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:

The tuberculosis necrotizing toxin kills macrophages by hydrolyzing NAD

Abstract

Mycobacterium tuberculosis (Mtb) induces necrosis of infected cells to evade immune responses. Recently, we found that Mtb uses the protein CpnT to kill human macrophages by secreting its C-terminal domain, named tuberculosis necrotizing toxin (TNT), which induces necrosis by an unknown mechanism. Here we show that TNT gains access to the cytosol of Mtb-infected macrophages, where it hydrolyzes the essential coenzyme NAD+. Expression or injection of a noncatalytic TNT mutant showed no cytotoxicity in macrophages or in zebrafish zygotes, respectively, thus demonstrating that the NAD+ glycohydrolase activity is required for TNT-induced cell death. To prevent self-poisoning, Mtb produces an immunity factor for TNT (IFT) that binds TNT and inhibits its activity. The crystal structure of the TNT–IFT complex revealed a new NAD+ glycohydrolase fold of TNT, the founding member of a toxin family widespread in pathogenic microorganisms.

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

Figure 1: TNT is a new β-NAD+ glycohydrolase of M. tuberculosis.
Figure 2: M. tuberculosis produces the endogenous immunity factor IFT, which binds to and inactivates TNT.
Figure 3: Atomic structure of the TNT–IFT complex.
Figure 4: Identification of TNT residues involved in NAD+ hydrolysis and cytotoxicity in macrophages.
Figure 5: M. tuberculosis secretes TNT into the macrophage cytosol.
Figure 6: TNT is toxic in zebrafish.

Similar content being viewed by others

Accession codes

Primary accessions

Protein Data Bank

Referenced accessions

Protein Data Bank

References

  1. Russell, D.G., Barry, C.E. III & Flynn, J.L. Tuberculosis: what we don't know can, and does, hurt us. Science 328, 852–856 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Behar, S.M., Divangahi, M. & Remold, H.G. Evasion of innate immunity by Mycobacterium tuberculosis: is death an exit strategy? Nat. Rev. Microbiol. 8, 668–674 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Butler, R.E. et al. The balance of apoptotic and necrotic cell death in Mycobacterium tuberculosis infected macrophages is not dependent on bacterial virulence. PLoS ONE 7, e47573 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Behar, S.M. et al. Apoptosis is an innate defense function of macrophages against Mycobacterium tuberculosis. Mucosal Immunol. 4, 279–287 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Divangahi, M., Behar, S.M. & Remold, H. Dying to live: how the death modality of the infected macrophage affects immunity to tuberculosis. Adv. Exp. Med. Biol. 783, 103–120 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Smith, J. et al. Evidence for pore formation in host cell membranes by ESX-1-secreted ESAT-6 and its role in Mycobacterium marinum escape from the vacuole. Infect. Immun. 76, 5478–5487 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Lee, J., Repasy, T., Papavinasasundaram, K., Sassetti, C. & Kornfeld, H. Mycobacterium tuberculosis induces an atypical cell death mode to escape from infected macrophages. PLoS ONE 6, e18367 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Abdallah, A.M. et al. Mycobacterial secretion systems ESX-1 and ESX-5 play distinct roles in host cell death and inflammasome activation. J. Immunol. 187, 4744–4753 (2011).

    Article  CAS  PubMed  Google Scholar 

  9. Henkel, J.S., Baldwin, M.R. & Barbieri, J.T. Toxins from bacteria. EXS 100, 1–29 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Gordon, S.V., Bottai, D., Simeone, R., Stinear, T.P. & Brosch, R. Pathogenicity in the tubercle bacillus: molecular and evolutionary determinants. BioEssays 31, 378–388 (2009).

    Article  CAS  PubMed  Google Scholar 

  11. Mukhopadhyay, S., Nair, S. & Ghosh, S. Pathogenesis in tuberculosis: transcriptomic approaches to unraveling virulence mechanisms and finding new drug targets. FEMS Microbiol. Rev. 36, 463–485 (2012).

