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:

Guanylate-binding proteins convert cytosolic bacteria into caspase-4 signaling platforms

Abstract

Bacterial lipopolysaccharide triggers human caspase-4 (murine caspase-11) to cleave gasdermin-D and induce pyroptotic cell death. How lipopolysaccharide sequestered in the membranes of cytosol-invading bacteria activates caspases remains unknown. Here we show that in interferon-γ-stimulated cells guanylate-binding proteins (GBPs) assemble on the surface of Gram-negative bacteria into polyvalent signaling platforms required for activation of caspase-4. Caspase-4 activation is hierarchically controlled by GBPs; GBP1 initiates platform assembly, GBP2 and GBP4 control caspase-4 recruitment, and GBP3 governs caspase-4 activation. In response to cytosol-invading bacteria, activation of caspase-4 through the GBP platform is essential to induce gasdermin-D-dependent pyroptosis and processing of interleukin-18, thereby destroying the replicative niche for intracellular bacteria and alerting neighboring cells, respectively. Caspase-11 and GBPs epistatically protect mice against lethal bacterial challenge. Multiple antagonists of the pathway encoded by Shigella flexneri, a cytosol-adapted bacterium, provide compelling evolutionary evidence for the importance of the GBP–caspase-4 pathway in antibacterial defense.

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: IFN-γ prevents proliferation of cytosol-invading S. Typhimurium.
Fig. 2: Cytosol-invading bacteria recruit CASP4 for GSDMD-dependent pyroptosis.
Fig. 3: GBPs recruit and activate CASP4 at the surface of cytosol-invading bacteria.
Fig. 4: Formation of LPS-dependent GBP–CASP4 complexes.
Fig. 5: Analysis of LPS-dependent GBP–CASP4 complexes.
Fig. 6: Structure–function analysis of the GBP-dependent CASP4 signaling platform.
Fig. 7: Gbps and Casp11 protect mice epistatically against bacterial infection.
Fig. 8: Processing and secretion of IL-18 require GBP-dependent CASP4 activity.

Similar content being viewed by others

Data availability

All data generated or analyzed during this study are included in this published article (and its supplementary information files). The source data that support the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Poltorak, A. et al. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282, 2085–2088 (1998).

    PubMed  CAS  Google Scholar 

  2. Takeuchi, O. & Akira, S. Pattern recognition receptors and inflammation. Cell 140, 805–820 (2010).

    PubMed  CAS  Google Scholar 

  3. Park, B. S. et al. The structural basis of lipopolysaccharide recognition by the TLR4–MD-2 complex. Nature 458, 1191–1195 (2009).

    PubMed  CAS  Google Scholar 

  4. Hagar, J. A., Powell, D. A., Aachoui, Y., Ernst, R. K. & Miao, E. A. Cytoplasmic LPS activates caspase-11: implications in TLR4-independent endotoxic shock. Science 341, 1250–1253 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  5. Shi, J. et al. Inflammatory caspases are innate immune receptors for intracellular LPS. Nature 514, 187–192 (2014).

    PubMed  CAS  Google Scholar 

  6. Kayagaki, N. et al. Noncanonical inflammasome activation by intracellular LPS independent of TLR4. Science 341, 1246–1249 (2013).

    PubMed  CAS  Google Scholar 

  7. Aachoui, Y. et al. Caspase-11 protects against bacteria that escape the vacuole. Science 339, 975–978 (2013).

    PubMed  PubMed Central  CAS  Google Scholar 

  8. Kayagaki, N. et al. Caspase-11 cleaves gasdermin D for non-canonical inflammasome signalling. Nature 526, 666–671 (2015).

    PubMed  CAS  Google Scholar 

  9. Kayagaki, N. et al. Non-canonical inflammasome activation targets caspase-11. Nature 479, 117–121 (2011).

    PubMed  CAS  Google Scholar 

  10. Shi, J. et al. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature 526, 660–665 (2015).

    PubMed  CAS  Google Scholar 

  11. Broz, P. & Dixit, V. M. Inflammasomes: mechanism of assembly, regulation and signalling. Nat. Rev. Immunol. 16, 407–420 (2016).

