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:

A Legionella toxin exhibits tRNA mimicry and glycosyl transferase activity to target the translation machinery and trigger a ribotoxic stress response

Abstract

A widespread strategy employed by pathogens to establish infection is to inhibit host-cell protein synthesis. Legionella pneumophila, an intracellular bacterial pathogen and the causative organism of Legionnaires’ disease, secretes a subset of protein effectors into host cells that inhibit translation elongation. Mechanistic insights into how the bacterium targets translation elongation remain poorly defined. We report here that the Legionella effector SidI functions in an unprecedented way as a transfer-RNA mimic that directly binds to and glycosylates the ribosome. The 3.1 Å cryo-electron microscopy structure of SidI reveals an N-terminal domain with an ‘inverted L’ shape and surface-charge distribution characteristic of tRNA mimicry, and a C-terminal domain that adopts a glycosyl transferase fold that licenses SidI to utilize GDP–mannose as a sugar precursor. This coupling of tRNA mimicry and enzymatic action endows SidI with the ability to block protein synthesis with a potency comparable to ricin, one of the most powerful toxins known. In Legionella-infected cells, the translational pausing activated by SidI elicits a stress response signature mimicking the ribotoxic stress response, which is activated by elongation inhibitors that induce ribosome collisions. SidI-mediated effects on the ribosome activate the stress kinases ZAKα and p38, which in turn drive an accumulation of the protein activating transcription factor 3 (ATF3). Intriguingly, ATF3 escapes the translation block imposed by SidI, translocates to the nucleus and orchestrates the transcription of stress-inducible genes that promote cell death, revealing a major role for ATF3 in the response to collided ribosome stress. Together, our findings elucidate a novel mechanism by which a pathogenic bacterium employs tRNA mimicry to hijack a ribosome-to-nuclear signalling pathway that regulates cell fate.

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: SidI potently inhibits protein synthesis and binds to the cellular translation machinery.
Fig. 2: Structural characterization of SidI by cryo-EM reveals tRNA mimicry and glycosyl transferase activity.
Fig. 3: Coupling of tRNA mimicry and enzymatic activity yields SidI potency to target ribosomes and inhibit translation.
Fig. 4: Glycosyl transferase activity of SidI induces ribosome stalling and collisions.
Fig. 5: Transcriptional profiling reveals a conserved stress signature activated by SidI-mediated ribosome stress.
Fig. 6: ATF3 accumulation is regulated by kinase signalling and preferential mRNA translation during collided ribosome stress.
Fig. 7: ATF3 orchestrates a transcriptional programme that regulates cell fate.

Similar content being viewed by others

Data availability

Key resource information is available in Supplementary Table 9. Further information and requests for resources and reagents will be fulfilled by Shaeri Mukherjee. The authors declare that the data supporting the findings of this study are available within the article and its Supplementary Information files. The atomic models of BshA (PDB ID: 5D00), PimB′ (PDB ID: 3OKA) and EEF2 bound to the ribosome (PDB ID: 6GZ5) were downloaded from the PDB. The atomic model of SidI has been deposited into the PDB (PDB ID: 8JHU) and the EM map has been deposited into the EMDB (EMDB ID: EMD-36294). The MS data files for Supplementary Table 1 (raw and search results) have been deposited to the ProteomeXchange Consortium (http://proteomecentral.proteomexchange.org) via the PRIDE partner repository with the dataset identifier PXD034240. The MS data files for Supplementary Table 3 (raw and search results) have been deposited to the MassIVE repository with the dataset identifier MSV000091992. The RNA-seq datasets have been deposited in the GenBank database under the accession code GSE205648. The nucleotide sequence of ATF3 mRNA was downloaded from the NCBI GenBank database (NM_001674.3). ChIP–seq data were mined from the publicly available ENCODE 3 database and the tracks were generated by the ENCODE Transcription Factor ChIP–seq Processing Pipeline using UCSC Genome Browser. Methods documentation and full metadata for each track can be found at the ENCODE project portal using the ENCODE file accessions: K562, ENCFF937OKC; A549, ENCFF851UTY; HepG2, ENCFF137OEY; and liver cells, ENCFF146URA. Source data are provided with this paper.

Code availability

The MaxQuant and Protein Prospector (v6.3.23) software required for the analysis of raw MS data are available at https://www.maxquant.org/ and https://prospector.ucsf.edu/prospector/mshome.htm, respectively. Code for the in-house PAVA software is available upon request. For cryo-EM image processing and data analysis, the micrograph frames were aligned using MotionCor2 (ref. 73; RRID: SCR_016499). The contrast transfer function (CTF) parameters were estimated using GCTF74 (RRID: SCR_016500). Particles were picked using Gautomatch (developed by K. Zhang; https://www2.mrc-lmb.cam.ac.uk/download/gautomatch-053/) without a template. Particles were extracted using a 64-pixel box size and classified in two dimensions using Relion75 (RRID: SCR_016274). Particles were imported to cryoSPARC76 (RRID: SCR_016501) for homogeneous refinement and heterogeneous refinement. Structure prediction analysis was performed using RaptorX77 (RRID: SCR_018118). Models were manually adjusted in Coot79 (RRID: SCR_014222). Model refinement was performed in Phenix80 (RRID: SCR_014224) Molecular graphics and analyses were performed with the UCSF Chimera package82 (RRID: SCR_004097). The ICE CRISPR analysis tool from Synthego was used to analyse the KO efficiency of cell lines and is available at https://www.synthego.com/products/bioinformatics/crispr-analysis.

