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.

  • Review Article
  • Published:

Subversion of host genome integrity by bacterial pathogens

Key Points

  • Bacteria have evolved a multitude of strategies for manipulating and exploiting the host to establish an infection and replicate, while keeping the host cells, and thus their replicative niche, alive.

  • Bacterial infections can cause DNA damage in host cells, either directly by secreting genotoxic proteins or through mechanisms involving host response to the infection.

  • Some bacterial species simultaneously disrupt the host's DNA damage response through multiple pathways. This can lead to genomic instability and, in consequence, tumorigenic transformation.

  • The environment established upon infection may promote the escape of such transformed cells from tumour surveillance mechanisms.

  • Chronic infections that evade the immune system seem to be particularly dangerous in this context, and several of them, such as chronic infections by Helicobacter pylori, Chlamydia trachomatis and Salmonella enterica subsp. enterica serovar Typhi, have been associated with the development of human cancers.

Abstract

Mammalian cells possess sophisticated genome surveillance and repair mechanisms, executed by the so-called DNA damage response (DDR), failure of which leads to accumulation of DNA damage and genomic instability. Mounting evidence suggests that bacterial infections can elicit DNA damage in host cells, and certain pathogens induce such damage as part of their multi-faceted infection programme. Bacteria-mediated DNA damage can occur either directly through the formation of toxins with genotoxic activities or indirectly as a result of the activation of cell-autonomous or immune defence mechanisms against the pathogen. Moreover, host-cell signalling routes involved in the DDR can be altered in response to an infection, and this, in the context of DNA damage elicited by the pathogen, has the potential to trigger mutations and cancer.

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: Bacteria-induced DNA damage.
Figure 2: Modulation of the DNA damage response by Helicobacter pylori.
Figure 3: Modulation of the DNA damage response by Chlamydia trachomatis.
Figure 4: Bacteria in tumorigenic transformation.

Similar content being viewed by others

References

  1. Lindahl, T. & Barnes, D. E. Repair of endogenous DNA damage. Cold Spring Harb. Symp. Quant. Biol. 65, 127–133 (2000).

    Article  CAS  PubMed  Google Scholar 

  2. Ciccia, A. & Elledge, S. J. The DNA damage response: making it safe to play with knives. Mol. Cell 40, 179–204 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Hoeijmakers, J. H. Genome maintenance mechanisms for preventing cancer. Nature 411, 366–374 (2001).

    Article  CAS  PubMed  Google Scholar 

  4. IARC. Schistosomes, liver flukes and Helicobacter pylori. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans. Lyon, 7–14 June 1994. IARC Monogr. Eval. Carcinog. Risks Hum. 61, 1–241 (1994).

  5. Azenabor, A. A. & Mahony, J. B. Generation of reactive oxygen species and formation and membrane lipid peroxides in cells infected with Chlamydia trachomatis. Int. J. Infect. Dis. 4, 46–50 (2000).

    Article  CAS  PubMed  Google Scholar 

  6. Wu, M. et al. Host DNA repair proteins in response to Pseudomonas aeruginosa in lung epithelial cells and in mice. Infect. Immun. 79, 75–87 (2011).

    Article  CAS  PubMed  Google Scholar 

  7. Hardbower, D. M., de Sablet, T., Chaturvedi, R. & Wilson, K. T. Chronic inflammation and oxidative stress: the smoking gun for Helicobacter pylori-induced gastric cancer? Gut Microbes 4, 475–481 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Abdul-Sater, A. A. et al. Enhancement of reactive oxygen species production and chlamydial infection by the mitochondrial Nod-like family member NLRX1. J. Biol. Chem. 285, 41637–41645 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Chumduri, C., Gurumurthy, R. K., Zadora, P. K., Mi, Y. & Meyer, T. F. Chlamydia infection promotes host DNA damage and proliferation but impairs the DNA damage response. Cell Host Microbe 13, 746–758 (2013). This study shows that C. trachomatis induces persistent DSBs in infected cells and simultaneously suppresses recruitment of ATM and 53BP1, while maintaining cells in a proliferative state.

