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:

Persistent bacterial infections and persister cells

Key Points

  • Many bacterial infections persist in the host for long periods of time despite antibiotic treatment.

  • This persistence is an important medical concern, as it leads to the overuse of antibiotics and therefore contributes to antimicrobial resistance.

  • The role of antibiotic-tolerant persister cells in the recalcitrance and relapse of bacterial infections has gained recognition in recent years.

  • Persisters are often growth-arrested bacteria with a reduced metabolism that are able to restart growth after a stress.

  • The stresses that bacteria encounter during the infection of a host are triggers for the formation of persisters.

  • Toxin–antitoxin modules have an important role in the formation of growth-arrested persisters.

  • Understanding how growth-arrested persisters regrow is necessary to design better therapeutic strategies.

Abstract

Many bacteria can infect and persist inside their hosts for long periods of time. This can be due to immunosuppression of the host, immune evasion by the pathogen and/or ineffective killing by antibiotics. Bacteria can survive antibiotic treatment if they are resistant or tolerant to a drug. Persisters are a subpopulation of transiently antibiotic-tolerant bacterial cells that are often slow-growing or growth-arrested, and are able to resume growth after a lethal stress. The formation of persister cells establishes phenotypic heterogeneity within a bacterial population and has been hypothesized to be important for increasing the chances of successfully adapting to environmental change. The presence of persister cells can result in the recalcitrance and relapse of persistent bacterial infections, and it has been linked to an increase in the risk of the emergence of antibiotic resistance during treatment. If the mechanisms of the formation and regrowth of these antibiotic-tolerant cells were better understood, it could lead to the development of new approaches for the eradication of persistent bacterial infections. In this Review, we discuss recent developments in our understanding of bacterial persisters and their potential implications for the treatment of persistent infections.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: Antibiotic killing kinetics of resistant, tolerant and persister cells.
Figure 2: Mechanisms of persister formation.
Figure 3: Mechanisms of persister regrowth.

Similar content being viewed by others

References

  1. Helaine, S. & Kugelberg, E. Bacterial persisters: formation, eradication, and experimental systems. Trends Microbiol. 22, 417–424 (2014).

    CAS  PubMed  Google Scholar 

  2. Harms, A., Maisonneuve, E. & Gerdes, K. Mechanisms of bacterial persistence during stress and antibiotic exposure. Science 354, aaf4268 (2016).

    PubMed  Google Scholar 

  3. Michiels, J. E., Van den Bergh, B., Verstraeten, N. & Michiels, J. Molecular mechanisms and clinical implications of bacterial persistence. Drug Resist. Updat. 29, 76–89 (2017).

    Google Scholar 

  4. Mechler, L. et al. A novel point mutation promotes growth phase-dependent daptomycin tolerance in Staphylococcus aureus. Antimicrob. Agents Chemother. 59, 5366–5376 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Van den Bergh, B. et al. Frequency of antibiotic application drives rapid evolutionary adaptation of Escherichia coli persistence. Nat. Microbiol. 1, 16020 (2016).

    CAS  PubMed  Google Scholar 

  6. Cohen, N. R., Lobritz, M. A. & Collins, J. J. Microbial persistence and the road to drug resistance. Cell Host Microbe 13, 632–642 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Levin-Reisman, I. et al. Antibiotic tolerance facilitates the evolution of resistance. Science 355, 826–830 (2017).

    CAS  PubMed  Google Scholar 

  8. Hobby, G. L., Meyer, K. & Chaffee, E. Observations on the mechanism of action of penicillin. Exp. Biol. Med. 50, 281–285 (1942).

    CAS  Google Scholar 

  9. Bigger, J. W. Treatment of staphylococcal infections with penicillin by intermittant sterilisation. Lancet 244, 497–500 (1944).

    Google Scholar 

  10. Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. & Leibler, S. Bacterial persistence as a phenotypic switch. Science 305, 1622–1625 (2004).

    CAS  PubMed  Google Scholar 

  11. Adams, K. N. et al. Drug tolerance in replicating mycobacteria mediated by a macrophage-induced efflux mechanism. Cell 145, 39–53 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Wakamoto, Y. et al. Dynamic persistence of antibiotic-stressed mycobacteria. Science 339, 91–95 (2013).

    CAS  PubMed  Google Scholar 

  13. Amato, S. et al. The role of metabolism in bacterial persistence. Front. Microbiol. 5, 70 (2014).

    PubMed  PubMed Central  Google Scholar 

  14. Helaine, S. et al. Internalization of Salmonella by macrophages induces formation of nonreplicating persisters. Science 343, 204–208 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Rhen, M., Eriksson, S., Clements, M., Bergström, S. & Normark, S. J. The basis of persistent bacterial infections. Trends Microbiol. 11, 80–86 (2003).

    CAS  PubMed  Google Scholar 

  16. Grant, S. S. & Hung, D. T. Persistent bacterial infections, antibiotic tolerance, and the oxidative stress response. Virulence 4, 273–283 (2013).