References

  1. Barry, K. C., Ingolia, N. T. & Vance, R. E. Global analysis of gene expression reveals mRNA superinduction is required for the inducible immune response to a bacterial pathogen. eLife https://doi.org/10.7554/eLife.22707 (2017).

  2. Belyi, Y. et al. Legionella pneumophila glucosyltransferase inhibits host elongation factor 1A. Proc. Natl Acad. Sci. USA 103, 16953–16958 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Fontana, M. F. et al. Secreted bacterial effectors that inhibit host protein synthesis are critical for induction of the innate immune response to virulent Legionella pneumophila. PLoS Pathog. 7, e1001289 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. McCusker, K. T., Braaten, B. A., Cho, M. W. & Low, D. A. Legionella pneumophila inhibits protein synthesis in Chinese hamster ovary cells. Infect. Immun. 59, 240–246 (1991).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Shen, X. et al. Targeting eEF1A by a Legionella pneumophila effector leads to inhibition of protein synthesis and induction of host stress response. Cell. Microbiol. 11, 911–926 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Moss, S. M. et al. A Legionella pneumophila kinase phosphorylates the hsp70 chaperone family to inhibit eukaryotic protein synthesis. Cell Host Microbe 25, 454–462 e456 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Belyi, Y., Tabakova, I., Stahl, M. & Aktories, K. Lgt: a family of cytotoxic glucosyltransferases produced by Legionella pneumophila. J. Bacteriol. 190, 3026–3035 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Guo, Z., Stephenson, R., Qiu, J., Zheng, S. & Luo, Z. Q. A Legionella effector modulates host cytoskeletal structure by inhibiting actin polymerization. Microbes Infect. 16, 225–236 (2014).

    CAS  PubMed  Google Scholar 

  9. Joseph, A. M. et al. The Legionella pneumophila metaeffector Lpg2505 (MesI) regulates SidI-mediated translation inhibition and novel glycosyl hydrolase activity. Infect. Immun. https://doi.org/10.1128/IAI.00853-19 (2020).

  10. Fontana, M. F., Shin, S. & Vance, R. E. Activation of host mitogen-activated protein kinases by secreted Legionella pneumophila effectors that inhibit host protein translation. Infect. Immun. 80, 3570–3575 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Joazeiro, C. A. P. Mechanisms and functions of ribosome-associated protein quality control. Nature Rev. Mol. Cell Biol. 20, 368–383 (2019).

    CAS  Google Scholar 

  12. Wu, C. C., Peterson, A., Zinshteyn, B., Regot, S. & Green, R. Ribosome collisions trigger general stress responses to regulate cell fate. Cell 182, 404–416 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Sinha, N. K. et al. EDF1 coordinates cellular responses to ribosome collisions. eLife https://doi.org/10.7554/eLife.58828 (2020).

  14. Costa-Mattioli, M. & Walter, P. The integrated stress response: from mechanism to disease. Science https://doi.org/10.1126/science.aat5314 (2020).

  15. Iordanov, M. S. et al. Ribotoxic stress response: activation of the stress-activated protein kinase JNK1 by inhibitors of the peptidyl transferase reaction and by sequence-specific RNA damage to the α-sarcin/ricin loop in the 28S rRNA. Mol. Cell. Biol. 17, 3373–3381 (1997).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Vind, A. C. et al. ZAKα recognizes stalled ribosomes through partially redundant sensor domains. Mol. Cell 78, 700–713 (2020).

    CAS  PubMed  Google Scholar 

  17. Olsnes, S., Refsnes, K. & Pihl, A. Mechanism of action of the toxic lectins abrin and ricin. Nature 249, 627–631 (1974).

    CAS  PubMed  Google Scholar 

  18. Moremen, K. W. & Haltiwanger, R. S. Emerging structural insights into glycosyltransferase-mediated synthesis of glycans. Nat. Chem. Biol. 15, 853–864 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Batt, S. M. et al. Acceptor substrate discrimination in phosphatidyl-myo-inositol mannoside synthesis: structural and mutational analysis of mannosyltransferase Corynebacterium glutamicum PimB′. J. Biol. Chem. 285, 37741–37752 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  20. Lu, W. et al. Structural basis of the action of glucosyltransferase Lgt1 from Legionella pneumophila. J. Mol. Biol. 396, 321–331 (2010).

    PubMed  Google Scholar 

  21. Flis, J. et al. tRNA translocation by the eukaryotic 80S ribosome and the impact of GTP hydrolysis. Cell Rep. 25, 2676–2688 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Song, H. et al. The crystal structure of human eukaryotic release factor eRF1—mechanism of stop codon recognition and peptidyl-tRNA hydrolysis. Cell 100, 311–321 (2000).

    CAS  PubMed  Google Scholar 

  23. Nissen, P. et al. Crystal structure of the ternary complex of Phe-tRNAPhe, EF-Tu, and a GTP analog. Science 270, 1464–1472 (1995).