    Article  CAS  PubMed  Google Scholar 

  10. Tattoli, I. et al. NLRX1 is a mitochondrial NOD-like receptor that amplifies NF-κB and JNK pathways by inducing reactive oxygen species production. EMBO Rep. 9, 293–300 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Lam, G. Y. et al. Listeriolysin O suppresses phospholipase C-mediated activation of the microbicidal NADPH oxidase to promote Listeria monocytogenes infection. Cell Host Microbe 10, 627–634 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Leitao, E. et al. Listeria monocytogenes induces host DNA damage and delays the host cell cycle to promote infection. Cell Cycle 13, 928–940 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Samba-Louaka, A. et al. Listeria monocytogenes dampens the DNA damage response. PLoS Pathog. 10, e1004470 (2014). This paper shows that the L. monocytogenes toxin LLO degrades the DDR sensor MRE11, despite DSB induction by the bacteria, and that this repression of the DDR promotes infection.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Plotkowski, M. C. et al. Early mitochondrial dysfunction, superoxide anion production, and DNA degradation are associated with non-apoptotic death of human airway epithelial cells induced by Pseudomonas aeruginosa exotoxin A. Am. J. Respir. Cell. Mol. Biol. 26, 617–626 (2002).

    Article  CAS  PubMed  Google Scholar 

  15. Saliba, A. M. et al. Implications of oxidative stress in the cytotoxicity of Pseudomonas aeruginosa ExoU. Microbes Infect. 8, 450–459 (2006).

    Article  CAS  PubMed  Google Scholar 

  16. Zorov, D. B., Juhaszova, M. & Sollott, S. J. Mitochondrial ROS-induced ROS release: an update and review. Biochim. Biophys. Acta 1757, 509–517 (2006).

    Article  CAS  PubMed  Google Scholar 

  17. Wei, L. & Dirksen, R. T. Perspectives on: SGP symposium on mitochondrial physiology and medicine: mitochondrial superoxide flashes: from discovery to new controversies. J. Gen. Physiol. 139, 425–434 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Roy, A. et al. Mitochondria-dependent reactive oxygen species-mediated programmed cell death induced by 3,3′-diindolylmethane through inhibition of F0F1-ATP synthase in unicellular protozoan parasite Leishmania donovani. Mol. Pharmacol. 74, 1292–1307 (2008).

    Article  CAS  PubMed  Google Scholar 

  19. Rai, P. et al. Streptococcus pneumoniae secretes hydrogen peroxide leading to DNA damage and apoptosis in lung cells. Proc. Natl Acad. Sci. USA 112, E3421–E3430 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Chaturvedi, R. et al. Spermine oxidase mediates the gastric cancer risk associated with Helicobacter pylori CagA. Gastroenterology 141, 1696–1708.e2 (2011).

    Article  CAS  PubMed  Google Scholar 

  21. Kim, J. M. et al. Vacuolating cytotoxin in Helicobacter pylori water-soluble proteins upregulates chemokine expression in human eosinophils via Ca2+ influx, mitochondrial reactive oxygen intermediates, and NF-κB activation. Infect. Immun. 75, 3373–3381 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Gong, M., Ling, S. S., Lui, S. Y., Yeoh, K. G. & Ho, B. Helicobacter pylori γ-glutamyl transpeptidase is a pathogenic factor in the development of peptic ulcer disease. Gastroenterology 139, 564–573 (2010).

    Article  CAS  PubMed  Google Scholar 

  23. Uberti, A. F. et al. Pro-inflammatory properties and neutrophil activation by Helicobacter pylori urease. Toxicon 69, 240–249 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Wang, G., Hong, Y., Olczak, A., Maier, S. E. & Maier, R. J. Dual roles of Helicobacter pylori NapA in inducing and combating oxidative stress. Infect. Immun. 74, 6839–6846 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Gorlach, A., Bertram, K., Hudecova, S. & Krizanova, O. Calcium and ROS: a mutual interplay. Redox Biol. 6, 260–271 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Ermak, G. & Davies, K. J. Calcium and oxidative stress: from cell signaling to cell death. Mol. Immunol. 38, 713–721 (2002).

    Article  CAS  PubMed  Google Scholar 

  27. Raza, Y. et al. Oxidative DNA damage as a potential early biomarker of Helicobacter pylori associated carcinogenesis. Pathol. Oncol. Res. 20, 839–846 (2014).

    Article  CAS  PubMed  Google Scholar 

  28. Kidane, D., Murphy, D. L. & Sweasy, J. B. Accumulation of abasic sites induces genomic instability in normal human gastric epithelial cells during Helicobacter pylori infection. Oncogenesis 3, e128 (2014). This report shows that DNA damage induced by H. pylori leads to the accumulation of abasic sites, and that BER processing of these sites generates DSBs.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Toller, I. M. et al. Carcinogenic bacterial pathogen Helicobacter pylori triggers DNA double-strand breaks and a DNA damage response in its host cells. Proc. Natl Acad. Sci. USA 108, 14944–14949 (2011). This paper finds that direct cell contact is sufficient for the induction of DSBs in H. pylori -infected cells and that prolonged infection leads to persistent unrepaired breaks.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Hartung, M. L. et al. H. pylori-induced DNA strand breaks are introduced by nucleotide excision repair endonucleases and promote NF-κB target gene expression. Cell Rep. 13, 70–79 (2015). This study reports that NER endonucleases induce DSB breaks in H. pylori -infected cells, which lead to increased NF-κB target gene activation, thus promoting the survival of host cells.

