google Scholar 

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

    Article  CAS  Google Scholar 

  29. Burdett, H., Hu, X., Rank, M. X., 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 (2021).

  30. Vajjhala, P. R., Ve, T., Bentham, A., Stacey, K. J. & Kobe, B. The molecular mechanisms of signaling by cooperative assembly formation in innate immunity pathways. Mol. Immunol. 86, 23–37 (2017).

    Article  CAS  Google Scholar 

  31. Morehouse, B. R. et al. Cryo-EM structure of an active bacterial TIR–STING filament complex. Nature https://doi.org/10.1038/s41586-022-04999-1 (2022).

  32. Grüschow, S., Adamson, C. S. & White, M. F. Specificity and sensitivity of an RNA targeting type III CRISPR complex coupled with a NucC endonuclease effector. Nucl. Acids Res. 49, 13122–13134 (2021).

  33. Rouillon, C., Athukoralage, J. S., Graham, S., Grüschow, S. & White, M. F. Investigation of the cyclic oligoadenylate signalling pathway of type III CRISPR systems. Methods Enzymol. 616, 191–218 (2019).

    Article  CAS  Google Scholar 

  34. Whiteley, A. T. et al. Bacterial cGAS-like enzymes synthesize diverse nucleotide signals. Nature 567, 194–199 (2019).

    Article  ADS  CAS  Google Scholar 

  35. Pernot, P. et al. Upgraded ESRF BM29 beamline for SAXS on macromolecules in solution. J. Synchrotron Radiat. 20, 660–664 (2013).

    Article  CAS  Google Scholar 

  36. Round, A. et al. BioSAXS Sample Changer: a robotic sample changer for rapid and reliable high-throughput X-ray solution scattering experiments. Acta Crystallogr. D 71, 67–75 (2015).

    Article  CAS  Google Scholar 

  37. Costa, L. et al. Combined small angle X-ray solution scattering with atomic force microscopy for characterizing radiation damage on biological macromolecules. BMC Struct. Biol. 16, 18 (2016).

    Article  Google Scholar 

  38. Hajizadeh, N. R., Franke, D., Jeffries, C. M. & Svergun, D. I. Consensus Bayesian assessment of protein molecular mass from solution X-ray scattering data. Sci. Rep. 8, 7204 (2018).

    Article  ADS  Google Scholar 

  39. Manalastas-Cantos, K. et al. ATSAS 3.0: expanded functionality and new tools for small-angle scattering data analysis. J. Appl. Crystallogr. 54, 343–355 (2021).

    Article  CAS  Google Scholar 

  40. Svergun, D., Barberato, C. & Koch, M. H. J. CRYSOL - a program to evaluate x-ray solution scattering of biological macromolecules from atomic coordinates. J. Appl. Crystallogr. 28, 768–773 (1995).

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

  42. 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  Google Scholar 

  43. Rubinstein, J. L. & Brubaker, M. A. Alignment of cryo-EM movies of individual particles by optimization of image translations. J. Struct. Biol. 192, 188–195 (2015).

    Article  Google Scholar 

  44. Punjani, A. & Fleet, D. J. 3D variability analysis: resolving continuous flexibility and discrete heterogeneity from single particle cryo-EM. J. Struct. Biol. 213, 107702 (2021).

    Article  CAS  Google Scholar 

  45. Casanal, A., Lohkamp, B. & Emsley, P. Current developments in Coot for macromolecular model building of electron cryo-microscopy and crystallographic data. Protein Sci. 29, 1069–1078 (2020).

    Article  CAS  Google Scholar 

  46. Nicholls, R. A., Tykac, M., Kovalevskiy, O. & Murshudov, G. N. Current approaches for the fitting and refinement of atomic models into cryo-EM maps using CCP-EM. Acta Crystallogr. D 74, 492–505 (2018).