    PubMed  CAS  Google Scholar 

  12. Shi, J. et al. Inflammatory caspases are innate immune receptors for intracellular LPS. Nature 514, 187–192 (2014).

    PubMed  CAS  Google Scholar 

  13. Kim, B. H. et al. A family of IFN-inducible 65-kD GTPases protects against bacterial infection. Science 332, 717–721 (2011).

    PubMed  CAS  Google Scholar 

  14. Martens, S. & Howard, J. The interferon-inducible GTPases. Annu. Rev. Cell Dev. Biol. 22, 559–589 (2006).

    PubMed  CAS  Google Scholar 

  15. Tretina, K., Park, E.-S., Maminska, A. & MacMicking, J. D. Interferon-induced guanylate-binding proteins: guardians of host defense in health and disease. J. Exp. Med. 216, 482–500 (2019).

    PubMed  PubMed Central  CAS  Google Scholar 

  16. Pilla, D. M. et al. Guanylate binding proteins promote caspase-11-dependent pyroptosis in response to cytoplasmic LPS. Proc. Natl Acad. Sci. USA 111, 6046–6051 (2014).

    PubMed  CAS  Google Scholar 

  17. Santos, J. C. et al. LPS targets host guanylate‐binding proteins to the bacterial outer membrane for non‐canonical inflammasome activation. EMBO J. 37, e98089 (2018).

    PubMed  PubMed Central  Google Scholar 

  18. Meunier, E. et al. Caspase-11 activation requires lysis of pathogen-containing vacuoles by IFN-induced GTPases. Nature 509, 366–370 (2014).

    PubMed  CAS  Google Scholar 

  19. Fisch, D. et al. Human GBP1 is a microbe‐specific gatekeeper of macrophage apoptosis and pyroptosis. EMBO J. 38, e100926 (2019).

    PubMed  PubMed Central  Google Scholar 

  20. Man, S. M. et al. IRGB10 liberates bacterial ligands for sensing by the AIM2 and caspase-11-NLRP3 inflammasomes. Cell 167, 382–396.e17 (2016).

    PubMed  PubMed Central  CAS  Google Scholar 

  21. Bekpen, C. et al. The interferon-inducible p47 (IRG) GTPases in vertebrates: loss of the cell autonomous resistance mechanism in the human lineage. Genome Biol. 6, R92 (2005).

    PubMed  PubMed Central  Google Scholar 

  22. Mitchell, G. & Isberg, R. R. Innate immunity to intracellular pathogens: balancing microbial elimination and inflammation. Cell Host Microbe 22, 166–175 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  23. Matsuzawa-Ishimoto, Y., Hwang, S. & Cadwell, K. Autophagy and inflammation. Annu. Rev. Immunol. 36, 73–101 (2018).

    PubMed  CAS  Google Scholar 

  24. Thurston, T. L. M., Wandel, M. P., von Muhlinen, N., Foeglein, A. & Randow, F. Galectin 8 targets damaged vesicles for autophagy to defend cells against bacterial invasion. Nature 482, 414–418 (2012).

    PubMed  PubMed Central  CAS  Google Scholar 

  25. Thurston, T. L. M., Ryzhakov, G., Bloor, S., von Muhlinen, N. & Randow, F. The TBK1 adaptor and autophagy receptor NDP52 restricts the proliferation of ubiquitin-coated bacteria. Nat. Immunol. 10, 1215–1221 (2009).

    PubMed  CAS  Google Scholar 

  26. Thurston, T. L. et al. Recruitment of TBK1 to cytosol‐invading Salmonella induces WIPI2‐dependent antibacterial autophagy. EMBO J. 35, e201694491 (2016).

    Google Scholar 

  27. Ravenhill, B. J. et al. The cargo receptor NDP52 initiates selective autophagy by recruiting the ULK complex to cytosol-invading bacteria. Mol. Cell 74, 320–329.e6 (2019).

    PubMed  PubMed Central  CAS  Google Scholar 

  28. Radtke, A. L., Delbridge, L. M., Balachandran, S., Barber, G. N. & O’Riordan, M. X. D. TBK1 protects vacuolar integrity during intracellular bacterial infection. PLoS Pathog. 3, e29 (2007).