    CAS  PubMed  Google Scholar 

  24. Blaha, G., Stanley, R. E. & Steitz, T. A. Formation of the first peptide bond: the structure of EF-P bound to the 70S ribosome. Science 325, 966–970 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Jank, T. et al. Domain organization of Legionella effector SetA. Cell. Microbiol. 14, 852–868 (2012).

    CAS  PubMed  Google Scholar 

  26. Levanova, N. et al. The Legionella effector LtpM is a new type of phosphoinositide-activated glucosyltransferase. J. Biol. Chem. 294, 2862–2879 (2019).

    CAS  PubMed  Google Scholar 

  27. Mukherjee, S. et al. Modulation of Rab GTPase function by a protein phosphocholine transferase. Nature 477, 103–106 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Juszkiewicz, S. et al. ZNF598 Is a quality control sensor of collided ribosomes. Mol. Cell 72, 469–481 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Ishimura, R., Nagy, G., Dotu, I., Chuang, J. H. & Ackerman, S. L. Activation of GCN2 kinase by ribosome stalling links translation elongation with translation initiation. eLife https://doi.org/10.7554/eLife.14295 (2016).

  30. Vogel, J. P., Andrews, H. L., Wong, S. K. & Isberg, R. R. Conjugative transfer by the virulence system of Legionella pneumophila. Science 279, 873–876 (1998).

    CAS  PubMed  Google Scholar 

  31. Roy, C. R., Berger, K. H. & Isberg, R. R. Legionella pneumophila DotA protein is required for early phagosome trafficking decisions that occur within minutes of bacterial uptake. Mol. Microbiol. 28, 663–674 (1998).

    CAS  PubMed  Google Scholar 

  32. Shao, S. et al. Decoding mammalian ribosome–mRNA states by translational GTPase complexes. Cell 167, 1229–1240 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Subramanian, A. et al. Gene set enrichment analysis: a knowledge-based approach for interpreting genome-wide expression profiles. Proc. Natl Acad. Sci. USA 102, 15545–15550 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Lipo, E. et al. 5′ Untranslated mRNA regions allow bypass of host cell translation inhibition by Legionella pneumophila. Infect. Immun. 90, e0017922 (2022).

    PubMed  Google Scholar 

  35. Garreau de Loubresse, N. et al. Structural basis for the inhibition of the eukaryotic ribosome. Nature 513, 517–522 (2014).

    CAS  PubMed  Google Scholar 

  36. Eferl, R. & Wagner, E. F. AP-1: a double-edged sword in tumorigenesis. Nat. Rev. Cancer 3, 859–868 (2003).

    CAS  PubMed  Google Scholar 

  37. Fonseca, G. J. et al. Diverse motif ensembles specify non-redundant DNA binding activities of AP-1 family members in macrophages. Nat. Commun. 10, 414 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Hai, T., Wolfgang, C. D., Marsee, D. K., Allen, A. E. & Sivaprasad, U. ATF3 and stress responses. Gene Expr. 7, 321–335 (1999).

    CAS  PubMed  Google Scholar 

  39. Scheuner, D. et al. Translational control is required for the unfolded protein response and in vivo glucose homeostasis. Mol. Cell 7, 1165–1176 (2001).

    CAS  PubMed  Google Scholar 

  40. Harding, H. P., Zhang, Y. & Ron, D. Protein translation and folding are coupled by an endoplasmic-reticulum-resident kinase. Nature 397, 271–274 (1999).

    CAS  PubMed  Google Scholar 

  41. Jiang, H. Y. et al. Activating transcription factor 3 is integral to the eukaryotic initiation factor 2 kinase stress response. Mol. Cell. Biol. 24, 1365–1377 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Sidrauski, C. et al. Pharmacological brake-release of mRNA translation enhances cognitive memory. eLife 2, e00498 (2013).

    PubMed  PubMed Central  Google Scholar 

  43. Noack, J. et al. Dynamic proteomics profiling of Legionella pneumophila infection unveils modulation of the host mitochondrial stress response pathway. Preprint at bioRxiv https://doi.org/10.1101/2020.05.19.105395 (2020).

  44. Shin, S. et al. Type IV secretion-dependent activation of host MAP kinases induces an increased proinflammatory cytokine response to Legionella pneumophila. PLoS Pathog. 4, e1000220 (2008).

    PubMed  PubMed Central  Google Scholar 

  45. Yang, J. et al. Design, synthesis, and structure-activity relationships of 1,2,3-triazole benzenesulfonamides as new selective leucine-zipper and sterile-α motif kinase (ZAK) inhibitors. J. Med. Chem. 63, 2114–2130 (2020).

    CAS  PubMed  Google Scholar 

  46. Pearlman, E., Jiwa, A. H., Engleberg, N. C. & Eisenstein, B. I. Growth of Legionella pneumophila in a human macrophage-like (U937) cell line. Microb. Pathog. 5, 87–95 (1988).

    CAS  PubMed  Google Scholar 

  47. Marra, A., Horwitz, M. A. & Shuman, H. A. The HL-60 model for the interaction of human macrophages with the Legionnaires’ disease bacterium. J. Immunol. 144, 2738–2744 (1990).