    Article  CAS  Google Scholar 

  47. 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 

Download references

Acknowledgements

This work was financed by the Biotechnology and Biological Sciences Research Council (references BB/S000313 and BB/T004789) and a European Research Council Advanced Grant (grant number 101018608) to M.F.W. We thank J. Athukoralage, T. Gloster and S. McQuarrie for discussions. We thank M. Tully for assistance in using beamline BM29 and acknowledge the European Synchrotron Radiation Facility for provision of synchrotron radiation facilities. We acknowledge the Scottish Centre for Macromolecular Imaging, M. Clarke and J. Streetley for assistance with cryo-EM experiments and access to instrumentation, financed by the Medical Research Council (MC_PC_17135) and the Scottish Funding Council (H17007). This work used the platforms of the Grenoble Instruct-ERIC centre (Integrated Structural Biology Grenoble; UAR 3518 CNRS-CEA-UGA-EMBL) within the Grenoble Partnership for Structural Biology, supported by the French Infrastructure for Integrated Structural Biology (ANR-10-INBS-0005-02) and the Grenoble Alliance for Integrated Structural Cell Biology, financed within the University Grenoble Alpes graduate school (Ecoles Universitaires de Recherche) CBH-EUR-GS (ANR-17-EURE-0003). Components of Figs. 1 and 3 were created with BioRender.com.

Author information

Authors and Affiliations

Authors

Contributions

G.H. planned, carried out and analysed the results of the biochemical and structural analyses and drafted the manuscript; A.G. collected and analysed electron microscopy data; S. Graham cloned, expressed and purified the WT and variant proteins; H.R. carried out the preliminary biochemical analysis of the effector protein; S. Grüschow analysed the cyclic nucleotides; Q.B. carried out the SAXS; L.S. planned and analysed the EM analyses; M.F.W. conceptualized and oversaw the project, obtained funding and analysed data along with the other authors. All authors contributed to the drafting and revision of the manuscript.

Corresponding authors

Correspondence to Laura Spagnolo or Malcolm F. White.

Ethics declarations

Competing interests

The authors declare no competing interests.

Peer review

Peer review information

Nature thanks Martin Jinek and Andrew Lovering for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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 Cyclase and NADase activity.

a, Cyclase products separated by thin layer chromatography. Radiolabelled ATP was mixed with 50 µM “cold” NTP (ATP, GTP, CTP, TTP, or UTP) and incubated with 20 µM cyclase for 2 h at 37 °C. As controls, cyclic oligoadenylate produced by the Type III CRISPR complex VmeCMR21 (C+) and the reaction without protein (C). b, Cyclic nucleotide screening for TIR-SAVED NADase activity. 0.5 µM TIR-SAVED was incubated without (apo condition) or with 5 µM cyclic: AMP, di-AMP, tri-AMP, tetra-AMP, hexa-AMP, GMP, di-GMP, AMP-GMP, di-AMP-GMP or ADP ribose. The fluorescence intensity (a.u.) at 90 min of reaction was plotted for each cyclic oligoadenylate. c, Initial rate of TIR NADase activity depends on cA3 concentration. 0.5 µM TIR-SAVED was incubated with 0, 0.1, 0.3, 0.5, 1.0, 2.5 µM cA3 and 500 µM ɛNAD+. d, Enzymatic characterization of TIR-SAVED NADase activity. Initial rate is plotted against a range of ɛNAD+ concentration. Experimental data were fitted using the Michaelis-Menten equation. e, TIR NADase activity is proportional to TIR-SAVED concentration. 0, 0.1, 0.2, 0.4, 0.5, 0.8, 1.0 µM TIR-SAVED were incubated with 1 µM cA3 and 2 mM ɛNAD+. Data are the means of triplicate experiments with standard deviation (b, c, d, e).

Source data

Extended Data Fig. 2 Plasmid immunity assay with TIR-SAVED activated by Type III CRISPR CSM.