    PubMed  PubMed Central  Google Scholar 

  17. Levine, M. M., Black, R. E. & Lanata, C. Precise estimation of the numbers of chronic carriers of Salmonella Typhi in Santiago, Chile, an endemic area. J. Infect. Dis. 146, 724–726 (1982).

    CAS  PubMed  Google Scholar 

  18. Caygill, C. P. J., Hill, M. J., Braddick, M. & Sharp, J. C. M. Cancer mortality in chronic typhoid and paratyphoid carriers. Lancet 343, 83–84 (1994).

    CAS  PubMed  Google Scholar 

  19. Bhan, M. K., Bahl, R. & Bhatnagar, S. Typhoid and paratyphoid fever. Lancet 366, 749–762 (2005).

    CAS  PubMed  Google Scholar 

  20. Nomura, A., Stemmermann, G. N., Chyou, P.-H., Perez-Perez, G. I. & Blaser, M. J. Helicobacter pylori infection and the risk for duodenal and gastric ulceration. Ann. Intern. Med. 120, 977–981 (1994).

    CAS  PubMed  Google Scholar 

  21. Wotherspoon, A. C., Ortiz-Hidalgo, C., Falzon, M. R. & Isaacson, P. G. Helicobacter pylori-associated gastritis and primary B-cell gastric lymphoma. Lancet 338, 1175–1176 (1991).

    CAS  PubMed  Google Scholar 

  22. Eslick, G. D., Lim, L. L.-Y., Byles, J. E., Xia, H. H.-X. & Talley, N. J. Association of Helicobacter pylori infection with gastric carcinoma: a meta-analysis. Am. J. Gastroenterol. 94, 2373–2379 (1999).

    CAS  PubMed  Google Scholar 

  23. Gomez, J. E. & McKinney, J. D. M. tuberculosis persistence, latency, and drug tolerance. Tuberculosis 84, 29–44 (2004).

    PubMed  Google Scholar 

  24. Gotuzzo, E. et al. Association between specific plasmids and relapse in typhoid fever. J. Clin. Microbiol. 25, 1779–1781 (1987).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Foxman, B. Recurring urinary tract infection: incidence and risk factors. Am. J. Public Health 80, 331–333 (1990).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Österlund, A., Popa, R., Nikkilä, T., Scheynius, A. & Engstrand, L. Intracellular reservoir of Streptococcus pyogenes in vivo: a possible explanation for recurrent pharyngotonsillitis. Laryngoscope 107, 640–647 (1997).

    PubMed  Google Scholar 

  27. Bryers, J. D. Medical biofilms. Biotechnol. Bioeng. 100, 1–18 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Lyczak, J. B., Cannon, C. L. & Pier, G. B. Lung infections associated with cystic fibrosis. Clin. Microbiol. Rev. 15, 194–222 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Jernberg, C., Löfmark, S., Edlund, C. & Jansson, J. K. Long-term impacts of antibiotic exposure on the human intestinal microbiota. Microbiology 156, 3216–3223 (2010).

    CAS  PubMed  Google Scholar 

  30. Norris, S. J. Antigenic variation with a twist — the Borrelia story. Mol. Microbiol. 60, 1319–1322 (2006).

    CAS  PubMed  Google Scholar 

  31. Redpath, S., Ghazal, P. & Gascoigne, N. R. J. Hijacking and exploitation of IL-10 by intracellular pathogens. Trends Microbiol. 9, 86–92 (2001).

    CAS  PubMed  Google Scholar 

  32. Scott, C. C., Botelho, R. J. & Grinstein, S. Phagosome maturation: a few bugs in the system. J. Membr. Biol. 193, 137–152 (2003).

    CAS  PubMed  Google Scholar 

  33. Bayer-Santos, E. et al. The Salmonella effector SteD mediates MARCH8-dependent ubiquitination of MHC II molecules and inhibits T cell activation. Cell Host Microbe 20, 584–595 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Reddick, L. E. & Alto, N. M. Bacteria fighting back — how pathogens target and subvert the host innate immune system. Mol. Cell 54, 321–328 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Stewart, P. S. Mechanisms of antibiotic resistance in bacterial biofilms. Int. J. Med. Microbiol. 292, 107–113 (2002).

    CAS  PubMed  Google Scholar 

  36. Monack, D. M., Mueller, A. & Falkow, S. Persistent bacterial infections: the interface of the pathogen and the host immune system. Nat. Rev. Microbiol. 2, 747–765 (2004).

    CAS  PubMed  Google Scholar 

  37. Salama, N. R., Hartung, M. L. & Muller, A. Life in the human stomach: persistence strategies of the bacterial pathogen Helicobacter pylori. Nat. Rev. Microbiol. 11, 385–399 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Costerton, J. W., Stewart, P. S. & Greenberg, E. P. Bacterial biofilms: a common cause of persistent infections. Science 284, 1318–1322 (1999).