    CAS  PubMed  Google Scholar 

  48. Zink, S. D., Pedersen, L., Cianciotto, N. P. & Abu-Kwaik, Y. The Dot/Icm type IV secretion system of Legionella pneumophila is essential for the induction of apoptosis in human macrophages. Infect. Immun. 70, 1657–1663 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Lord, J. M., Roberts, L. M. & Robertus, J. D. Ricin: structure, mode of action, and some current applications. FASEB J. 8, 201–208 (1994).

    CAS  PubMed  Google Scholar 

  50. Shames, S. R. et al. Multiple Legionella pneumophila effector virulence phenotypes revealed through high-throughput analysis of targeted mutant libraries. Proc. Natl Acad. Sci. USA 114, E10446–E10454 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. McCloskey, A., Perri, K., Chen, T., Han, A. & Luo, Z. Q. The metaeffector MesI regulates the activity of the Legionella effector SidI through direct protein–protein interactions. Microbes Infect. 23, 104794 (2021).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Mailliot, J. et al. Crystal structures of the uL3 mutant ribosome: illustration of the importance of ribosomal proteins for translation efficiency. J. Mol. Biol. 428, 2195–2202 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Meskauskas, A., Russ, J. R. & Dinman, J. D. Structure/function analysis of yeast ribosomal protein L2. Nucleic Acids Res. 36, 1826–1835 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Diedrich, G. et al. Ribosomal protein L2 is involved in the association of the ribosomal subunits, tRNA binding to A and P sites and peptidyl transfer. EMBO J. 19, 5241–5250 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Schulze, H. & Nierhaus, K. H. Minimal set of ribosomal components for reconstitution of the peptidyltransferase activity. EMBO J. 1, 609–613 (1982).

    CAS  PubMed  PubMed Central  Google Scholar 

  56. Lawrence, M. G. et al. The extended loops of ribosomal proteins uL4 and uL22 of Escherichia coli contribute to ribosome assembly and protein translation. Nucleic Acids Res. 44, 5798–5810 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Gregory, S. T. & Dahlberg, A. E. Erythromycin resistance mutations in ribosomal proteins L22 and L4 perturb the higher order structure of 23 S ribosomal RNA. J. Mol. Biol. 289, 827–834 (1999).

    CAS  PubMed  Google Scholar 

  58. Zaman, S., Fitzpatrick, M., Lindahl, L. & Zengel, J. Novel mutations in ribosomal proteins L4 and L22 that confer erythromycin resistance in Escherichia coli. Mol. Microbiol. 66, 1039–1050 (2007).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Wong, W. et al. Mefloquine targets the Plasmodium falciparum 80S ribosome to inhibit protein synthesis. Nat. Microbiol. 2, 17031 (2017).

    PubMed  PubMed Central  Google Scholar 

  60. Mo, P., Wang, H., Lu, H., Boyd, D. D. & Yan, C. MDM2 mediates ubiquitination and degradation of activating transcription factor 3. J. Biol. Chem. 285, 26908–26915 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  61. Vu, T. T. M. & Varshavsky, A. The ATF3 transcription factor is a short-lived substrate of the Arg/N-degron pathway. Biochemistry 59, 2796–2812 (2020).

    CAS  PubMed  Google Scholar 

  62. Losick, V. P. & Isberg, R. R. NF-κB translocation prevents host cell death after low-dose challenge by Legionella pneumophila. J. Exp. Med. 203, 2177–2189 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Farbrother, P. et al. Dictyostelium transcriptional host cell response upon infection with Legionella. Cell. Microbiol. 8, 438–456 (2006).

    CAS  PubMed  Google Scholar 

  64. Li, Z. et al. The amoebal MAP kinase response to Legionella pneumophila is regulated by DupA. Cell Host Microbe 6, 253–267 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  65. Case, C. L. et al. Caspase-11 stimulates rapid flagellin-independent pyroptosis in response to Legionella pneumophila. Proc. Natl Acad. Sci. USA 110, 1851–1856 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Sol, A. et al. Legionella pneumophila translocated translation inhibitors are required for bacterial-induced host cell cycle arrest. Proc. Natl Acad. Sci. USA 116, 3221–3228 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Ibe, N. U., Subramanian, A. & Mukherjee, S. Non-canonical activation of the ER stress sensor ATF6 by Legionella pneumophila effectors. Life Sci. Alliance https://doi.org/10.26508/lsa.202101247 (2021).

  68. Treacy-Abarca, S. & Mukherjee, S. Legionella suppresses the host unfolded protein response via multiple mechanisms. Nat. Commun. 6, 7887 (2015).

    CAS  PubMed  Google Scholar 

  69. Hubber, A. & Roy, C. R. Modulation of host cell function by Legionella pneumophila type IV effectors. Ann. Rev. Cell Dev. Biol. 26, 261–283 (2010).

    CAS  Google Scholar 

  70. Hempstead, A. D. & Isberg, R. R. Inhibition of host cell translation elongation by Legionella pneumophila blocks the host cell unfolded protein response. Proc. Natl Acad. Sci. USA 112, E6790–E6797 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. De Leon, J. A. et al. Positive and negative regulation of the master metabolic regulator mTORC1 by two families of Legionella pneumophila effectors. Cell Rep. 21, 2031–2038 (2017).