    Article  CAS  PubMed  Google Scholar 

  31. Matsumoto, Y. et al. Helicobacter pylori infection triggers aberrant expression of activation-induced cytidine deaminase in gastric epithelium. Nat. Med. 13, 470–476 (2007).

    Article  CAS  PubMed  Google Scholar 

  32. Shimizu, T. et al. Accumulation of somatic mutations in TP53 in gastric epithelium with Helicobacter pylori infection. Gastroenterology 147, 407–417.e3 (2014).

    Article  CAS  PubMed  Google Scholar 

  33. Mechtcheriakova, D., Svoboda, M., Meshcheryakova, A. & Jensen-Jarolim, E. Activation-induced cytidine deaminase (AID) linking immunity, chronic inflammation, and cancer. Cancer Immunol. Immunother. 61, 1591–1598 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Liu, M. et al. Two levels of protection for the B cell genome during somatic hypermutation. Nature 451, 841–845 (2008).

    Article  CAS  PubMed  Google Scholar 

  35. Gurumurthy, R. K. et al. Dynamin-mediated lipid acquisition is essential for Chlamydia trachomatis development. Mol. Microbiol. 94, 186–201 (2014).

    Article  CAS  PubMed  Google Scholar 

  36. Elwell, C., Mirrashidi, K. & Engel, J. Chlamydia cell biology and pathogenesis. Nat. Rev. Microbiol. 14, 385–400 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Zahlten, J. et al. Streptococcus pneumoniae-induced oxidative stress in lung epithelial cells depends on pneumococcal autolysis and is reversible by resveratrol. J. Infect. Dis. 211, 1822–1830 (2015).

    Article  CAS  PubMed  Google Scholar 

  38. Oosthuysen, W. F., Mueller, T., Dittrich, M. T. & Schubert-Unkmeir, A. Neisseria meningitidis causes cell cycle arrest of human brain microvascular endothelial cells at S phase via p21 and cyclin G2. Cell. Microbiol. 18, 46–65 (2016).

    Article  CAS  PubMed  Google Scholar 

  39. Grasso, F. & Frisan, T. Bacterial genotoxins: merging the DNA damage response into infection biology. Biomolecules 5, 1762–1782 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Homburg, S., Oswald, E., Hacker, J. & Dobrindt, U. Expression analysis of the colibactin gene cluster coding for a novel polyketide in Escherichia coli. FEMS Microbiol. Lett. 275, 255–262 (2007).

    Article  CAS  PubMed  Google Scholar 

  41. Nougayrede, J. P. et al. Escherichia coli induces DNA double-strand breaks in eukaryotic cells. Science 313, 848–851 (2006). This study identified the genotoxic activity of the colibactin biosynthetic pathway encoded by the pks genomic island of E. coli.

    Article  CAS  PubMed  Google Scholar 

  42. Putze, J. et al. Genetic structure and distribution of the colibactin genomic island among members of the family Enterobacteriaceae. Infect. Immun. 77, 4696–4703 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Vizcaino, M. I. & Crawford, J. M. The colibactin warhead crosslinks DNA. Nat. Chem. 7, 411–417 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Guidi, R. et al. Salmonella enterica delivers its genotoxin through outer membrane vesicles secreted from infected cells. Cell. Microbiol. 15, 2034–2050 (2013).

    Article  CAS  PubMed  Google Scholar 

  45. Bezine, E., Vignard, J. & Mirey, G. The cytolethal distending toxin effects on mammalian cells: a DNA damage perspective. Cells 3, 592–615 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. DiRienzo, J. M. Uptake and processing of the cytolethal distending toxin by mammalian cells. Toxins (Basel) 6, 3098–3116 (2014).

    Article  CAS  Google Scholar 

  47. Fedor, Y. et al. From single-strand breaks to double-strand breaks during S-phase: a new mode of action of the Escherichia coli cytolethal distending toxin. Cell. Microbiol. 15, 1–15 (2013).

    Article  CAS  PubMed  Google Scholar 

  48. Sandvig, K. Shiga toxins. Toxicon 39, 1629–1635 (2001).

    Article  CAS  PubMed  Google Scholar 

  49. Johannes, L. & Romer, W. Shiga toxins — from cell biology to biomedical applications. Nat. Rev. Microbiol. 8, 105–116 (2010).