    CAS  PubMed  Google Scholar 

  39. Jesaitis, A. J. et al. Compromised host defense on Pseudomonas aeruginosa biofilms: characterization of neutrophil and biofilm interactions. J. Immunol. 171, 4329–4339 (2003).

    CAS  PubMed  Google Scholar 

  40. Domenech, M., Ramos-Sevillano, E., García, E., Moscoso, M. & Yuste, J. Biofilm formation avoids complement immunity and phagocytosis of Streptococcus pneumoniae. Infect. Immun. 81, 2606–2615 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Lawn, S. D., Butera, S. T. & Shinnick, T. M. Tuberculosis unleashed: the impact of human immunodeficiency virus infection on the host granulomatous response to Mycobacterium tuberculosis. Microbes Infect. 4, 635–646 (2002).

    CAS  PubMed  Google Scholar 

  42. Kardas, P. Patient compliance with antibiotic treatment for respiratory tract infections. J. Antimicrob. Chemother. 49, 897–903 (2002).

    CAS  PubMed  Google Scholar 

  43. Gonçalves-Pereira, J. & Póvoa, P. Antibiotics in critically ill patients: a systematic review of the pharmacokinetics of β-lactams. Crit. Care 15, R206 (2011).

    PubMed  PubMed Central  Google Scholar 

  44. Blair, J. M. A., Webber, M. A., Baylay, A. J., Ogbolu, D. O. & Piddock, L. J. V. Molecular mechanisms of antibiotic resistance. Nat. Rev. Microbiol. 13, 42–51 (2015).

    CAS  PubMed  Google Scholar 

  45. Brauner, A., Fridman, O., Gefen, O. & Balaban, N. Q. Distinguishing between resistance, tolerance and persistence to antibiotic treatment. Nat. Rev. Microbiol. 14, 320–330 (2016). This article explains the differences between antibiotic resistance, tolerance and persistence, and proposes a quantitative indicator of tolerance (the minimum duration for killing (MDK)) for use in the clinic.

    CAS  PubMed  Google Scholar 

  46. Lewis, K. Persister cells. Annu. Rev. Microbiol. 64, 357–372 (2010).

    CAS  PubMed  Google Scholar 

  47. Claudi, B. et al. Phenotypic variation of Salmonella in host tissues delays eradication by antimicrobial chemotherapy. Cell 158, 722–733 (2014).

    CAS  PubMed  Google Scholar 

  48. Okoro, C. K. et al. High-resolution single nucleotide polymorphism analysis distinguishes recrudescence and reinfection in recurrent invasive nontyphoidal Salmonella typhimurium disease. Clin. Infect. Dis. 54, 955–963 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Russo, T. A., Stapleton, A., Wenderoth, S., Hooton, T. M. & Stamm, W. E. Chromosomal restriction fragment length polymorphism analysis of Escherichia coli strains causing recurrent urinary tract infections in young women. J. Infect. Dis. 172, 440–445 (1995).

    CAS  PubMed  Google Scholar 

  50. Bingen, E. et al. DNA restriction fragment length polymorphism differentiates recurrence from relapse in treatment failures of Streptococcus pyogenes pharyngitis. J. Med. Microbiol. 37, 162–164 (1992).

    CAS  PubMed  Google Scholar 

  51. Fridman, O., Goldberg, A., Ronin, I., Shoresh, N. & Balaban, N. Q. Optimization of lag time underlies antibiotic tolerance in evolved bacterial populations. Nature 513, 418–421 (2014). This evolutionary study uses the recently developed ScanLag technique to monitor bacterial regrowth after repeated antibiotic treatments and uncovers the importance of lag in the tolerance of bacterial populations.

    CAS  PubMed  Google Scholar 

  52. Moyed, H. S. & Bertrand, K. P. hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J. Bacteriol. 155, 768–775 (1983).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Mulcahy, L. R., Burns, J. L., Lory, S. & Lewis, K. Emergence of Pseudomonas aeruginosa strains producing high levels of persister cells in patients with cystic fibrosis. J. Bacteriol. 192, 6191–6199 (2010). This study uncovers a strong link between the formation of persisters and persistent bacterial infections.

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Schumacher, M. A. et al. HipBA–promoter structures reveal the basis of heritable multidrug tolerance. Nature 524, 59–64 (2015). This study identifies how a mutation that was found in an E. coli UTI clinical isolate affects dimerization of the toxin HipA and, consequently, disrupts toxin–antitoxin–promoter interactions, which leads to an increase in toxicity and antibiotic tolerance.

    CAS  PubMed  PubMed Central  Google Scholar 

  55. LaFleur, M. D., Qi, Q. & Lewis, K. Patients with long-term oral carriage harbor high-persister mutants of Candida albicans. Antimicrob. Agents Chemother. 54, 39–44 (2010).