    Article  CAS  PubMed  Google Scholar 

  12. Forrellad, M.A. et al. Virulence factors of the Mycobacterium tuberculosis complex. Virulence 4, 3–66 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Danilchanka, O. et al. An outer membrane channel protein of Mycobacterium tuberculosis with exotoxin activity. Proc. Natl. Acad. Sci. USA 111, 6750–6755 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Belenky, P., Bogan, K.L. & Brenner, C. NAD+ metabolism in health and disease. Trends Biochem. Sci. 32, 12–19 (2007).

    Article  CAS  PubMed  Google Scholar 

  15. Lin, S.J. & Guarente, L. Nicotinamide adenine dinucleotide, a metabolic regulator of transcription, longevity and disease. Curr. Opin. Cell Biol. 15, 241–246 (2003).

    Article  CAS  PubMed  Google Scholar 

  16. Du, L. et al. Intra-mitochondrial poly(ADP-ribosylation) contributes to NAD+ depletion and cell death induced by oxidative stress. J. Biol. Chem. 278, 18426–18433 (2003).

    Article  CAS  PubMed  Google Scholar 

  17. Zong, W.X. & Thompson, C.B. Necrotic death as a cell fate. Genes Dev. 20, 1–15 (2006).

    Article  CAS  PubMed  Google Scholar 

  18. Hara, N. et al. Elevation of cellular NAD levels by nicotinic acid and involvement of nicotinic acid phosphoribosyltransferase in human cells. J. Biol. Chem. 282, 24574–24582 (2007).

    Article  CAS  PubMed  Google Scholar 

  19. Russell, D.G. New ways to arrest phagosome maturation. Nat. Cell Biol. 9, 357–359 (2007).

    Article  CAS  PubMed  Google Scholar 

  20. Artman, M. & Bekierkunst, A. Mycobacterium tuberculosis H37Rv grown in vivo: nature of the inhibitor of lactic dehydrogenase of Mycobacterium phlei. Proc. Soc. Exp. Biol. Med. 106, 610–614 (1961).

    Article  CAS  PubMed  Google Scholar 

  21. Gopinathan, K.P., Sirsi, M. & Vaidyanathan, C.S. Nicotinamide-adenine dinucleotide glycohydrolase of Mycobacterium tuberculosis H37Rv. Biochem. J. 91, 277–282 (1964).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Stevens, D.L., Salmi, D.B., McIndoo, E.R. & Bryant, A.E. Molecular epidemiology of nga and NAD glycohydrolase/ADP-ribosyltransferase activity among Streptococcus pyogenes causing streptococcal toxic shock syndrome. J. Infect. Dis. 182, 1117–1128 (2000).

    Article  CAS  PubMed  Google Scholar 

  23. Meehl, M.A., Pinkner, J.S., Anderson, P.J., Hultgren, S.J. & Caparon, M.G. A novel endogenous inhibitor of the secreted streptococcal NAD-glycohydrolase. PLoS Pathog. 1, e35 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Gopinathan, K.P., Sirsi, M. & Ramakrishnan, T. Nicotin-amide-adenine nucleotides of Mycobacterium tuberculosis H37Rv. Biochem. J. 87, 444–448 (1963).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gopinathan, K.P., Ramakrishnan, T. & Vaidyanathan, C.S. Purification and properties of an inhibitor for nicotinamide-adenine dinucleotidase from Mycobacterium tuberculosis H-37-Rv. Arch. Biochem. Biophys. 113, 376–382 (1966).

    Article  CAS  PubMed  Google Scholar 

  26. Reddy, B.G. et al. 1.55 A resolution X-ray crystal structure of Rv3902c from Mycobacterium tuberculosis. Acta Crystallogr. F Struct. Biol. Commun. 70, 414–417 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Smith, C.L. et al. Structural basis of Streptococcus pyogenes immunity to its NAD+ glycohydrolase toxin. Structure 19, 192–202 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Choe, S. et al. The crystal structure of diphtheria toxin. Nature 357, 216–222 (1992).