    Article  CAS  PubMed  Google Scholar 

  50. Brigotti, M. et al. Shiga toxin 1: damage to DNA in vitro. Toxicon 39, 341–348 (2001).

    Article  CAS  PubMed  Google Scholar 

  51. Brigotti, M. et al. Damage to nuclear DNA induced by Shiga toxin 1 and ricin in human endothelial cells. FASEB J. 16, 365–372 (2002).

    Article  CAS  PubMed  Google Scholar 

  52. Bergounioux, J. et al. Calpain activation by the Shigella flexneri effector VirA regulates key steps in the formation and life of the bacterium's epithelial niche. Cell Host Microbe 11, 240–252 (2012).

    Article  CAS  PubMed  Google Scholar 

  53. Vielfort, K. et al. Neisseria gonorrhoeae infection causes DNA damage and affects the expression of p21, 27 and p53 in non-tumor epithelial cells. J. Cell Sci. 126, 339–347 (2013).

    Article  CAS  PubMed  Google Scholar 

  54. Weyler, L. et al. Restriction endonucleases from invasive Neisseria gonorrhoeae cause double-strand breaks and distort mitosis in epithelial cells during infection. PLoS ONE 9, e114208 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Guerra, L. et al. Myc is required for activation of the ATM-dependent checkpoints in response to DNA damage. PLoS ONE 5, e8924 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Hassane, D. C., Lee, R. B. & Pickett, C. L. Campylobacter jejuni cytolethal distending toxin promotes DNA repair responses in normal human cells. Infect. Immun. 71, 541–545 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Li, L. et al. The Haemophilus ducreyi cytolethal distending toxin activates sensors of DNA damage and repair complexes in proliferating and non-proliferating cells. Cell. Microbiol. 4, 87–99 (2002).

    Article  CAS  PubMed  Google Scholar 

  58. Alaoui-El-Azher, M. et al. Role of the ATM-checkpoint kinase 2 pathway in CDT-mediated apoptosis of gingival epithelial cells. PLoS ONE 5, e11714 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Fahrer, J. et al. Cytolethal distending toxin (CDT) is a radiomimetic agent and induces persistent levels of DNA double-strand breaks in human fibroblasts. DNA Repair (Amst.) 18, 31–43 (2014).

    Article  CAS  Google Scholar 

  60. Blazkova, H. et al. Bacterial intoxication evokes cellular senescence with persistent DNA damage and cytokine signalling. J. Cell. Mol. Med. 14, 357–367 (2010).

    Article  CAS  PubMed  Google Scholar 

  61. Cortes-Bratti, X., Frisan, T. & Thelestam, M. The cytolethal distending toxins induce DNA damage and cell cycle arrest. Toxicon 39, 1729–1736 (2001).

    Article  CAS  PubMed  Google Scholar 

  62. Cortes-Bratti, X., Karlsson, C., Lagergard, T., Thelestam, M. & Frisan, T. The Haemophilus ducreyi cytolethal distending toxin induces cell cycle arrest and apoptosis via the DNA damage checkpoint pathways. J. Biol. Chem. 276, 5296–5302 (2001).

    Article  CAS  PubMed  Google Scholar 

  63. Sato, T. et al. p53-independent expression of p21CIP1/WAF1 in plasmacytic cells during G2 cell cycle arrest induced by Actinobacillus actinomycetemcomitans cytolethal distending toxin. Infect. Immun. 70, 528–534 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Guidi, R. et al. Chronic exposure to the cytolethal distending toxins of Gram-negative bacteria promotes genomic instability and altered DNA damage response. Cell. Microbiol. 15, 98–113 (2013). This report finds that long-term exposure of cells to sublethal doses of Cdt leads to mutations and chromosomal aberrations while cell survival is maintained by p38 MAPK, leading to the emergence of transformed cells exhibiting anchorage-independent growth.

    Article  CAS  PubMed  Google Scholar 

  65. Guerra, L. et al. A bacterial cytotoxin identifies the RhoA exchange factor Net1 as a key effector in the response to DNA damage. PLoS ONE 3, e2254 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Arthur, J. C. et al. Intestinal inflammation targets cancer-inducing activity of the microbiota. Science 338, 120–123 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Fernet, M., Megnin-Chanet, F., Hall, J. & Favaudon, V. Control of the G2/M checkpoints after exposure to low doses of ionising radiation: implications for hyper-radiosensitivity. DNA Repair (Amst.) 9, 48–57 (2010).