Extended dataset from Fig. 1g. a, Schematic of the plasmid and competent cells used for the transformation. On the left, tetracycline resistant plasmid used for to transform the recipient cell. The tir-saved gene was cloned into the multiple cloning site 1 of the plasmid 2 and 3. In the plasmid 3, the catalytic residue E84 was mutated to prevent NADase activity. Recipient cell A and B both expressed the MtbCsm targeting the tetracycline resistant plasmid. In B, MtbCsm Cas10 was mutated (D630A) to prevent cOA production and is used as a control. In recipient cell C, the tetracycline resistant plasmid is not recognized as a target by MtbCsm. b, Cloning strategy of TIR-SAVED variants into the pRAT-Duet plasmid (left panel). To co-express two TIR-SAVED variants, one version was cloned into the multiple cloning site 1 (MSC-1) under the pBAD promotor while the R variant was cloned into the multiple cloning site 2 (MSC-2) under the T7 promoter. c, Transformed colonies after incubation overnight on induced plate. The different recipient cell/plasmid combination are annotated as “A.1” (recipient cell A transformed by plasmid 1). Results from two independent experiments with technical duplicates.

Extended Data Fig. 3 TIR-SAVED is a monomeric protein in diluted solution.

a, SEC-SAXS profile. Rg (radiation gyrus) was calculated based on Guinier approximation (see Material & Method section) and plotted against eluted volume. In red, the fractions used to estimate the global Rg and the molecular weight range of the protein. The theoretical MW value of the recombinant TIR-SAVED is 47.3 kDa. b, AF2 model fitted with the SEC-SAXS experimental dataset selected (red fractions in a.). c, Elution profile of TIR-SAVED after size exclusion chromatography in absence (blue) or presence (orange) of cA3 for a molar ratio of 1:1.5 (protein:cA3). Additional analysis of TIR-SAVED molecular weight by analytical gel filtration are shown in Supplemental Data Fig. 4.

Source data

Extended Data Fig. 4 Cryo-EM analysis of the TIR-SAVED filament.

a, Cryo-EM micrograph with single particle picked. b, Extract of single particle selected. c, 2D classes from both top and side views particles. d, 3D model of the filament. In blue, the extracted map used for refinement. e, Final refined 3D map coloured by local resolution (blue 2.5 A to red, 5 Å). f, Atomic model fitted into the density map for four TIR-SAVED/cA3 subunits.

Extended Data Fig. 5 Surface representation of one TIR-SAVED subunit.

a, position of the cA3 molecule bound to the SAVED protein. b, Representative map densities. Example map densities that allowed construction of the atomic model. The labels refer to the chain identities and residue numbers. The regions part of TIR-SAVED main features (BB loop, DE loop and cA3 binding pocket) are highlighted.

Extended Data Fig. 6 Sequence alignment of TIR-SAVED proteins.

Secondary structure features are displayed for the TIR domain to match the conserved features of protein TIR family. Conserved residues are shaded. Mutated residues used in biochemical experiments are highlighted.

Extended Data Fig. 7 Activity and oligomerisation of TIR mutants.

a, Dynamic Light Scattering analysis confirms the oligomerisation of Y115A and D45AL46A in presence of 1:1.5 protein:cA3 molar ratio. Experiment in technical triplicates for each condition. b, NADase activity comparison of the Y115A and D45AL46A mutants with the WT for a 3-fold serial dilution in protein concentrations (0.16, 0.5, 1.5, 4.5, 13.5 µM). 27 µM cA3 was used to activate TIR-SAVED incubated with 500 µM ɛNAD+ substrate. c, Enzymatic properties of Y115A mutant. Based on a NAD range concentration experiment, the initial rate of fluorescent ADP ribose production was calculated and fitted following a Michaelis-Menten model. In these experiments, as used previously for the WT, 0.5 µM Y115A was mixed in presence of 1 µM cA3 to hydrolyse 25, 75, 225, 500, 1000, 1500, 2000 and 3000 µM ɛNAD+. The right panel compares the final Michaelis-Menten parameters of Y115A with the WT protein. Data are the means of three experiments with Standard deviation indicated.