    Article  CAS  PubMed  Google Scholar 

  29. Zhang, R.G. et al. The three-dimensional crystal structure of cholera toxin. J. Mol. Biol. 251, 563–573 (1995).

    Article  CAS  PubMed  Google Scholar 

  30. Fieldhouse, R.J., Turgeon, Z., White, D. & Merrill, A.R. Cholera- and anthrax-like toxins are among several new ADP-ribosyltransferases. PLoS Comput. Biol. 6, e1001029 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  31. Tweten, R.K., Barbieri, J.T. & Collier, R.J. Diphtheria toxin: effect of substituting aspartic acid for glutamic acid 148 on ADP-ribosyltransferase activity. J. Biol. Chem. 260, 10392–10394 (1985).

    Article  CAS  PubMed  Google Scholar 

  32. Blanke, S.R., Huang, K. & Collier, R.J. Active-site mutations of diphtheria toxin: role of tyrosine-65 in NAD binding and ADP-ribosylation. Biochemistry 33, 15494–15500 (1994).

    Article  CAS  PubMed  Google Scholar 

  33. Van Gool, F. et al. Intracellular NAD levels regulate tumor necrosis factor protein synthesis in a sirtuin-dependent manner. Nat. Med. 15, 206–210 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Lord, J.M., Smith, D.C. & Roberts, L.M. Toxin entry: how bacterial proteins get into mammalian cells. Cell. Microbiol. 1, 85–91 (1999).

    Article  CAS  PubMed  Google Scholar 

  35. Torgersen, M.L., Skretting, G., van Deurs, B. & Sandvig, K. Internalization of cholera toxin by different endocytic mechanisms. J. Cell Sci. 114, 3737–3747 (2001).

    Article  CAS  PubMed  Google Scholar 

  36. Barth, H. & Aktories, K. New insights into the mode of action of the actin ADP-ribosylating virulence factors Salmonella enterica SpvB and Clostridium botulinum C2 toxin. Eur. J. Cell Biol. 90, 944–950 (2011).

    Article  CAS  PubMed  Google Scholar 

  37. Manzanillo, P.S. et al. The ubiquitin ligase parkin mediates resistance to intracellular pathogens. Nature 501, 512–516 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Kremer, L., Maughan, W.N., Wilson, R.A., Dover, L.G. & Besra, G.S. The M. tuberculosis antigen 85 complex and mycolyltransferase activity. Lett. Appl. Microbiol. 34, 233–237 (2002).

    Article  CAS  PubMed  Google Scholar 

  39. Manzanillo, P.S., Shiloh, M.U., Portnoy, D.A. & Cox, J.S. Mycobacterium tuberculosis activates the DNA-dependent cytosolic surveillance pathway within macrophages. Cell Host Microbe 11, 469–480 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Spitsbergen, J.M. & Kent, M.L. The state of the art of the zebrafish model for toxicology and toxicologic pathology research–advantages and current limitations. Toxicol. Pathol. 31 (suppl.), 62–87 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Saslowsky, D.E. et al. Intoxication of zebrafish and mammalian cells by cholera toxin depends on the flotillin/reggie proteins but not Derlin-1 or -2. J. Clin. Invest. 120, 4399–4409 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Vanden Broeck, D., Horvath, C. & De Wolf, M.J. Vibrio cholerae: cholera toxin. Int. J. Biochem. Cell Biol. 39, 1771–1775 (2007).

    Article  CAS  PubMed  Google Scholar 

  43. Moraco, A.H. & Kornfeld, H. Cell death and autophagy in TB. Semin. Immunol. 26, 497–511 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Ha, H.C. & Snyder, S.H. Poly(ADP-ribose) polymerase is a mediator of necrotic cell death by ATP depletion. Proc. Natl. Acad. Sci. USA 96, 13978–13982 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Herceg, Z. & Wang, Z.Q. Failure of poly(ADP-ribose) polymerase cleavage by caspases leads to induction of necrosis and enhanced apoptosis. Mol. Cell. Biol. 19, 5124–5133 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. McCommis, K.S. & Finck, B.N. Mitochondrial pyruvate transport: a historical perspective and future research directions. Biochem. J. 466, 443–454 (2015).