    CAS  PubMed  Google Scholar 

  56. Mouton, J. M., Helaine, S., Holden, D. W. & Sampson, S. L. Elucidating population-wide mycobacterial replication dynamics at the single-cell level. Microbiology 162, 966–978 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Brooun, A., Liu, S. & Lewis, K. A. Dose–response study of antibiotic resistance in Pseudomonas aeruginosa biofilms. Antimicrob. Agents Chemother. 44, 640–646 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  58. Spoering, A. L. & Lewis, K. Biofilms and planktonic cells of Pseudomonas aeruginosa have similar resistance to killing by antimicrobials. J. Bacteriol. 183, 6746–6751 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Keren, I., Kaldalu, N., Spoering, A., Wang, Y. & Lewis, K. Persister cells and tolerance to antimicrobials. FEMS Microbiol. Lett. 230, 13–18 (2004).

    CAS  PubMed  Google Scholar 

  60. Maisonneuve, E., Castro-Camargo, M. & Gerdes, K. (p)ppGpp controls bacterial persistence by stochastic induction of toxin–antitoxin activity. Cell 154, 1140–1150 (2013). This study identifies a signalling pathway for the activation of toxin–antitoxin modules, leading to the formation of persisters.

    CAS  PubMed  Google Scholar 

  61. Conlon, B. P. et al. Activated ClpP kills persisters and eradicates a chronic biofilm infection. Nature 503, 365–370 (2013). This study uses ADEP4 as a treatment to eradicate a chronic biofilm infection, which represents a great advance in the development of anti-persister therapies.

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Burmølle, M. et al. Biofilms in chronic infections — a matter of opportunity — monospecies biofilms in multispecies infections. FEMS Immunol. Med. Microbiol. 59, 324–336 (2010).

    PubMed  Google Scholar 

  63. Hunstad, D. A. & Justice, S. S. Intracellular lifestyles and immune evasion strategies of uropathogenic Escherichia coli. Annu. Rev. Microbiol. 64, 203–221 (2010).

    CAS  PubMed  Google Scholar 

  64. Davies, J. C. Pseudomonas aeruginosa in cystic fibrosis: pathogenesis and persistence. Paediatr. Respir. Rev. 3, 128–134 (2002).

    PubMed  Google Scholar 

  65. Oh, J. D., Karam, S. M. & Gordon, J. I. Intracellular Helicobacter pylori in gastric epithelial progenitors. Proc. Natl Acad. Sci. USA 102, 5186–5191 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Cammarota, G., Sanguinetti, M., Gallo, A. & Posteraro, B. Review article: biofilm formation by Helicobacter pylori as a target for eradication of resistant infection. Aliment. Pharmacol. Ther. 36, 222–230 (2012).

    CAS  PubMed  Google Scholar 

  67. Prouty, A. M., Schwesinger, W. H. & Gunn, J. S. Biofilm formation and interaction with the surfaces of gallstones by Salmonella spp. Infect. Immun. 70, 2640–2649 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  68. Amato, S. M., Orman, M. A. & Brynildsen, M. P. Metabolic control of persister formation in Escherichia coli. Mol. Cell 50, 475–487 (2013).

    CAS  PubMed  Google Scholar 

  69. Dörr, T., Vulic´, M. & Lewis, K. Ciprofloxacin causes persister formation by inducing the TisB toxin in Escherichia coli. PLoS Biol. 8, e1000317 (2010).

    PubMed  PubMed Central  Google Scholar 

  70. Fasani, R. A. & Savageau, M. A. Molecular mechanisms of multiple toxin–antitoxin systems are coordinated to govern the persister phenotype. Proc. Natl Acad. Sci. USA 110, E2528–E2537 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Nguyen, D. et al. Active starvation responses mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. Science 334, 982–986 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  72. Ogura, T. & Hiraga, S. Mini-F plasmid genes that couple host cell division to plasmid proliferation. Proc. Natl Acad. Sci. USA 80, 4784–4788 (1983).

    CAS  PubMed  PubMed Central  Google Scholar 

  73. Gerdes, K., Rasmussen, P. B. & Molin, S. Unique type of plasmid maintenance function: postsegregational killing of plasmid-free cells. Proc. Natl Acad. Sci. USA 83, 3116–3120 (1986).

    CAS  PubMed  PubMed Central  Google Scholar 

  74. Pandey, D. P. & Gerdes, K. Toxin–antitoxin loci are highly abundant in free-living but lost from host-associated prokaryotes. Nucleic Acids Res. 33, 966–976 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  75. Gerdes, K., Christensen, S. K. & Lobner-Olesen, A. Prokaryotic toxin–antitoxin stress response loci. Nat. Rev. Microbiol. 3, 371–382 (2005).

    CAS  PubMed  Google Scholar 

  76. Yamaguchi, Y. & Inouye, M. Regulation of growth and death in Escherichia coli by toxin–antitoxin systems. Nat. Rev. Microbiol. 9, 779–790 (2011).