    PubMed  PubMed Central  Google Scholar 

  72. Isaac, D. T. et al. Macrophage cell death and transcriptional response are actively triggered by the fungal virulence factor Cbp1 during H. capsulatum infection. Mol. Microbiol. 98, 910–929 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Zhang, K. Gctf: real-time CTF determination and correction. J. Struct. Biol. 193, 1–12 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  76. 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).

    CAS  PubMed  Google Scholar 

  77. Kallberg, M., Margaryan, G., Wang, S., Ma, J. & Xu, J. RaptorX server: a resource for template-based protein structure modeling. Methods Mol. Biol. 1137, 17–27 (2014).

    CAS  PubMed  Google Scholar 

  78. Winchell, K. R. et al. A structural, functional, and computational analysis of BshA, the first enzyme in the bacillithiol biosynthesis pathway. Biochemistry 55, 4654–4665 (2016).

    CAS  PubMed  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  82. Pettersen, E. F. et al. UCSF Chimera–a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    CAS  PubMed  Google Scholar 

  83. Han, C. et al. Polysome profiling followed by quantitative PCR for identifying potential micropeptide encoding long non-coding RNAs in suspension cell lines. STAR Protoc. 3, 101037 (2022).

    CAS  PubMed  Google Scholar 

  84. Li, B. & Dewey, C. N. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 12, 323 (2011).

    CAS  Google Scholar 

  85. Bray, N. L., Pimentel, H., Melsted, P. & Pachter, L. Near-optimal probabilistic RNA-seq quantification. Nat. Biotechnol. 34, 525–527 (2016).

    CAS  PubMed  Google Scholar 

  86. Ritchie, M. E. et al. limma powers differential expression analyses for RNA-sequencing and microarray studies. Nucleic Acids Res. 43, e47 (2015).

    PubMed  PubMed Central  Google Scholar 

  87. Smyth, G. K. Linear models and empirical bayes methods for assessing differential expression in microarray experiments. Stat. Appl. Genet. Mol. Biol. https://doi.org/10.2202/1544-6115.1027 (2004).

  88. Jolliffe, I. T. Principal Component Analysis 2nd edn (Springer, 2002).

  89. Van Der Walt, S., Colbert, S. C. & Varoquaux, G. The NumPy array: a structure for efficient numerical computation. Comput. Sci. Eng. 13, 22–30 (2011).

    Google Scholar 

  90. Reuter, J. S. & Mathews, D. H. RNAstructure: software for RNA secondary structure prediction and analysis. BMC Bioinform. 11, 129 (2010).

    Google Scholar 

  91. Guo, X. et al. Mitochondrial stress is relayed to the cytosol by an OMA1–DELE1–HRI pathway. Nature 579, 427–432 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  92. Ochoa, D. et al. An atlas of human kinase regulation. Mol. Syst. Biol. 12, 888 (2016).

    PubMed  PubMed Central  Google Scholar 

  93. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

  94. Davis, C. A. et al. The Encyclopedia of DNA elements (ENCODE): data portal update. Nucleic Acids Res. 46, D794–D801 (2018).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all of the members of the S.M. and P.W. laboratories for their critical evaluation and discussions of the data. We thank Z.-Q. Luo for his gift of isogenic ΔsidI and SidI-complemented L. pneumophila strains. We thank S. Venkatraman and L. Calviello from the Floor laboratory for helpful discussions. We thank B. Al-Sady and E. Simental from the Al-Sady laboratory for help with radioactivity experiments. We thank J. Yee from the Spitzer laboratory for help with flow cytometry experiments. We thank J. Noack, C. J. Sarabia and B. Wang for technical assistance with the project. L.W. acknowledges a fellowship from the Damon Runyon Cancer Research Foundation (DRG-2312-17). S.S. is supported by a Helen Hay Whitney Postdoc fellowship and K99 grant (grant no. 1K99GM143527-01A1) from NIGMS. N.J.K. acknowledges financial support from the National Institutes of Health (grant no. U19 AI135990). A.L.B. acknowledges financial support from the Dr. Miriam and Sheldon G. Adelson Medical Research Foundation. S.N.F. is a Pew Scholar in the Biomedical Sciences, supported by The Pew Charitable Trusts and acknowledges financial support from the National Institutes of Health (grant no. DP2GM132932). A. Sil acknowledges financial support from the National Institutes of Health (grant no. R01AI136735) and a gift fund from the Chan Zuckerberg Biohub. P.W. acknowledges financial support from the National Institutes of Health (grant no. R01GM032384) and the Howard Hughes Medical Institute. S.M. acknowledges financial support from the National Institutes of Health (grant nos R01GM140440 and R01GM144378), the Pew Charitable Trust (grant no. A129837), a Bowes Biomedical Investigator award and a gift fund from the Chan–Zuckerberg Biohub. We thank the Vincent J. Coates Genomics Sequencing Laboratory at the California Institute for Quantitative Biosciences (QB3) for help with the RNA-seq experiments. We also thank the QB3 shared cluster for computational support.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: A. Subramanian and S.M. Methodology: A. Subramanian, L.W., T.M., M.V., S.S., D.L.S., R.J.C., N.J.K., S.N.F., A. Sil, P.W. and S.M. Investigation: A. Subramanian, L.W., T.M., M.V., E.S., E.H.P., S.K., D.L.S., R.J.C. and K.H.L. Formal analysis: A. Subramanian, L.W., M.V., D.L.S. and R.J.C. Writing (original draft): A. Subramanian, L.W. and S.M. Writing (review and editing): A. Subramanian, L.W., M.V., A. Sil, P.W. and S.M. Visualization: A. Subramanian, L.W. and S.M. Funding acquisition: P.W. and S.M. Resources: S.S., A.L.B. and S.N.F. Supervision: A. Subramanian, D.L.S., N.J.K., A. Sil, P.W. and S.M.