    Article  CAS  PubMed  Google Scholar 

  47. Kennedy, C.L., Smith, D.J., Lyras, D., Chakravorty, A. & Rood, J.I. Programmed cellular necrosis mediated by the pore-forming alpha-toxin from Clostridium septicum. PLoS Pathog. 5, e1000516 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Wong, K.W. & Jacobs, W.R. Jr. Critical role for NLRP3 in necrotic death triggered by Mycobacterium tuberculosis. Cell. Microbiol. 13, 1371–1384 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Heine, K., Pust, S., Enzenmuller, S. & Barth, H. ADP-ribosylation of actin by the Clostridium botulinum C2 toxin in mammalian cells results in delayed caspase-dependent apoptotic cell death. Infect. Immun. 76, 4600–4608 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Morimoto, H. & Bonavida, B. Diphtheria toxin- and Pseudomonas A toxin-mediated apoptosis. ADP ribosylation of elongation factor-2 is required for DNA fragmentation and cell lysis and synergy with tumor necrosis factor-alpha. J. Immunol. 149, 2089–2094 (1992).

    CAS  PubMed  Google Scholar 

  51. Pessina, A. et al. Bcl-2 down modulation in WEHI-3B/CTRES cells resistant to Cholera Toxin (CT)-induced apoptosis. Cell Res. 16, 306–312 (2006).

    Article  CAS  PubMed  Google Scholar 

  52. Wolf, A.J. et al. Mycobacterium tuberculosis infects dendritic cells with high frequency and impairs their function in vivo. J. Immunol. 179, 2509–2519 (2007).

    Article  CAS  PubMed  Google Scholar 

  53. Eum, S.Y. et al. Neutrophils are the predominant infected phagocytic cells in the airways of patients with active pulmonary TB. Chest 137, 122–128 (2010).

    Article  PubMed  Google Scholar 

  54. Ajdic, D., McShan, W.M., Savic, D.J., Gerlach, D. & Ferretti, J.J. The NAD-glycohydrolase (nga) gene of Streptococcus pyogenes. FEMS Microbiol. Lett. 191, 235–241 (2000).

    Article  CAS  PubMed  Google Scholar 

  55. Finn, R.D. et al. Pfam: the protein families database. Nucleic Acids Res. 42, D222–D230 (2014).

    Article  CAS  PubMed  Google Scholar 

  56. Flórido, M., Cooper, A.M. & Appelberg, R. Immunological basis of the development of necrotic lesions following Mycobacterium avium infection. Immunology 106, 590–601 (2002).

    Article  PubMed  PubMed Central  Google Scholar 

  57. Sridharan, H. & Upton, J.W. Programmed necrosis in microbial pathogenesis. Trends Microbiol. 22, 199–207 (2014).

    Article  CAS  PubMed  Google Scholar 

  58. Danilchanka, O. et al. An outer membrane channel protein of Mycobacterium tuberculosis with exotoxin activity. Proc. Natl. Acad. Sci. USA 111, 6750–6755 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Studier, F.W. Protein production by auto-induction in high density shaking cultures. Protein Expr. Purif. 41, 207–234 (2005).

    Article  CAS  PubMed  Google Scholar 

  60. Otwinowski, Z. & Minor, W. Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307–326 (1997).

    Article  CAS  PubMed  Google Scholar 

  61. McCoy, A.J. Solving structures of protein complexes by molecular replacement with Phaser. Acta Crystallogr. D Biol. Crystallogr. 63, 32–41 (2007).

    Article  CAS  PubMed  Google Scholar 

  62. Cowtan, K. The Buccaneer software for automated model building. 1. Tracing protein chains. Acta Crystallogr. D Biol. Crystallogr. 62, 1002–1011 (2006).

    Article  CAS  PubMed  Google Scholar 

  63. Adams, P.D. et al. PHENIX: building new software for automated crystallographic structure determination. Acta Crystallogr. D Biol. Crystallogr. 58, 1948–1954 (2002).

    Article  CAS  PubMed  Google Scholar 

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

    Article  PubMed  CAS  Google Scholar 

  65. Chen, V.B. et al. MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D Biol. Crystallogr. 66, 12–21 (2010).

    Article  CAS  PubMed  Google Scholar 

  66. Krissinel, E. & Henrick, K. Inference of macromolecular assemblies from crystalline state. J. Mol. Biol. 372, 774–797 (2007).