    CAS  PubMed  Google Scholar 

  77. Rotem, E. et al. Regulation of phenotypic variability by a threshold-based mechanism underlies bacterial persistence. Proc. Natl Acad. Sci. USA 107, 12541–12546 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  78. Cataudella, I., Trusina, A., Sneppen, K., Gerdes, K. & Mitarai, N. Conditional cooperativity in toxin–antitoxin regulation prevents random toxin activation and promotes fast translational recovery. Nucleic Acids Res. 40, 6424–6434 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  79. Page, R. & Peti, W. Toxin–antitoxin systems in bacterial growth arrest and persistence. Nat. Chem. Biol. 12, 208–214 (2016). A comprehensive review of toxin–antitoxin systems in the context of persister formation.

    CAS  PubMed  Google Scholar 

  80. Keren, I., Shah, D., Spoering, A., Kaldalu, N. & Lewis, K. Specialized persister cells and the mechanism of multidrug tolerance in Escherichia coli. J. Bacteriol. 186, 8172–8180 (2004).

    CAS  PubMed  PubMed Central  Google Scholar 

  81. Shah, D. et al. Persisters: a distinct physiological state of E. coli. BMC Microbiol. 6, 53 (2006).

    PubMed  PubMed Central  Google Scholar 

  82. Maisonneuve, E., Shakespeare, L. J., Jørgensen, M. G. & Gerdes, K. Bacterial persistence by RNA endonucleases. Proc. Natl Acad. Sci. USA 108, 13206–13211 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  83. Verstraeten, N. et al. Obg and membrane depolarization are part of a microbial bet-hedging strategy that leads to antibiotic tolerance. Mol. Cell 59, 9–21 (2015). This study describes the regulation of the HokB–SokB toxin–antitoxin system by ObgE and (p)ppGpp in E. coli , and suggests an overlap in the regulation of both type I and type II toxin–antitoxin modules and central metabolism.

    CAS  PubMed  Google Scholar 

  84. Maurizi, M. R. Proteases and protein degradation in Escherichia coli. Experientia 48, 178–201 (1992).

    CAS  PubMed  Google Scholar 

  85. Kint, C., Verstraeten, N., Hofkens, J., Fauvart, M. & Michiels, J. Bacterial Obg proteins: GTPases at the nexus of protein and DNA synthesis. Crit. Rev. Microbiol. 40, 207–224 (2014).

    CAS  PubMed  Google Scholar 

  86. Theodore, A., Lewis, K. & Vulic´, M. Tolerance of Escherichia coli to fluoroquinolone antibiotics depends on specific components of the SOS response pathway. Genetics 195, 1265–1276 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  87. Shan, Y., Lazinski, D., Rowe, S., Camilli, A. & Lewis, K. Genetic basis of persister tolerance to aminoglycosides in Escherichia coli. mBio 6, e00078-15 (2015).

    PubMed  PubMed Central  Google Scholar 

  88. Mizusawa, S. & Gottesman, S. Protein degradation in Escherichia coli: the lon gene controls the stability of SulA protein. Proc. Natl Acad. Sci. USA 80, 358–362 (1983).

    CAS  PubMed  PubMed Central  Google Scholar 

  89. Germain, E., Castro-Roa, D., Zenkin, N. & Gerdes, K. Molecular mechanism of bacterial persistence by HipA. Mol. Cell 52, 248–254 (2013).

    CAS  PubMed  Google Scholar 

  90. Pedersen, K. et al. The bacterial toxin RelE displays codon-specific cleavage of mRNAs in the ribosomal A site. Cell 112, 131–140 (2003).

    CAS  PubMed  Google Scholar 

  91. Christensen, S. K. & Gerdes, K. RelE toxins from bacteria and Archaea cleave mRNAs on translating ribosomes, which are rescued by tmRNA. Mol. Microbiol. 48, 1389–1400 (2003).

    CAS  PubMed  Google Scholar 

  92. Christensen-Dalsgaard, M., Jørgensen, M. G. & Gerdes, K. Three new RelE–homologous mRNA interferases of Escherichia coli differentially induced by environmental stresses. Mol. Microbiol. 75, 333–348 (2010).

    CAS  PubMed  Google Scholar 

  93. Winther, K. S. & Gerdes, K. Enteric virulence associated protein VapC inhibits translation by cleavage of initiator tRNA. Proc. Natl Acad. Sci. USA 108, 7403–7407 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  94. Winther, K., Tree, J. J., Tollervey, D. & Gerdes, K. VapCs of Mycobacterium tuberculosis cleave RNAs essential for translation. Nucleic Acids Res. 9860–9871 (2016).

  95. Castro-Roa, D. et al. The Fic protein Doc uses an inverted substrate to phosphorylate and inactivate EF-Tu. Nat. Chem. Biol. 9, 811–817 (2013).

    CAS  PubMed  Google Scholar 

  96. Bernard, P. et al. The F plasmid CcdB protein induces efficient ATP-dependent DNA cleavage by gyrase. J. Mol. Biol. 234, 534–541 (1993).