Corresponding authors

Correspondence to Peter Walter or Shaeri Mukherjee.

Ethics declarations

Competing interests

P.W. is an inventor on US Patent 9708247 held by the Regents of the University of California that describes ISRIB and its analogues. Rights to the invention have been licensed by UCSF to Calico. The N.J.K. laboratory has received research support from Vir Biotechnology, F. Hoffmann-La Roche and Rezo Therapeutics. N.J.K. has financially compensated consulting agreements with the Icahn School of Medicine at Mount Sinai, New York, Maze Therapeutics, Interline Therapeutics, Rezo Therapeutics, GEn1E Lifesciences, Inc, and Twist Bioscience Corp. N.J.K. is on the Board of Directors of Rezo Therapeutics and is a shareholder in Tenaya Therapeutics, Maze Therapeutics, Rezo Therapeutics and Interline Therapeutics. D.L.S. has a consulting agreement with Maze Therapeutics and Rezo Therapeutics. All other authors declare no competing interests.

Peer review

Peer review information

Nature Cell Biology thanks the anonymous reviewers 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

Extended Data Fig. 1 Analysis of protein synthesis rates in cells after heterologous expression of L. pneumophila EITs and MS analyses of SidI-interacting partners.

ab, Analysis of protein synthesis rates in cells after heterologous expression of L. pneumophila EITs. a, Cells expressing epitope-tagged L. pneumophila EITs were subjected to puromycin pulse–chase analyses. Immunoblotting of lysates was performed using antibodies to puromycin and tubulin. Data are representative of two independent experiments. The values beneath the immunoblots represent densitometric quantification of puromycylated peptides normalized to the amount of tubulin in cells. b, The expression levels of L. pneumophila toxins were determined by qRT-PCR. Data represent the mean ± s.e.m. n, independent replicates; GFP (n = 25); lgt1, lgt2, lgt3, legK4 and ravX (n = 4); sidI and sidL (n = 6). Data were normalized to the average Ct values of the reference gene RPS29. c,d, MS analyses of SidI-interacting partners highlights the translation machinery as targets of SidI. c, Venn diagram depicting the common interacting partners enriched in GST–SidI-pulldown eluates from HEK293T cell lysates (n = 2) and RRLs (n = 1). n, independent replicates. d, List of translation-machinery components selectively enriched by GST–SidI. GST–SidI enrichment is calculated as a ratio of the peptide intensities observed in GST–SidI-pulldown eluates over GST-pulldown eluates or marked as peptide intensities uniquely measured only in GST–SidI-pulldown eluates. n.d., not detected.

Source data

Extended Data Fig. 2 Cryo-EM workflow and characterization of SidI structure.

a, Representative micrograph (from n = 5,620 micrograph images) showing the quality of data used for the final reconstruction of the SidI structure. Scale bar, 200 Å. b, Data processing scheme of the SidI structure. c, Fourier shell correlation (FSC) plots of the 3D reconstructions of SidI unmasked (grey) and masked (orange). d, Orientation angle distribution of the SidI reconstruction. e, EM maps of different regions of the SidI structure showing the quality of the data and the fit of the model. f, Local resolution map of the SidI structure.

Extended Data Fig. 3 Biochemical and biophysical characterization of SidI.

ac, GDP hydrolysis assay with 1 µM GST, GST–SidI and GST–Lgt2. Nucleotide sugars (50 µM) were provided as precursors. Graphs depict amount of nucleotide release. Data represent the mean ± s.e.m.; n = 3 independent replicates per treatment condition. d, Overlay of SidI (gold) and GDP–PimB′ (light grey) showing the structural similarity between the enzymatic pockets and the conserved amino acids. e, Different views of the SidI structure showing the three distinct kinks. f, Surface-charge distribution on the tRNA-mimicry domains of SidI, eEF2, EF-G, eRF1 and EF-P. The boxes highlight the regions in eEF2, EF-G, eRF1 and EF-P that contact the ribosome. g, Melting curve analyses of GST–SidI and GST–SidI mutants (n = 2 independent replicates).

Source data

Extended Data Fig. 4 Comparative structural and functional enzymatic analyses of SidI reveals a potential mode of ribosome binding.

a, Overlay of eEF2 and SidI structures showing their similarity in size and shape. The positively charged loops are boxed in black boxes. b, Structure of SidI overlayed on the structure of eEF2 in the eEF2-bound ribosome structure (PDB ID: 6GZ5). c, Zoomed-in view of the eEF2–ribosome complex showing the positively charged regions on eEF2 making contacts with: (1) eS31, (2) 18S rRNA and (3) tRNA at the P-site. d,e, MS analyses of Con-A-enriched proteins after in vitro mannosylation reactions with 1 µM GST–SidI or GST–SidI R453P co-incubated with ribosomes (240 nM) and GDP–mannose (50 µM). Graphs depict the average peptide counts of ribosomal proteins and SidI (n = 2 independent replicates). f,g, Con-A-enriched ribosomal proteins at/near the A-site/P-site and peptide exit tunnel of the ribosome (PDB ID: 6GZ5) are highlighted in green and blue, respectively.