    PubMed  PubMed Central  Google Scholar 

  29. Isberg, R. R. & Falkow, S. A single genetic locus encoded by Yersinia pseudotuberculosis permits invasion of cultured animal cells by Escherichia coli K-12. Nature 317, 262–264 (1985).

    PubMed  CAS  Google Scholar 

  30. Isberg, R. R., Voorhis, D. L. & Falkow, S. Identification of invasin: a protein that allows enteric bacteria to penetrate cultured mammalian cells. Cell 50, 769–778 (1987).

    PubMed  CAS  Google Scholar 

  31. Perrin, A., Jiang, X., Birmingham, C., So, N. & Brumell, J. Recognition of bacteria in the cytosol of mammalian cells by the ubiquitin system. Curr. Biol. 14, 806–811 (2004).

    PubMed  CAS  Google Scholar 

  32. von Muhlinen, N. et al. LC3C, bound selectively by a noncanonical LIR motif in NDP52, is required for antibacterial autophagy. Mol. Cell 48, 329–342 (2012).

    PubMed  PubMed Central  CAS  Google Scholar 

  33. Kobayashi, T. et al. The Shigella OspC3 effector inhibits caspase-4, antagonizes inflammatory cell death, and promotes epithelial infection. Cell Host Microbe 13, 570–583 (2013).

    PubMed  CAS  Google Scholar 

  34. Briken, V. et al. Interferon regulatory factor 1 is required for mouse Gbp gene activation by gamma interferon. Mol. Cell. Biol. 15, 975–982 (1995).

    PubMed  PubMed Central  CAS  Google Scholar 

  35. Li, P. et al. Ubiquitination and degradation of GBPs by a Shigella effector to suppress host defence. Nature 551, 378–383 (2017).

    PubMed  CAS  Google Scholar 

  36. Wandel, M. P. et al. GBPs inhibit motility of Shigella flexneri but are targeted for degradation by the bacterial ubiquitin ligase IpaH9.8. Cell Host Microbe 22, 507–518.e5 (2017).

    PubMed  PubMed Central  CAS  Google Scholar 

  37. Piro, A. S. et al. Detection of cytosolic Shigella flexneri via a C-terminal triple-arginine motif of GBP1 inhibits actin-based motility. Mbio 8, e01979-17 (2017).

    PubMed  PubMed Central  Google Scholar 

  38. Kane, C. D., Schuch, R., Day, W. A. & Maurelli, A. T. MxiE regulates intracellular expression of factors secreted by the Shigella flexneri 2a type III secretion system. J. Bacteriol. 184, 4409–4419 (2002).

    PubMed  PubMed Central  CAS  Google Scholar 

  39. Ghosh, A., Praefcke, G. J. K., Renault, L., Wittinghofer, A. & Herrmann, C. How guanylate-binding proteins achieve assembly-stimulated processive cleavage of GTP to GMP. Nature 440, 101–104 (2006).

    PubMed  CAS  Google Scholar 

  40. Prakash, B., Praefcke, G. J., Renault, L., Wittinghofer, A. & Herrmann, C. Structure of human guanylate-binding protein 1 representing a unique class of GTP-binding proteins. Nature 403, 567–571 (2000).

    PubMed  CAS  Google Scholar 

  41. Knodler, L. A. et al. Noncanonical inflammasome activation of caspase-4/caspase-11 mediates epithelial defenses against enteric bacterial pathogens. Cell Host Microbe 16, 249–256 (2014).

    PubMed  PubMed Central  CAS  Google Scholar 

  42. Ramirez, M. L. G. et al. Extensive peptide and natural protein substrate screens reveal that mouse caspase-11 has much narrower substrate specificity than caspase-1. J. Biol. Chem. 293, 7058–7067 (2018).

    PubMed  PubMed Central  Google Scholar 

  43. Jorgensen, I., Zhang, Y., Krantz, B. A. & Miao, E. A. Pyroptosis triggers pore-induced intracellular traps (PITs) that capture bacteria and lead to their clearance by efferocytosis. J. Exp. Med. 213, 2113–2128 (2016).