    CAS  PubMed  Google Scholar 

  97. Jiang, Y., Pogliano, J., Helinski, D. R. & Konieczny, I. ParE toxin encoded by the broad-host-range plasmid RK2 is an inhibitor of Escherichia coli gyrase. Mol. Microbiol. 44, 971–979 (2002).

    CAS  PubMed  Google Scholar 

  98. Harms, A. et al. Adenylylation of gyrase and Topo IV by FicT toxins disrupts bacterial DNA topology. Cell Rep. 12, 1497–1507 (2015).

    CAS  PubMed  Google Scholar 

  99. Cheverton, A. M. et al. A Salmonella toxin promotes persister formation through acetylation of tRNA. Mol. Cell 63, 86–96 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  100. Pedersen, K., Christensen, S. K. & Gerdes, K. Rapid induction and reversal of a bacteriostatic condition by controlled expression of toxins and antitoxins. Mol. Microbiol. 45, 501–510 (2002).

    CAS  PubMed  Google Scholar 

  101. Harrison, J. J. et al. The chromosomal toxin gene yafQ is a determinant of multidrug tolerance for Escherichia coli growing in a biofilm. Antimicrob. Agents Chemother. 53, 2253–2258 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  102. Lobato-Márquez, D., Moreno-Córdoba, I., Figueroa, V., Díaz-Orejas, R. & García-del Portillo, F. Distinct type I and type II toxin–antitoxin modules control Salmonella lifestyle inside eukaryotic cells. Sci. Rep. 5, 9374 (2015).

    PubMed  PubMed Central  Google Scholar 

  103. Kaspy, I. et al. HipA-mediated antibiotic persistence via phosphorylation of the glutamyl-tRNA-synthetase. Nat. Commun. 4, 3001 (2013).

    PubMed  Google Scholar 

  104. Germain, E., Roghanian, M., Gerdes, K. & Maisonneuve, E. Stochastic induction of persister cells by HipA through (p)ppGpp-mediated activation of mRNA endonucleases. Proc. Natl Acad. Sci. USA 112, 5171–5176 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  105. Conlon, B. P. et al. Persister formation in Staphylococcus aureus is associated with ATP depletion. Nat. Microbiol. 1, 16051 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  106. Fu, Z., Tamber, S., Memmi, G., Donegan, N. P. & Cheung, A. L. Overexpression of MazFSa in Staphylococcus aureus induces bacteriostasis by selectively targeting mRNAs for cleavage. J. Bacteriol. 191, 2051–2059 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  107. Donegan, N. P. & Cheung, A. L. Regulation of the mazEF toxin–antitoxin module in Staphylococcus aureus and its impact on sigB expression. J. Bacteriol. 191, 2795–2805 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  108. Chowdhury, N., Kwan, B. W. & Wood, T. K. Persistence increases in the absence of the alarmone guanosine tetraphosphate by reducing cell growth. Sci. Rep. 6, 20519 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  109. Baba, T. et al. Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Mol. Syst. Biol. http://dx.doi.org/10.1038/msb4100050 (2006).

  110. Hansen, S., Lewis, K. & Vulic´, M. Role of global regulators and nucleotide metabolism in antibiotic tolerance in Escherichia coli. Antimicrob. Agents Chemother. 52, 2718–2726 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  111. Radzikowski, J. L., Schramke, H. & Heinemann, M. Bacterial persistence from a system-level perspective. Curr. Opin. Biotechnol. 46, 98–105 (2017).

    CAS  PubMed  Google Scholar 

  112. Radzikowski, J. L. et al. Bacterial persistence is an active σS stress response to metabolic flux limitation. Mol. Syst. Biol. 12, 882 (2016).

    PubMed  PubMed Central  Google Scholar 

  113. Amato, S. M. & Brynildsen, M. P. Persister heterogeneity arising from a single metabolic stress. Curr. Biol. 25, 2090–2098 (2015).

    CAS  PubMed  Google Scholar 

  114. Shan, Y. et al. ATP-dependent persister formation in Escherichia coli. mBio 8, e02267-16 (2017).

    PubMed  PubMed Central  Google Scholar 

  115. Pu, Y. et al. Enhanced efflux activity facilitates drug tolerance in dormant bacterial cells. Mol. Cell 62, 284–294 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  116. Band, V. I. et al. Antibiotic failure mediated by a resistant subpopulation in Enterobacter cloacae. Nat. Microbiol. 1, 16053 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Pader, V. et al. Staphylococcus aureus inactivates daptomycin by releasing membrane phospholipids. Nat. Microbiol. 2, 16194 (2016).

    CAS  PubMed  Google Scholar 

  118. Queck, S. Y. et al. RNAIII-independent target gene control by the agr quorum-sensing system: insight into the evolution of virulence regulation in Staphylococcus aureus. Mol. Cell 32, 150–158 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  119. Painter, K. L., Krishna, A., Wigneshweraraj, S. & Edwards, A. M. What role does the quorum-sensing accessory gene regulator system play during Staphylococcus aureus bacteremia? Trends Microbiol. 22, 676–685 (2014).