Source data

Extended Data Fig. 5 SidI mutants precipitate eEF1A and tRNA ligases from HEK293T lysates.

a,b, Enrichment of interaction partners from HEK293T lysates with GST, GST–SidI and GST–SidI K-rich loop mutant (a) or GST–SidI R453P pulldowns (b). Immunoblotting was performed using antibodies to RTCB, VARS, CARS, eEF1A and GST. Data are representative of two independent experiments.

Source data

Extended Data Fig. 6 Analyses of RNA-seq datasets reveal a dynamic induction of stress transcripts during L. pneumophila infection in an effector-dependent manner.

a, Principal component analyses on all differentially expressed genes in HEK293-FcγR cells infected with WT and ΔdotA L. pneumophila for 1, 2, 4 or 8 h, and in HEK293T cells treated with ANS (0.1 µM) and DD-B (0.1 µM) for 4 h. Coloured circles depict replicates. b, Fold induction of upregulated and downregulated transcripts in common with transcripts induced by ribosome-stalling insults. Expression levels of transcripts are after WT and ΔdotA L. pneumophila infections normalized to uninfected controls; n = 2 independent replicates per time point.

Source data

Extended Data Fig. 7 Characterization of stress-transcript induction and ATF3 accumulation by L. pneumophila and inhibitors of ribosome function.

a, Heat map of absolute stress-transcript abundances (RPKM values) from L. pneumophila-infected macrophages extracted from RNA-seq datasets previously published by Lipo and colleagues34. b, Immunoblotting of ATF3 in lysates of Casp1/Casp11−/− primary BMDM cells infected with WT or ΔdotA L. pneumophila (MOI = 10), or treated with ANS (0.01 µM, 0.1 µM, 1 µM or 10 µM). The samples were derived from parallel experiments and the gels were processed in parallel. Data are representative of two independent experiments. c, Immunoblotting of ATF3 in the lysates of HEK293-FcγR cells infected with the indicated L. pneumophila strains. The values beneath the immunoblots depict densitometric quantification of ATF3 band intensities normalized to controls. Data are representative of two independent experiments. dh, Fold induction of ATF3, TNFRSF12A, HSPB8, PPP1R15A and GADD45A mRNA, measured by qRT-PCR, in HEK293T cells treated with vehicle control (n = 6), ANS 0.1 µM (n = 4), DD-B 0.1 µM (n = 4), DON 1 µM (n = 4), CHX 0.1 µM (n = 4) or HTN 0.1 µM (n = 4) for 4 h, or expressing FLAG–SidI (n = 6) or FLAG–SidI R453P (n = 3). Data represent the mean ± s.e.m. for n independent replicates. Transcript expression values were normalized internally to the reference gene RPS29 and expressed as the log2-transformed fold change over the levels in controls. i, Immunoblotting of ATF3 in the lysates of HEK293T cells treated with ANS (0.01 µM, 0.1 µM, 1 µM and 10 µM); DD-B (0.001 µM, 0.01 µM, 0.1 µM and 1 µM); CHX (0.01 µM, 0.1 µM, 1 µM and 10 µM); DON (0.01 µM, 0.1 µM, 1 µM and 10 µM) and HTN (0.01 µM, 0.1 µM, 1 µM and 10 µM) for 4 h. Data are representative of three independent experiments. Samples from the same experiment were processed in parallel on different gels. j, Heat map of the ATF3 band intensities measured by densitometric analyses of the blots in i. k, Indirect immunofluorescence of ATF3 in HeLa cells treated with ANS, DD-B, CHX, DON or HTN at the indicated concentrations. Scale bar, 10 µM. Data are representative of two independent experiments. The nuclear region is marked by arrows.

Source data

Extended Data Fig. 8 Characterization of translation rates, stress signalling and cell viability during L. pneumophila infections.