    Article  CAS  PubMed  Google Scholar 

  67. Laskowski, R.A. PDBsum new things. Nucleic Acids Res. 37, D355–D359 (2009).

    Article  CAS  PubMed  Google Scholar 

  68. Trott, O. & Olson, A.J. AutoDock Vina: improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 31, 455–461 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Moriarty, N.W., Grosse-Kunstleve, R.W. & Adams, P.D. electronic Ligand Builder and Optimization Workbench (eLBOW): a tool for ligand coordinate and restraint generation. Acta Crystallogr. D Biol. Crystallogr. 65, 1074–1080 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Ghosh, J., Anderson, P.J., Chandrasekaran, S. & Caparon, M.G. Characterization of Streptococcus pyogenes beta-NAD+ glycohydrolase: re-evaluation of enzymatic properties associated with pathogenesis. J. Biol. Chem. 285, 5683–5694 (2010).

    Article  CAS  PubMed  Google Scholar 

  71. Meehl, M.A., Pinkner, J.S., Anderson, P.J., Hultgren, S.J. & Caparon, M.G. A novel endogenous inhibitor of the secreted streptococcal NAD-glycohydrolase. PLoS Pathog. 1, e35 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Manzanillo, P.S. et al. The ubiquitin ligase parkin mediates resistance to intracellular pathogens. Nature 501, 512–516 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank L. DeLucas for providing the Rv3902c structure before its publication, J. Pascal for stimulating discussions, Y. Wang for initial purification of IFT, the scientific staff at National Synchrotron Light Source beamlines X6A and X29 for beam time, the University of Alabama at Birmingham Zebrafish Core for generous support of the zebrafish experiments, F. Wolschendorf for use of his microscope, and P. Patel and U. Tak for technical assistance. Some work was carried out at the Sidney Kimmel Cancer Center X-ray Crystallography and Molecular Interaction Facility, which is supported in part by the US National Cancer Institute Cancer Center Support Grant P30 CA56036. This work was supported by the US National Institutes of Health grants GM100888 (G.C.), AI63432 (M.N.), AI083632 (M.N.) and AI074805 (M.N.).

Author information

Authors and Affiliations

Authors

Contributions

J.S., A. Siroy, A. Speer, K.S.D., G.C. and M.N. conceived and designed experiments and analyzed data; J.S., A. Siroy, R.K.L., K.S.D., A. Speer and G.C. performed experiments; J.S., G.C. and M.N. wrote and edited the paper.

Corresponding authors

Correspondence to Gino Cingolani or Michael Niederweis.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 TNT reduces cellular NAD+ levels by degradation.

(a) Global gene expression after induction of tnt in E. coli was determined by deep sequencing. Gene expression changes were compared to E. coli containing an empty vector and are plotted as a ratio (log2-fold changes) against the significance of the difference in expression (Log Odds; −log10 q-value). Only genes with expression ratios larger than four-fold (log2 ratios ≥ 2 or ≤ −2) and log Odds > 1.3 (q-values < 0.05) were included. (b) The amount of NAD+ extracted from 5 × 108 E. coli or 1 × 106 Jurkat cells 2 h or 24 h post-induction of tnt expression, respectively, was measured by a colorimetric enzyme coupling assay. Data represent the means ± s.d. of three independent experiments. (c) NAD+ hydrolysis by TNT releases ADP-ribose (and nicotinamide) as determined by thin-layer chromatography using radiolabeled NAD+, cyclic ADP-ribose, and ADP-ribose as standards as described in Methods.

Supplementary Figure 2 Expression and purification of TNT–IFT complex.

(a) Schematic flow chart of the purification procedure with accompanying yield at various steps. (b) Fractions were taken from each step of the purification process and analyzed in Coomassie-stained SDS-polyacrylamide gels. The final protein complex after gel filtration chromatography (pooled fractions) was used for X-ray crystallization. (c) Fractions from purification of free TNT were analyzed by Coomassie staining. TNT was separated from the TNT-IFT complex by heating at 70°C for 10 min. TNT was recovered from the soluble fraction while IFT remained in the precipitate after heating. WCL, whole cell lysate; SN, soluble.

Supplementary Figure 3 Kinetic analysis and pH dependency of NAD+ hydrolysis by TNT.