Source data

Extended Data Fig. 8 cA3 binding properties of TIR-SAVED mutant R388E and W394A.

a, Schematic of TIR-SAVED mutations. b, Dynamic Light Scattering profile of TIR-SAVED variants in presence of 1:1.5 protein:cA3 molar ratio. c, NADase activity of E84Q and R388E compared to the WT TIR-SAVED enzyme. A range of protein concentration (0, 0.16, 0.5, 1.5, 4, 13.5 µM) was incubated with 27 µM cA3 and 500 µM ɛ-NAD+ for 60 min. Data are means plotted with standard deviations for triplicate experiments. d, The K199E/W394A variant is catalytically dead. The protein (1.5 µM) was incubated with increased cA3 concentration (from 0.03 to 13.5 µM) and 500 µM ɛ-NAD+. Data are means from duplicate experiments. e, Comparison of the thermal denaturation profile of TIR-SAVED in presence of three concentrations of cA3 indicated as protein:cA3 molar ratio of 1:0.2, 1:1, 1:5. The melting temperature Tm for the condition 1:0 and 1:5 was plotted for each protein (right panel). Experiments were done in triplicates with three measures for each temperature.

Source data

Extended Data Fig. 9 Combination of TIR-SAVED variants.

a, Purification by gel filtration of the dimeric TIR-SAVED variant in comparison to the wild-type apo protein. The R388E variant (R) was incubated with K199E/W394A (KW) in presence of cA3. b, Analysis of the eluted fraction from (a) by native PAGE. As controls, separated TIR-SAVED were loaded in absence or presence of cA3. c, The combination of R+KW is catalytically dead. The R variant (0.25 µM) was incubated with increased KW concentration (from 0.125 to 2 µM) in presence of 4.5 µM cA3 and 500 µM ɛ-NAD+. d, Kinetic analysis of subunit mixing experiments over 60 min. The WT+E and R+E combinations yielded similar activation kinetics. At later time points, weak activation was observed for the KW+E combination, whilst the E and KW controls showed no activity. In each mixing experiment, the indicated variant (0.25 µM) was incubated with 8 µM E variant in presence of 16.5 µM cA3. Created with BioRender.com.

Source data

Extended Data Table 1 Cryo-EM data collection, refi nement and validation statistics

Supplementary information

Supplementary Information

This file contains Supplementary Figs. 1–6, legends for Supplementary Videos 1–7, and Supplementary Tables 1 and 2.

Reporting Summary

Peer Review File

Supplementary Video 1

Dynamic of the two tiers TIR–SAVED/cA3 filament, eigenvector 1. Video associated with the Supplementary Fig. 5. A prominent bending of the long spring-shaped assembly was observed along the long axis.

Supplementary Video 2

Dynamic of the two tiers TIR–SAVED/cA3 filament, eigenvector 2. Video associated with the Supplementary Fig. 5. A prominent bending of the long spring-shaped assembly was observed along the long axis.

Supplementary Video 3

Dynamic of the two tiers TIR–SAVED/cA3 filament, eigenvector 3. Video associated with the Supplementary Fig. 5. A prominent bending of the long spring-shaped assembly was observed along the long axis.

Supplementary Video 4

Dynamic of the one tier TIR-SAVED/cA3 filament, eigenvector 1. Video associated with the Supplementary Fig. 6. A slight change in the pitch of one round of the spring-shaped assembly was observed along the long axis.

Supplementary Video 5

Dynamic of the one tier TIR–SAVED/cA3 filament, eigenvector 2. Video associated with the Supplementary Fig. 6. A slight change in the pitch of one round of the spring-shaped assembly was observed along the long axis.

Supplementary Video 6

Dynamic of the one tier TIR–SAVED/cA3 filament, eigenvector 3. Videos associated with the Supplementary Fig. 6. A slight change in the pitch of one round of the spring-shaped assembly was observed along the long axis.

Supplementary Video 7

cA3 binding by two TIR–SAVED subunits. The cA3 is located into the binding pocket of TIR–SAVED 1 (pink) closed by the tail side of TIR–SAVED 2 (green). Video associated with Fig. 2e.

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

Hogrel, G., Guild, A., Graham, S. et al. Cyclic nucleotide-induced helical structure activates a TIR immune effector. Nature 608, 808–812 (2022). https://doi.org/10.1038/s41586-022-05070-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-022-05070-9

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