    PubMed  PubMed Central  CAS  Google Scholar 

  44. Miao, E. A. et al. Caspase-1-induced pyroptosis is an innate immune effector mechanism against intracellular bacteria. Nat. Immunol. 11, 1136–1142 (2010).

    PubMed  PubMed Central  CAS  Google Scholar 

  45. Randow, F. & Sale, J. E. Retroviral transduction of DT40. Subcell. Biochem. 40, 383–386 (2006).

    PubMed  Google Scholar 

  46. Nowarski, R. et al. Epithelial IL-18 equilibrium controls barrier function in colitis. Cell 163, 1444–1456 (2015).

    PubMed  PubMed Central  CAS  Google Scholar 

  47. Kraiczy, J. et al. DNA methylation defines regional identity of human intestinal epithelial organoids and undergoes dynamic changes during development. Gut 68, 49–61 (2019).

    PubMed  CAS  Google Scholar 

  48. Ettayebi, K. et al. Replication of human noroviruses in stem cell-derived human enteroids. Science 353, 1387–1393 (2016).

    PubMed  PubMed Central  Google Scholar 

  49. Koestler, B. J., Ward, C. M. & Payne, S. M. Shigella pathogenesis modeling with tissue culture assays. Curr. Protoc. Microbiol. 50, e57 (2018).

    PubMed  PubMed Central  Google Scholar 

  50. Brinkman, E. K., Chen, T., Amendola, M. & van Steensel, B. Easy quantitative assessment of genome editing by sequence trace decomposition. Nucleic Acids Res. 42, e168 (2014).

    PubMed  PubMed Central  Google Scholar 

  51. Sidik, S. et al. A Shigella flexneri virulence plasmid encoded factor controls production of outer membrane vesicles. G3 (Bethesda) 4, 2493–2503 (2014).

    PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank J. Kendrick-Jones (MRC Laboratory of Molecular Biology, Cambridge) for providing antiserum against NDP52. This work was supported by the MRC (grant no. U105170648) and the Wellcome Trust (grant no. WT104752MA) to F.R., and by the NIH National Institutes of Allergy and Infectious Diseases (grant nos. R01AI068041-13 and R01AI108834-05) to J.D.M. J.D.M. is an investigator of the Howard Hughes Medical Institute.

Author information

Authors and Affiliations

Authors

Contributions

M.P.W. performed and analyzed all experiments with the following exceptions: K.B.B. performed and analyzed experiments in enterocytes and bacterial proliferation assays in knockout cells; B.-H.K., E.-S.P. and J.D.M. designed, performed and analyzed Salmonella and Shigella infections in mice; A.H. and J.R. generated Shigella mutants; K.N. and M.Z. generated human enteroids; and B.L. and T.H. generated U937 knockouts. M.P.W. and F.R. designed the study and wrote the manuscript.

Corresponding authors

Correspondence to Michal P. Wandel or Felix Randow.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Zoltan Fehervari was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 IFN-γ prevents proliferation of cytosol-invading S. Typhimurium.