a, Immunoblotting of the lysates of cells infected with WT, ΔdotA and Δ5 L. pneumophila was performed using antibodies to puromycin. Ponceau S staining of membranes serves as a loading control. Data are representative of two independent experiments. b, Immunoblotting of lysates of cells treated with ANS (0.01 µM, 0.1 µM, 1 µM and 10 µM), DD-B (0.001 µM, 0.01 µM, 0.1 µM and 1 µM), CHX (0.01 µM, 0.1 µM, 1 µM and 10 µM), DON (0.01 µM, 0.1 µM, 1 µM and 10 µM) or HTN (0.01 µM, 0.1 µM, 1 µM and 10 µM) for 4 h was performed using antibodies to puromycin. Ponceau S staining of membranes serves as a loading control. Gels were processed in parallel. Data are representative of two independent experiments. c, Immunoblotting of ATF3, ATF4, phospho-eIF2 and GAPDH in the lysates of WT and eIF2 S51A+/+ MEF cells treated with ANS (0.01 µM, 0.1 µM, 1 µM and 10 µM) or thapsigargin (Tg; 1 µM) for 4 h. Data are representative of two independent experiments. d,e, Immunoblotting of ATF3 in the lysates of WT and eIF2 S51A+/+ MEF cells treated with DD-B (0.001 µM, 0.01 µM, 0.1 µM and 1 µM) for 4 h (d); and HEK293T cells treated with ANS (0.1 µM) or transfected with FLAG–SidI in the absence or presence of ISRIB (200 nM; e). Data are representative of two independent experiments. f, Immunoblotting of p-p38 and total p38 in lysates of HEK293T cells treated with ANS (0.01 µM, 0.1 µM, 1 µM and 10 µM), DD-B (0.001 µM, 0.01 µM, 0.1 µM and 1 µM), CHX (0.01 µM, 0.1 µM, 1 µM and 10 µM), DON (0.01 µM, 0.1 µM, 1 µM and 10 µM) or HTN (0.01 µM, 0.1 µM, 1 µM and 10 µM) for 4 h. Data are representative of two independent experiments. Samples from the same experiment were processed in parallel on different gels. g, Heat map of the p-p38 band intensities, measured by densitometric analyses, of the blots in f. h, Phosphoproteomics pipeline and inferred kinase activation in WT L. pneumophila-infected cells (MOI = 100) based on the analyses of phosphorylation sites on proteins. Kinases of the mitogen and stress activated protein kinase pathways are positively regulated and highlighted in green. i, Immunoblotting of p-p38 and total p38 in the lysates of HEK293-FcγR cells infected with WT, Δ5 or Δ5+sidI L. pneumophila strains (MOI = 50). The values beneath the immunoblots represent densitometric quantification of p-p38 levels normalized to the total p38 levels. Data are representative of two independent experiments. j, Cell viability measurements of HEK293-FcγR cells infected with WT L. pneumophila (MOI = 50; U.I., 8 h and 10 h samples (n = 4); 1 h, 2 h and 4 h samples (n = 2); n, independent replicates per time point. For samples with n ≥ 3, data represent the mean ± s.e.m. P values were calculated using a Student’s t-test.

Source data

Extended Data Fig. 9 ATF3 binds to genomic regions of stress-inducible transcripts.

Mapping and enrichment of ATF3-binding zones on the ATF3, TNFRSF12A, HSPB8, PPP1R15A and GADD45A genes across four cell types. H3K27Ac tracks mark acetylation of lysine 27 of the H3 histone protein across 7 cell types (GM12878, H1-hESC, HSMM, HSMM, K562, NHEK and NHLF cells). Tracks of each cell type are marked with a different colour and displayed as a transparent overlay. The enrichment scores for ATF3 peaks were extracted from the ENCODE project datasets and visualized using the UCSC genome browser (human hg38 assembly). The display for this track shows the site location with the point-source of the peak marked with a coloured vertical bar and the level of enrichment at the site indicated by the darkness of the item. The enrichment values were computed based on signal values assigned by the ENCODE pipeline. The input signal values were multiplied by a normalization factor calculated as the ratio of the maximum score value (1,000) to the signal value at 1 s.d. from the mean, with values exceeding 1,000 capped at 1,000.

Extended Data Fig. 10 ATF3 depletion attenuates stress-transcript induction.

ad, Fold-change induction of TNFRSF12A, HSPB8, PPP1R15A and GADD45A mRNA in parent and ATF3-KO HEK293T cells expressing GFP, FLAG–SidI or FLAG–SidI R453P, or treated with ANS (0.1 µM) measured by qRT-PCR (n = 2 independent replicates). Transcript expression values were normalized internally to the reference gene RPS29 and expressed as fold change over the levels in GFP-expressing cells.

Source data

Supplementary information

Reporting Summary

Peer Review File

Supplementary Table 1

MS analyses of SidI-interacting partners in HEK293T lysates and RRLs.

Supplementary Tables 2, 7, 8 and 9

Supplementary Table 2. Cryo-EM data collection, reconstruction and model refinement statistics. Supplementary Table 7. List of siRNAs and oligonucleotides. Supplementary Table 8. Amino-acid sequences of Legionella effector proteins used in this study. Supplementary Table 9. Key resources used in this study.

Supplementary Table 3

MS analyses of Con-A pulldowns after in vitro mannosylation reactions with SidI and ribosomes.

Supplementary Table 4

Fold changes in transcript expression levels in cells infected with L. pneumophila or treated with the ribosome-stalling inducers ANS and DD-B. Differentially expressed genes were identified by comparing replicate means for contrasts of interest using LIMMA version 3.30.8. Genes were considered significantly differentially expressed if they were statistically significant (at 5% FDR) with an effect size of at least 2× (absolute log2(fold change) ≥ 1) for a given contrast. LIMMA implements a two-sided moderated t-test with empirical Bayesian shrinkage.

Supplementary Table 5

Gene-set-enrichment analyses on the ranked list of genes from the RNA-seq datasets.

Supplementary Table 6

Inferred kinase-activity scores for cells infected with L. pneumophila for 1 h by reanalysis of phosphoproteomics datasets in the study by Noack and colleagues43.

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

Subramanian, A., Wang, L., Moss, T. et al. A Legionella toxin exhibits tRNA mimicry and glycosyl transferase activity to target the translation machinery and trigger a ribotoxic stress response. Nat Cell Biol 25, 1600–1615 (2023). https://doi.org/10.1038/s41556-023-01248-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-023-01248-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