    CAS  PubMed  Google Scholar 

  120. Overgaard, M., Borch, J., Jørgensen, M. G. & Gerdes, K. Messenger RNA interferase RelE controls relBE transcription by conditional cooperativity. Mol. Microbiol. 69, 841–857 (2008).

    CAS  PubMed  Google Scholar 

  121. Hall, A. M. J., Gollan, B. & Helaine, S. Toxin–antitoxin systems: reversible toxicity. Curr. Opin. Microbiol. 36, 102–110 (2017).

    CAS  PubMed  Google Scholar 

  122. Christensen, S. K., Pedersen, K., Hansen, F. G. & Gerdes, K. Toxin–antitoxin loci as stress-response-elements: ChpAK/MazF and ChpBK cleave translated RNAs and are counteracted by tmRNA. J. Mol. Biol. 332, 809–819 (2003).

    CAS  PubMed  Google Scholar 

  123. Allison, K. R., Brynildsen, M. P. & Collins, J. J. Metabolite-enabled eradication of bacterial persisters by aminoglycosides. Nature 473, 216–220 (2011). This was the first study to develop a method of persister eradication and test it in an animal model.

    CAS  PubMed  PubMed Central  Google Scholar 

  124. Pan, J., Bahar, A. A., Syed, H. & Ren, D. Reverting antibiotic tolerance of Pseudomonas aeruginosa PAO1 persister cells by (Z)-4-bromo-5-(bromomethylene)-3-methylfuran-2(5H)-one. PLoS ONE 7, e45778 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Harris, J. & Keane, J. How tumour necrosis factor blockers interfere with tuberculosis immunity. Clin. Exp. Immunol. 161, 1–9 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  126. Belland, R. J. et al. Transcriptome analysis of chlamydial growth during IFNγ-mediated persistence and reactivation. Proc. Natl Acad. Sci. USA 100, 15971–15976 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  127. Layton, J. C. & Foster, P. L. Error-prone DNA polymerase IV is controlled by the stress-response sigma factor, RpoS, in Escherichia coli. Mol. Microbiol. 50, 549–561 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Goodman, M. F. Error-prone repair DNA polymerases in prokaryotes and eukaryotes. Annu. Rev. Biochem. 71, 17–50 (2002).

    CAS  PubMed  Google Scholar 

  129. Dörr, T., Lewis, K. & Vulic´, M. SOS response induces persistence to fluoroquinolones in Escherichia coli. PLOS Genet. 5, e1000760 (2009).

    PubMed  PubMed Central  Google Scholar 

  130. Beaber, J. W., Hochhut, B. & Waldor, M. K. SOS response promotes horizontal dissemination of antibiotic resistance genes. Nature 427, 72–74 (2004).

    CAS  PubMed  Google Scholar 

  131. Völzing, K. G. & Brynildsen, M. P. Stationary-phase persisters to ofloxacin sustain DNA damage and require repair systems only during recovery. mBio 6, e00731-15 (2015).

    PubMed  PubMed Central  Google Scholar 

  132. Pearl Mizrahi, S., Gefen, O., Simon, I. & Balaban, N. Q. Persistence to anti-cancer treatments in the stationary to proliferating transition. Cell Cycle 15, 3442–3453 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  133. Michiels, J. & Fauvart, M. (eds) Bacterial Persistence: Methods and Protocols (Springer New York, 2016).

    Google Scholar 

  134. Helaine, S. et al. Dynamics of intracellular bacterial replication at the single cell level. Proc. Natl Acad. Sci. USA 107, 3746–3751 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  135. Manina, G., Dhar, N. & McKinney, J. D. Stress and host immunity amplify Mycobacterium tuberculosis phenotypic heterogeneity and induce nongrowing metabolically active forms. Cell Host Microbe 17, 32–46 (2015).

    CAS  PubMed  Google Scholar 

  136. Wang, Z., Gerstein, M. & Snyder, M. RNA-Seq: a revolutionary tool for transcriptomics. Nat. Rev. Genet. 10, 57–63 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  137. Levin-Reisman, I. et al. Automated imaging with ScanLag reveals previously undetectable bacterial growth phenotypes. Nat. Methods 7, 737–739 (2010).

    CAS  PubMed  Google Scholar 

  138. Gurnev, P. A., Ortenberg, R., Dörr, T., Lewis, K. & Bezrukov, S. M. Persister-promoting bacterial toxin TisB produces anion-selective pores in planar lipid bilayers. FEBS Lett. 586, 2529–2534 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  139. Hurdle, J. G., O'Neill, A. J., Chopra, I. & Lee, R. E. Targeting bacterial membrane function: an underexploited mechanism for treating persistent infections. Nat. Rev. Microbiol. 9, 62–75 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  140. Chen, X., Zhang, M., Zhou, C., Kallenbach, N. R. & Ren, D. Control of bacterial persister cells by Trp/Arg-containing antimicrobial peptides. Appl. Environ. Microbiol. 77, 4878–4885 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  141. Mukherjee, D., Zou, H., Liu, S., Beuerman, R. & Dick, T. Membrane-targeting AM-0016 kills mycobacterial persisters and shows low propensity for resistance development. Future Microbiol. 11, 643–650 (2016).