a, Lysates of HeLa cells treated with the indicated siRNAs. Blots were probed with the indicated antibodies, PCNA – loading control. b, Colony-forming units (CFU) of S. Typhimurium in HeLa cells at 1h p.i.. c, Fold replication of S. Typhimurium ΔprgH +inv in HeLa cells. d, Confocal micrographs of HeLa cells infected with S. Typhimurium taken at 1 h p.i. stained with DAPI and antibodies against Galectin-8 and ubiquitin (FK2 antibody) (top panel) or over-expressing GFP::LC3C and stained with DAPI and antibody against NDP52 (bottom panel). e, Percentage of Annexin V positive Hela cells expressing CFP::Galectin-8 amongst cells harbouring intracellular S. Typhimurium. Negative or positive – none or at least one bacterium per cell positive for CFP::Galectin-8. Live imaged every 6 min for 6 h, 12 fields per condition. f, Percentage of PI positive nuclei in HeLa cells infected with S. Typhimurium at 2h p.i.. Cells were treated with DMSO, 50 μM NEC-1s, 10 μM NSA or 50 μM Z-VAD-FMK as indicated. g, Confocal micrograph of HeLa cells infected with S. Typhimurium in the presence of FAM-VAD-FMK and stained with DAPI and antibody against Galectin-8. Image taken at 90 min p.i.. Statistical significance was assessed by two-tailed unpaired Student’s t-test (b), one-way (e,f) or two-way (c) analysis of variance (ANOVA) with Tukey’s multiple comparisons test; ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (c, e, f) or five (b) independent experiments, or representative of two (a) or three (d, g) independent experiments. HeLa cells were treated with IFN-γ (g) or treated with IFN-γ as indicated (a-c, e, f). Bacteria were counted based on their ability to grow on agar plates (b, c). Scale bar, 10 μm (d, g). Uncropped blots (a) are shown in the Source Data. PI - propidium iodide, p.i. - post-infection, S.T. - S. Typhimurium.

Source data

Extended Data Fig. 2 Cytosol-invading bacteria recruit caspase-4.

a, Confocal micrographs of HeLa cells over-expressing GFP::Caspase-4 or -5 at 1 h p.i. with S. Typhimurium and stained with DAPI. Scale bar, 10 μm. b, f, g, Lysates of HeLa cells expressing the indicated GFP::Caspase constructs (b), of cells treated with the indicated siRNAs (f), or of the indicated control or knock-out cells (g). Blots were probed with indicated antibodies, PCNA (b, f), Actin (g) – loading control. Samples in Extended Data Fig. 2g, Fig. 2p and Fig. 8c were obtained from the same experiment. c, d, Percentage of S. Typhimurium positive for the indicated GFP::Caspase constructs (c) or staining positive for endogenous Galectin-8 and/or Caspase-4 (d) in HeLa cells at 1 h p.i.. n > 100 bacteria per coverslip, in triplicate. e, Percentage of FAM-VAD-FMK positive S. Typhimurium amongst bacteria staining positive for endogenous Galectin-8 at 90 min p.i. in HeLa cells treated with siRNAs against caspases as indicated. n > 100 Galectin-8 +ve bacteria per coverslip, in triplicate. h, Percentage of PI positive nuclei in the indicated control or knock-out HeLa cells uninfected or infected with S. Typhimurium at 2h p.i.. i, Fold replication of S. Typhimurium in HeLa cells treated with the indicated siRNAs against caspases. Bacteria were counted based on their ability to grow on agar plates. Statistical significance was assessed by one-way (e, h) or two-way (i) analysis of variance (ANOVA) with Tukey’s multiple comparisons test; ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (c, d, e, i) independent experiments, or representative of two (b, f, g) or three (a) independent experiments. HeLa cells were treated with IFN-γ (c) or treated with IFN-γ as indicated (a, d-i). Uncropped blots (b, f, g) are shown in the Source Data. PI - propidium iodide, p.i. - post-infection, +ve – positive.

Source data

Extended Data Fig. 3 The S. flexneri effector OspC3 inhibits interferon-induced pyroptosis.

a, PI positive nuclei in HeLa cells infected with the indicated S. flexneri strains at 2h p.i.. b, Percentage of PI positive nuclei in CFP::Galectin-8 expressing HeLa cells infected with the indicated S. flexneri strains in the presence of PI. n = three (WT) and four (ΔospC3) independent repeats. Live imaged every 5 min for 5 h, 10 fields per condition. c, Percentage of Zombie Green positive (that is dead) cells in monolayers of differentiated human epithelial organoids at 2 h p.i. with the indicated S. flexneri strains. n > 50 infected cells per coverslip. d, Confocal micrograph of a monolayer of differentiated human epithelial organoids stained with Zombie Green and antibody against ZO-1 at 1 h p.i. with the indicated S. flexneri strains. Scale bar, 10 μm. e, f, Fold replication of the indicated S. flexneri strains in HeLa cells. Bacteria were counted based on their ability to grow on agar plates. Statistical significance was assessed by one-way (a, b) or two-way (e, f) analysis of variance (ANOVA) with Tukey’s multiple comparisons test; ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (a, b, e, f) or four (b) independent experiments, or representative of two (d) independent experiments. HeLa cells were treated with IFN-γ as indicated (a-f). PI - propidium iodide, p.i. - post-infection.