    Article  CAS  Google Scholar 

  68. Cuevas-Ramos, G. et al. Escherichia coli induces DNA damage in vivo and triggers genomic instability in mammalian cells. Proc. Natl Acad. Sci. USA 107, 11537–11542 (2010). This paper reports that infection of enterocytes with low doses of colibactin-positive E. coli leads to DSBs that are not completely repaired, leading to chromosomal instability and the emergence of anchorage-independent growth.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Talukder, K. A. et al. Activation of p53/ATM-dependent DNA damage signaling pathway by shiga toxin in mammalian cells. Microb. Pathog. 52, 311–317 (2012).

    Article  CAS  PubMed  Google Scholar 

  70. Machado, A. M. et al. Helicobacter pylori infection induces genetic instability of nuclear and mitochondrial DNA in gastric cells. Clin. Cancer Res. 15, 2995–3002 (2009).

    Article  CAS  PubMed  Google Scholar 

  71. Sepulveda, A. R. et al. CpG methylation and reduced expression of O6-methylguanine DNA methyltransferase is associated with Helicobacter pylori infection. Gastroenterology 138, 1836–1844 (2010).

    Article  CAS  PubMed  Google Scholar 

  72. Kim, J. J. et al. Helicobacter pylori impairs DNA mismatch repair in gastric epithelial cells. Gastroenterology 123, 542–553 (2002).

    Article  CAS  PubMed  Google Scholar 

  73. Park, D. I. et al. Effect of Helicobacter pylori infection on the expression of DNA mismatch repair protein. Helicobacter 10, 179–184 (2005).

    Article  CAS  PubMed  Google Scholar 

  74. Hanada, K. et al. Helicobacter pylori infection introduces DNA double-strand breaks in host cells. Infect. Immun. 82, 4182–4189 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Koeppel, M., Garcia-Alcalde, F., Glowinski, F., Schlaermann, P. & Meyer, T. F. Helicobacter pylori infection causes characteristic DNA damage patterns in human cells. Cell Rep. 11, 1703–1713 (2015).

    Article  CAS  PubMed  Google Scholar 

  76. Gaillard, H., Garcia-Muse, T. & Aguilera, A. Replication stress and cancer. Nat. Rev. Cancer 15, 276–289 (2015).

    Article  CAS  PubMed  Google Scholar 

  77. Lakin, N. D. & Jackson, S. P. Regulation of p53 in response to DNA damage. Oncogene 18, 7644–7655 (1999).

    Article  CAS  PubMed  Google Scholar 

  78. Buti, L. et al. Helicobacter pylori cytotoxin-associated gene A (CagA) subverts the apoptosis-stimulating protein of p53 (ASPP2) tumor suppressor pathway of the host. Proc. Natl Acad. Sci. USA 108, 9238–9243 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Wei, J. et al. Regulation of p53 tumor suppressor by Helicobacter pylori in gastric epithelial cells. Gastroenterology 139, 1333–1343 (2010).

    Article  CAS  PubMed  Google Scholar 

  80. Bhardwaj, V. et al. Helicobacter pylori bacteria alter the p53 stress response via ERK–HDM2 pathway. Oncotarget 6, 1531–1543 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  81. Wei, J. et al. Pathogenic bacterium Helicobacter pylori alters the expression profile of p53 protein isoforms and p53 response to cellular stresses. Proc. Natl Acad. Sci. USA 109, E2543–E2550 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Wei, J., Zaika, E. & Zaika, A. p53 Family: role of protein isoforms in human cancer. J. Nucleic Acids 2012, 687359 (2012).

    Article  CAS  PubMed  Google Scholar 

  83. Gurumurthy, R. K. et al. A loss-of-function screen reveals Ras- and Raf-independent MEK–ERK signaling during Chlamydia trachomatis infection. Sci. Signal. 3, ra21 (2010).

    Article  CAS  PubMed  Google Scholar 

  84. Di Micco, R. et al. Interplay between oncogene-induced DNA damage response and heterochromatin in senescence and cancer. Nat. Cell Biol. 13, 292–302 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Schuster-Bockler, B. & Lehner, B. Chromatin organization is a major influence on regional mutation rates in human cancer cells. Nature 488, 504–507 (2012).

    Article  CAS  PubMed  Google Scholar 

  86. Cao, X. et al. The use of transformed IMR90 cell model to identify the potential extra-telomeric effects of hTERT in cell migration and DNA damage response. BMC Biochem. 15, 17 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Padberg, I., Janssen, S. & Meyer, T. F. Chlamydia trachomatis inhibits telomeric DNA damage signaling via transient hTERT upregulation. Int. J. Med. Microbiol. 303, 463–474 (2013).