(a) 10 nM of purified TNT were incubated with different amounts of NAD+ at pH 7.4. After 1 min, 2.5 min, 5 min, and 10 min, aliquots were quenched by addition of 5 M NaOH and developed for 1 h in the dark. The relative fluorescence of each sample was measured and correlated to a standard concentration curve for NAD+. (b) The observed rate constant (kobs) of TNT enzymatic activity was plotted against NAD+ concentration to determine the kcat. (c) The pH dependency of TNT activity was determined by measuring the initial velocity of NAD+ hydrolysis at a TNT concentration of 10 nM and a substrate concentration of 200 µM. All data represent the means ± s.d. of three independent experiments.

Supplementary Figure 4 Model of the TNT–NAD+ complex.

(a) Docking of NAD+ inside TNT putative active site crevice performed using AutoDock Vina. NAD+ is shown in sticks and TNT as a green "Connolly" surface. (b) Ribbon representation of the docking model in panel (a) with magnified view of TNT residues in the palm-domain predicted to be important in the NAD+ binding and hydrolysis.

Supplementary Figure 5 Binding of TNT-mutant proteins to the antitoxin IFT.

The purified mutant proteins TNTQ822A (a) or TNTH792N Q822K (b) were used in an in vitro pull-down assay with purified MBP-IFT. MBP-IFT alone bound amylose resin while the TNTQ822A or TNTH792N Q822K proteins did not. Incubation of MBP-IFT (1 µg) with equimolar amounts of TNTQ822A or TNTH792N Q822K proteins resulted in binding of both TNT proteins to amylose resin and elution with maltose.

Supplementary Figure 6 Effect of nicotinamide (NM) and nicotinic acid (NA) on cellular NAD+ content and TNT activity.

(a) The amount of NAD+ extracted from 1 × 106 Jurkat cells supplemented with NM (5 mM) or NA (10 µM) or in combination for 3 h and 6 h were quantified by the EnzyFluo assay kit. These baseline NAD+ levels are similar to those previously reported in HEK293 cells and correspond to an NAD+ concentration of 503 µM. (b) NAD+-glycohydrolase activities were determined by the amount of NAD+ remaining after incubation of BSA or TNT (100 ng) with 1 µM NAD+ at 37°C for 30 min in the presence or absence of NM (5 mM) and/or NA (10 µM). Data in (a-b) represent the means ± s.d. of three independent experiments. (c) Jurkat cells expressing tnt were untreated (Ctrl), incubated with 5 mM nicotinamide or 10 µM nicotinic acid, or in combination. Then, tnt expression was induced with 100 ng/mL doxycycline for 2 h. Following removal of doxycycline, cells were incubated in fresh media for 24 h. Cell death was assessed using the 7AAD dye and flow cytometry. 7AAD: 7-amino-actinomycin D, SSC: Side scatter.

Supplementary Figure 7 TNT requires cytosolic localization to induce cell death.

(a) Exogenous purified TNT or TNT-IFT complex (10 µg/mL) was added to THP-1 macrophages and incubated for 24 h. Thereafter, cell death was quantified by 7AAD staining and flow cytometry. (b, c) purified TNT or TNT-IFT complex (250 µg/mL) was adsorbed onto 4 µm latex beads and tested for NAD+ glycohydrolase activity (b) as described in Methods and subsequently used to infect THP-1 macrophages in order to load TNT intracellularly. 48 h post infection, macrophages were assessed for viability with 7AAD staining (c). Data in (b) represent the means ± s.d. of three independent experiments.

Supplementary Figure 8 Plasmid construction for this study.

Plasmid names are shown in dark grey boxes. The oligonucleotide pairs used for PCR amplification and restriction enzymes utilized for cloning are listed in white boxes. The template DNA for PCR is indicated below the oligonucleotide pairs. When chromosomal DNA was used as a template, the originating organism is indicated.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–8, Supplementary Tables 1–5 and Supplementary Note (PDF 4004 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sun, J., Siroy, A., Lokareddy, R. et al. The tuberculosis necrotizing toxin kills macrophages by hydrolyzing NAD. Nat Struct Mol Biol 22, 672–678 (2015). https://doi.org/10.1038/nsmb.3064

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nsmb.3064

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