Extended Data Fig. 4 Cytosol-invading bacteria trigger caspase-4 and gasdermin-D dependent pyroptosis.

a, b, g, Fold replication of S. flexneri ΔospC3 in HeLa cells treated with the indicated siRNAs against caspases (a) and in control or knock-out HeLa cells (b, g). Bacteria were counted based on their ability to grow on agar plates. c, d, Lysates of HeLa cells at 1 h p.i. with the indicated S. flexneri strains. Pro - full length pro-form of GSDMD (shorter exposure), NT - N-terminal domain of GSDMD (longer exposure). Samples in Extended Data Figs. 4c,d and 9a, c were obtained from the same experiment. e, h, Lysates of HeLa cells treated with the indicated siRNAs. * unspecific band. f, PI positive nuclei in the indicated control or knock-out HeLa cells at 2 h p.i. with S. flexneri ΔospC3. Statistical significance was assessed by one-way (f) or two-way (a, b, g) analysis of variance (ANOVA) with Tukey’s multiple comparisons test; ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (b, g, f) or four (a) independent experiments, or representative of two (e, h) or three (c, d) independent experiments. HeLa cells were treated with IFN-γ as indicated (a-h). Blots were probed with indicated antibodies, PCNA (e, h), Actin (c, d) – loading control. Uncropped blots (c-e, h) are shown in the Source Data. PI - propidium iodide, p.i. - post-infection.

Source data

Extended Data Fig. 5 GBP1 recruits GBP2-4 to S. Typhimurium.

a, b, Lysates of HeLa cells treated with the indicated siRNAs (a), or from the indicated control or knock-out cells (b). Blots were probed with indicated antibodies, PCNA (a), Actin (b) – loading control. Samples in Extended Data Fig. 5b, 8c and 9f were obtained from the same experiment. c, Percentage of S. Typhimurium positive for the indicated GFP::GBP constructs at 1 h p.i. in the indicated control or knock-out HeLa cells. n > 100 bacteria per coverslip, in triplicate. d, Structured illumination micrograph of HeLa cells expressing GFP::GBP1 and antibody-stained for Galectin-8 at 1 h p.i. with S. Typhimurium. Scale bar, 1 μm. Statistical significance was assessed by one-way analysis of variance (ANOVA) with Tukey’s multiple comparisons test (c); ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (c) independent experiments, or representative of two (a, b) or three (d) independent experiments. HeLa cells were treated with IFN-γ (c, d) or treated with IFN-γ as indicated (b). Uncropped blots (a, b) are shown in the Source Data. p.i. - post-infection.

Source data

Extended Data Fig. 6 GBPs target cytosol-invading S. Typhimurium.

Confocal micrographs of HeLa cells over-expressing GFP::GBP1-7 and stained with DAPI and antibody against NDP52 at 1 h p.i. with S. Typhimurium. Representative of three independent experiments. Scale bar, 10 μm.

Extended Data Fig. 7 GBPs recruit and activate caspase-4.