    Article  CAS  PubMed  Google Scholar 

  88. Gonzalez, E. et al. Chlamydia infection depends on a functional MDM2–p53 axis. Nat. Commun. 5, 5201 (2014).

    Article  CAS  PubMed  Google Scholar 

  89. Siegl, C., Prusty, B. K., Karunakaran, K., Wischhusen, J. & Rudel, T. Tumor suppressor p53 alters host cell metabolism to limit Chlamydia trachomatis infection. Cell Rep. 9, 918–929 (2014). This reference and reference 88 show that C. trachomatis infection depends on depletion of the tumour suppressor p53 in host cells.

    Article  CAS  PubMed  Google Scholar 

  90. Binnicker, M. J., Williams, R. D. & Apicella, M. A. Infection of human urethral epithelium with Neisseria gonorrhoeae elicits an upregulation of host anti-apoptotic factors and protects cells from staurosporine-induced apoptosis. Cell. Microbiol. 5, 549–560 (2003).

    Article  CAS  PubMed  Google Scholar 

  91. Chen, A. & Seifert, H. S. Neisseria gonorrhoeae-mediated inhibition of apoptotic signalling in polymorphonuclear leukocytes. Infect. Immun. 79, 4447–4458 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Parsonnet, J. et al. Helicobacter pylori infection and the risk of gastric carcinoma. N. Engl. J. Med. 325, 1127–1131 (1991).

    Article  CAS  PubMed  Google Scholar 

  93. Uemura, N. et al. Helicobacter pylori infection and the development of gastric cancer. N. Engl. J. Med. 345, 784–789 (2001).

    Article  CAS  PubMed  Google Scholar 

  94. Escobar-Paramo, P. et al. Large-scale population structure of human commensal Escherichia coli isolates. Appl. Environ. Microbiol. 70, 5698–5700 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Andia, M. E., Hsing, A. W., Andreotti, G. & Ferreccio, C. Geographic variation of gallbladder cancer mortality and risk factors in Chile: a population-based ecologic study. Int. J. Cancer 123, 1411–1416 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Randi, G., Franceschi, S. & La Vecchia, C. Gallbladder cancer worldwide: geographical distribution and risk factors. Int. J. Cancer 118, 1591–1602 (2006).

    Article  CAS  PubMed  Google Scholar 

  97. Chaturvedi, A. K. et al. Chlamydia pneumoniae infection and risk for lung cancer. Cancer Epidemiol. Biomarkers Prev. 19, 1498–1505 (2010).

    Article  CAS  PubMed  Google Scholar 

  98. Idahl, A. et al. Chlamydia trachomatis and Mycoplasma genitalium plasma antibodies in relation to epithelial ovarian tumors. Infect. Dis. Obstet. Gynecol. 2011, 824627 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Smith, J. S. et al. Chlamydia trachomatis and invasive cervical cancer: a pooled analysis of the IARC multicentric case-control study. Int. J. Cancer 111, 431–439 (2004).

    Article  CAS  PubMed  Google Scholar 

  100. Touati, E. When bacteria become mutagenic and carcinogenic: lessons from H. pylori. Mutat. Res. 703, 66–70 (2010).

    Article  CAS  PubMed  Google Scholar 

  101. Touati, E. et al. Chronic Helicobacter pylori infections induce gastric mutations in mice. Gastroenterology 124, 1408–1419 (2003).

    Article  CAS  PubMed  Google Scholar 

  102. Wang, K. et al. Whole-genome sequencing and comprehensive molecular profiling identify new driver mutations in gastric cancer. Nat. Genet. 46, 573–582 (2014).

    Article  CAS  PubMed  Google Scholar 

  103. Matsumoto, Y. et al. Up-regulation of activation-induced cytidine deaminase causes genetic aberrations at the CDKN2b-CDKN2a in gastric cancer. Gastroenterology 139, 1984–1994 (2010).

    Article  CAS  PubMed  Google Scholar 

  104. Endo, Y., Marusawa, H. & Chiba, T. Involvement of activation-induced cytidine deaminase in the development of colitis-associated colorectal cancers. J. Gastroenterol. 46 (Suppl. 1), 6–10 (2011).

    Article  CAS  PubMed  Google Scholar 

  105. Brown, H. M. et al. Multinucleation during C. trachomatis infections is caused by the contribution of two effector pathways. PLoS ONE 9, e100763 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Knowlton, A. E., Fowler, L. J., Patel, R. K., Wallet, S. M. & Grieshaber, S. S. Chlamydia induces anchorage independence in 3T3 cells and detrimental cytological defects in an infection model. PLoS ONE 8, e54022 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Secher, T., Samba-Louaka, A., Oswald, E. & Nougayrede, J. P. Escherichia coli producing colibactin triggers premature and transmissible senescence in mammalian cells. PLoS ONE 8, e77157 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Cougnoux, A. et al. Bacterial genotoxin colibactin promotes colon tumour growth by inducing a senescence-associated secretory phenotype. Gut 63, 1932–1942 (2014).

    Article  CAS  PubMed  Google Scholar 

  109. Dalmasso, G., Cougnoux, A., Delmas, J., Darfeuille-Michaud, A. & Bonnet, R. The bacterial genotoxin colibactin promotes colon tumor growth by modifying the tumor microenvironment. Gut Microbes 5, 675–680 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  110. Santiago, A., Li, D., Zhao, L. Y., Godsey, A. & Liao, D. p53 SUMOylation promotes its nuclear export by facilitating its release from the nuclear export receptor CRM1. Mol. Biol. Cell 24, 2739–2752 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Santiago, A., Godsey, A. C., Hossain, J., Zhao, L. Y. & Liao, D. Identification of two independent SUMO-interacting motifs in Daxx: evolutionary conservation from Drosophila to humans and their biochemical functions. Cell Cycle 8, 76–87 (2009).

    Article  CAS  PubMed  Google Scholar 

  112. Siegl, C. & Rudel, T. Modulation of p53 during bacterial infections. Nat. Rev. Microbiol. 13, 741–748 (2015).

    Article  CAS  PubMed  Google Scholar 

  113. Scanu, T. et al. Salmonella manipulation of host signaling pathways provokes cellular transformation associated with gallbladder carcinoma. Cell Host Microbe 17, 763–774 (2015). This paper shows that manipulation of the AKT and MAPK pathways during S . Typhi infection of primary cells with predisposing mutations can directly lead to transformation.

    Article  CAS  PubMed  Google Scholar 

  114. Boccellato, F. & Meyer, T. F. Bacteria moving into focus of human cancer. Cell Host Microbe 17, 728–730 (2015).

    Article  CAS  PubMed  Google Scholar 

  115. Del Bel Belluz, L. et al. The typhoid toxin promotes host survival and the establishment of a persistent asymptomatic infection. PLoS Pathog. 12, e1005528 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  116. Hartlova, A. et al. DNA damage primes the type I interferon system via the cytosolic DNA sensor STING to promote anti-microbial innate immunity. Immunity 42, 332–343 (2015).

    Article  CAS  PubMed  Google Scholar 

  117. Barber, G. N. STING: infection, inflammation and cancer. Nat. Rev. Immunol. 15, 760–770 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Uehara, T. et al. H. pylori infection is associated with DNA damage of Lgr5-positive epithelial stem cells in the stomach of patients with gastric cancer. Dig. Dis. Sci. 58, 140–149 (2013).

    Article  CAS  PubMed  Google Scholar 

  119. Sigal, M. et al. Helicobacter pylori activates and expands Lgr5+ stem cells through direct colonization of the gastric glands. Gastroenterology 148, 1392–1404.e21 (2015).

    Article  CAS  PubMed  Google Scholar 

  120. Klein, G. Evolutionary aspects of cancer resistance. Semin. Cancer Biol. 25, 10–14 (2014).

    Article  CAS  PubMed  Google Scholar 

  121. Bartfeld, S. et al. In vitro expansion of human gastric epithelial stem cells and their responses to bacterial infection. Gastroenterology 148, 126–136.e6 (2015).

    Article  PubMed  Google Scholar 

  122. Schlaermann, P. et al. A novel human gastric primary cell culture system for modelling Helicobacter pylori infection in vitro. Gut 65, 202–213 (2016).

    Article  CAS  PubMed  Google Scholar 

  123. Kryston, T. B., Georgiev, A. B., Pissis, P. & Georgakilas, A. G. Role of oxidative stress and DNA damage in human carcinogenesis. Mutat. Res. 711, 193–201 (2011).

    Article  CAS  PubMed  Google Scholar 

  124. Jackson, S. P. & Bartek, J. The DNA-damage response in human biology and disease. Nature 461, 1071–1078 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Altieri, F., Grillo, C., Maceroni, M. & Chichiarelli, S. DNA damage and repair: from molecular mechanisms to health implications. Antioxid. Redox Signal. 10, 891–937 (2008).

    Article  CAS  PubMed  Google Scholar 

  126. Nouspikel, T. DNA repair in mammalian cells: Nucleotide excision repair: variations on versatility. Cell. Mol. Life Sci. 66, 994–1009 (2009).