a, Percentage of S. Typhimurium positive for GFP::GBP1-4 and/or staining positive for endogenous Caspase-4 in HeLa cells at 1 h p.i.. n > 100 bacteria per coverslip, in triplicate. b, e, Percentage of endogenous Caspase-4 (b) or FAM-VAD-FMK (e) positive S. Typhimurium amongst bacteria staining positive for endogenous Galectin-8 in the indicated control or knock-out HeLa cells at 1h (b) or 90 min (e) p.i.. n > 100 Galectin-8 +ve bacteria per coverslip, in triplicate. c, f Percentage of endogenous Caspase-4 (c) or FAM-VAD-FMK (f) positive bacteria of the indicated S. flexneri strains in HeLa cells at 1 h p.i.. n > 100 bacteria per coverslip, in triplicate. d, Confocal micrograph of a monolayer of differentiated human epithelial organoids antibody-stained for GBP1 and Caspase-4 at 1 h p.i. with S. flexneri ΔipaH9.8. g, Confocal micrographs of HeLa cells treated with DMSO or Carfilzomib as indicated at 1h p.i. with S. flexneri ΔospC3 in the presence of FAM-VAD-FMK and stained with DAPI. h-j, Percentage of FAM-VAD-FMK positive S. flexneri ΔospC3 ΔipaH9.8 (h, j) or S. flexneri ΔospC3 (i) at 1 h p.i. in the indicated control or knock-out HeLa cells (h,j) or in HeLa cells treated with the indicated siRNAs against caspases and 1 μM Carfilzomib (i). n > 100 bacteria per coverslip, in triplicate. Statistical significance was assessed by one-way analysis of variance (ANOVA) with Tukey’s multiple comparisons test (b, c, e, f, h-j); ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (a-c, e, f, h-j) independent experiments, or representative of two (d, g) independent experiments. HeLa cells were treated with IFN-γ (a, b, d, e, g, h) or treated with IFN-γ as indicated (c, f, i, j). Cells were treated with DMSO or 1 μM Carfilzomib as indicated (c,f, g). Scale bar, 10 μm (d, g). p.i. - post-infection, +ve – positive.

Extended Data Fig. 8 GBPs govern gasdermin-D dependent pyroptosis.

a, b, Percentage of FAM-VAD-FMK positive cells among HeLa cells containing S. Typhimurium positive for endogenous Galectin-8 at 90 min p.i. (a) or containing S. flexneri ΔospC3 at 1 h p.i. (b); cells treated with siRNAs against GBPs as indicated. n > 100 cells with Galectin-8 +ve bacteria (a) or n > 100 infected cells (b) per coverslip, in triplicate. c, Lysates of the indicated control or knock-out HeLa cells infected with S. flexneri ΔospC3 for 1h. Pro - full length pro-form of GSDMD (shorter exposure), NT - N-terminal domain of GSDMD (longer exposure). Samples in Extended Data Fig. 8c, Extended Data Fig. 5b and Extended Data Fig. 9f were obtained from the same experiment. d, e, Percentage of PI positive nuclei in the indicated control or knock-out HeLa cells at 2 h p.i. with S. Typhimurium (e) or S. flexneri ΔospC3 (d). f, Lysates of the indicated control or knock-out U937 cells. g, Sanger sequencing chromatogram of control and GBP3 knock-out U937 cells. Statistical significance was assessed by one-way analysis of variance (ANOVA) with Tukey’s multiple comparisons test (a, b, d, e); ns, not significant, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (b, d, e) or four (a) independent experiments, or representative of two (c, f) independent experiments. HeLa cells were treated with IFN-γ as indicated (a-f). Blots were probed with indicated antibodies, Actin – loading control (c, f). Uncropped blots (c, f) are shown in the Source Data. PI - propidium iodide, p.i. - post-infection, +ve – positive. S. T - S. Typhimurium.

Source data

Extended Data Fig. 9 Processing and secretion of IL-18 during S. flexneri infection.

a, c, Lysates of HeLa cells prepared at 1 h p.i. with the indicated S. flexneri strains. Samples in Extended Data Fig. 9a, c and Extended Data Fig. 4c, d were obtained from the same experiment. b, d, Release of IL-18 from HeLa cells infected with the indicated S. flexneri strains for 1h. e, Lysates of HeLa cells expressing the indicated FLAG-tagged caspase alleles and treated with the indicated siRNAs prepared at 1 h p.i. with S. flexneri ΔospC3. f, Lysates of the indicated control or knock-out HeLa cells prepared at 1 h p.i. with S. flexneri ΔospC3. Samples in Extended Data Fig. 9f, Extended Data Fig. 5b and Extended Data Fig. 8c and were obtained from the same experiment. g, Release of IL-18 from the indicated control or knock-out HeLa cells infected with S. flexneri ΔospC3 for 1h. Statistical significance was assessed by one-way analysis of variance (ANOVA) with Tukey’s multiple comparisons test (b, d, g); ns, not significant, *P < 0.05, **P < 0.01 (exact p values are provided in Supplementary Table 1). Data are expressed as the Mean ± SEM of three (b, d, g) independent experiments, or representative of two (e, f) or three (a, c) independent experiments. HeLa cells were treated with IFN-γ as indicated (a-g). Blots were probed with indicated antibodies, Actin – loading control (a, c, e, f). Uncropped blots (a, c, e, f) are shown in the Source Data.