    Article  CAS  PubMed  Google Scholar 

  127. Blackwood, J. K. et al. End-resection at DNA double-strand breaks in the three domains of life. Biochem. Soc. Trans. 41, 314–320 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Lamarche, B. J., Orazio, N. I. & Weitzman, M. D. The MRN complex in double-strand break repair and telomere maintenance. FEBS Lett. 584, 3682–3695 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Schlacher, K. et al. Double-strand break repair-independent role for BRCA2 in blocking stalled replication fork degradation by MRE11. Cell 145, 529–542 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Schlacher, K., Wu, H. & Jasin, M. A distinct replication fork protection pathway connects Fanconi anemia tumor suppressors to RAD51-BRCA1/2. Cancer Cell 22, 106–116 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank D. Schad for help with the figures to this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas F. Meyer.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Bacterial secretion systems

Molecular machineries used to transport bacterial effector molecules to their environment or simultaneously across the host cell membrane.

NLRX1

(Nucleotide-binding oligomerization domain, leucine-rich repeat containing X1 protein). A member of the NOD-like receptor family that translocates to the mitochondria and has been shown to augment reactive oxygen species production from the mitochondria in response to infections.

Phagosome

Membrane-bound organelle that degrades proteins from internalized particles into peptides.

Nosocomial diseases

Diseases caused by hospital-acquired infections, frequently caused by antibiotic-resistant bacteria, posing a high risk to susceptible patients.

Exotoxins

Bacterial secreted toxins that can bind to a cell surface receptor and stimulate intracellular signalling pathways (type I), damage the cell membrane (type II) or exert effects intracellularly (type III), causing damage to the host by destroying cells or disrupting their normal metabolism.

Bacterial virulence factor

Chromosome- or plasmid-encoded molecules, including bacterial toxins that contribute to colonization or pathogenicity.

CagA

(Cytotoxin-associated gene A). A 120–145 kDa protein expressed by Cag pathogenicity island-positive Helicobacter pylori strains. CagA is translocated into host cells and is associated with increased inflammation and risk of gastric cancer.

Cag pathogenicity island

(CagPAI). A 40 kb DNA segment of Helicobacter pylori, encoding 20 genes, including CagA and proteins involved in its translocation to the host cell.

Somatic hypermutation

Mutation of variable immunoglobulin gene regions in immune cells induced by activation-induced cytidine deaminase (AID), which drives B cell diversification and immune adaptation.

Class switch recombination

Programmed removal of sections of the antibody heavy chain locus from B cell chromosomes through generation of double-strand breaks, followed by rejoining of the segments by non-homologous end-joining, leading to generation of a different antibody class.

Replication fork stalling

Pausing (sometimes indefinite) of the progress of the replication fork complex (where DNA unwinding and DNA synthesis take place), owing to the presence of impediments, such as DNA lesions.

pks genomic island

A 54-kb genomic island of Escherichia coli, which encodes several genes that are responsible for the synthesis of the genotoxin colibactin.

Gram-negative bacteria

Bacterial species that do not have a thick cell wall, as detected by the Gram stain for peptidoglycan.

p38 MAPK

A class of MAPKs that respond to stress stimuli and have a major role in apoptosis, differentiation, survival, proliferation, development and inflammation.

γH2AX

Ser139 phosphorylated histone H2AX, which serves as an early cellular response and a sensitive marker for DNA double-strand breaks.

Anaphase bridges

Chromatin bridges formed during anaphase, caused by fusion of telomeres of sister chromatids, leading to failure to completely segregate them into their respective daughter cells.

Spindle assembly checkpoint

A process ensuring accurate chromosome segregation into daughter cells, which acts by delaying cell division during mitosis and meiosis until proper attachment of chromosomes to the microtubule spindle apparatus is achieved.

Micronuclei

Small, extranuclear bodies, resulting from chromosome fragments or whole chromosomes being separated from daughter nuclei during mitosis, either by breaking of an anaphase bridge or by a double-strand break in the DNA.

Microsatellite instability

Hypermutability of short tandem repeat DNA sequences, which results from defects in mismatch repair.

Senescence-associated secretory phenotype

(SASP). Expression of a set of secreted growth factors, inflammatory cytokines, proteases and matrix components by senescent cells, which is associated with pro-malignant changes in surrounding cells.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chumduri, C., Gurumurthy, R., Zietlow, R. et al. Subversion of host genome integrity by bacterial pathogens. Nat Rev Mol Cell Biol 17, 659–673 (2016). https://doi.org/10.1038/nrm.2016.100

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrm.2016.100

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