Source Data

Extended Data Fig. 10 Schematic illustration of the GBP-CASP4 pathway.

Interferon-induced guanylate-binding proteins (GBPs) transform Gram-negative bacteria into a caspase activation platform by coating their surface with a polyvalent protein array. The bacterial GBP coat may serve to foster contacts between CASP4 and its microbial ligand, the hydrophobic lipid A moiety of LPS, an integral and otherwise inaccessible component of the bacterial outer membrane. GBPs control CASP4 activation in a hierarchical manner; GBP1 initiates platform assembly, GBP2 and GBP4 control CASP4 recruitment, whereas GBP3 governs CASP4 activation. Once activated CASP4 cleaves GSDMD and IL-18 to cause pyroptotic cell death and cytokine release, thereby destroying the bacterial niche and alerting neighbouring cells of imminent danger. The cytosol-adapted bacterium Shigella flexneri antagonizes the pathway through secretion of the CASP4 inhibitor OspC3.

Supplementary information

Reporting Summary

Supplementary Table 1

Table containing precise P values.

Supplementary Video 1

Success and failure of galectin-8 to restrict proliferation of S. Typhimurium. Comparison of the outcome of cytosolic entry, marked by recruitment of galectin-8, for events 1 and 2. Live imaging on a confocal spinning disk microscope of HeLa cells expressing CFP::galectin-8 (white), infected with GFP-expressing S. Typhimurium (green) and live imaged every 6 min. Representative of six independent experiments. Scale bar, 10 μm.

Supplementary Video 2

Cytosolic entry of S. Typhimurium kills IFN-γ-treated cells. Illustration of the outcome of cytosolic entry, marked by recruitment of galectin-8, for two events. Live imaging on a confocal spinning disk microscope of IFN-γ-treated HeLa cells expressing CFP::galectin-8 (white), infected with GFP-expressing S. Typhimurium (green) in the presence of propidium iodide (red) and live imaged every 5 min. Representative of six independent experiments. Scale bar, 10 μm.

Supplementary Video 3

Caspase-4 recruitment to S. Typhimurium precedes cell death. Illustration of sequential events: (1) entry into the cytosol, marked by galectin-8; (2) recruitment of caspase-4; and (3) cell death, marked by nuclear accumulation of propidium iodide. Live imaging on a confocal spinning disk microscope of IFN-γ-treated HeLa cells expressing CFP::galectin-8 (white) and GFP::caspase-4 N104 (green), infected with mCherry-expressing S. Typhimurium (red) in the presence of propidium iodide (red) and live imaged every 5 min. Representative of three independent experiments. Scale bar, 10 μm.

Source data

Source Data Fig. 2

Unprocessed western blots

Source Data Fig. 3

Unprocessed western blots

Source Data Fig. 4

Unprocessed western blots

Source Data Fig. 5

Unprocessed western blots

Source Data Fig. 6

Unprocessed western blots

Source Data Fig. 8

Unprocessed western blots

Source Data Extended Data Fig. 1

Unprocessed western blots

Source Data Extended Data Fig. 2

Unprocessed western blots

Source Data Extended Data Fig. 4

Unprocessed western blots

Source Data Extended Data Fig. 5

Unprocessed western blots

Source Data Extended Data Fig. 8

Unprocessed western blots

Source Data Extended Data Fig. 9

Unprocessed western blots

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wandel, M.P., Kim, BH., Park, ES. et al. Guanylate-binding proteins convert cytosolic bacteria into caspase-4 signaling platforms. Nat Immunol 21, 880–891 (2020). https://doi.org/10.1038/s41590-020-0697-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41590-020-0697-2

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