    CAS  PubMed  Google Scholar 

  142. Kim, W. et al. NH125 kills methicillin-resistant Staphylococcus aureus persisters by lipid bilayer disruption. Future Med. Chem. 8, 257–269 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  143. Schmidt, N. W. et al. Engineering persister-specific antibiotics with synergistic antimicrobial functions. ACS Nano 8, 8786–8793 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  144. Lehar, S. M. et al. Novel antibody–antibiotic conjugate eliminates intracellular S. aureus. Nature 527, 323–328 (2015).

    CAS  PubMed  Google Scholar 

  145. Yang, H., Yu, J. & Wei, H. Engineered bacteriophage lysins as novel anti-infectives. Front. Microbiol. 5, 542 (2014).

    PubMed  PubMed Central  Google Scholar 

  146. Briers, Y. et al. Art-175 is a highly efficient antibacterial against multidrug-resistant strains and persisters of Pseudomonas aeruginosa. Antimicrob. Agents Chemother. 58, 3774–3784 (2014).

    PubMed  PubMed Central  Google Scholar 

  147. De Soyza, A. et al. Lung transplantation for patients with cystic fibrosis and Burkholderia cepacia complex infection: a single-center experience. J. Heart Lung Transplant. 29, 1395–1404 (2010).

    PubMed  Google Scholar 

  148. King, P. Haemophilus influenzae and the lung (Haemophilus and the lung). Clin. Transl Med. 1, 10 (2012).

    PubMed  PubMed Central  Google Scholar 

  149. Ficht, T. A. Intracellular survival of Brucella: defining the link with persistence. Vet. Microbiol. 92, 213–223 (2003).

    CAS  PubMed  Google Scholar 

  150. Pawlowski, S. W., Warren, C. A. & Guerrant, R. Diagnosis and treatment of acute or persistent diarrhea. Gastroenterology 136, 1874–1886 (2009).

    CAS  PubMed  Google Scholar 

  151. Cocanour, C. S. Best strategies in recurrent or persistent Clostridium difficile infection. Surg. Infect. (Larchmt) 12, 235–239 (2011).

    Google Scholar 

  152. Marzel, A. et al. Persistent infections by nontyphoidal Salmonella in humans: epidemiology and genetics. Clin. Infect. Dis. 62, 879–886 (2016).

    PubMed  PubMed Central  Google Scholar 

  153. Chong, Y. P. et al. Persistent Staphylococcus aureus bacteremia: a prospective analysis of risk factors, outcomes, and microbiologic and genotypic characteristics of isolates. Medicine (Baltimore) 92, 98–108 (2013).

    CAS  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  155. Radolf, J. D. et al. Treponema pallidum, the syphilis spirochete: making a living as a stealth pathogen. Nat. Rev. Microbiol. 14, 744–759 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Helaine laboratory for critical reading of the manuscript. R.A.F. is supported by a UK Medical Research Council (MRC) Centre for Molecular Bacteriology and Infection (CMBI) studentship (grant MR/J006874/1). S.H. and B.G. are supported by an MRC Career Development Award (grant MR/M009629/1).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sophie Helaine.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

PowerPoint slides

Glossary

Nosocomial infections

Infections acquired when under medical care, also known as hospital-acquired infections (HAIs).

Interleukin-10

(IL-10). A cytokine that suppresses the interferon (IFN)-mediated transcriptional response.

Interferon-γ response

(IFNγ response). The transcriptional changes that take place in a cell due to the detection of the cytokine interferon-γ (IFNγ), some of these changes increase the antibacterial capabilities of phagocytes.

Type III secretion system

(T3SS). A protein needle-like appendage that is used by bacteria to translocate effector proteins into eukaryotic cells to manipulate host cellular processes.

MHC class II antigen presentation

The process of presenting antigen bound to MHC class II on the cell plasma membrane for recognition by cells of the adaptive immune system.

Granulomas

Organized groupings of macrophages and other cell types around foreign material that the immune system is unable to eliminate.

Bactericidal antibiotic

An antibiotic that kills bacteria, as opposed to a bacteriostatic antibiotic, which only inhibits growth.

Salmonella-containing vacuole

(SCV). A membrane-bound compartment that resembles that of a modified late endosome, in which Salmonella spp. typically reside within host cells.

Fluoroquinolones

A class of antibiotics that target DNA gyrase activity to induce the formation of lethal double-stranded breaks in bacterial DNA.

Genovar

A classification used for differentiating between strains of the same serovar that differ substantially in their genetic content.

Lag time

The time taken before resumption of the growth of growth-arrested bacteria.

Diauxic shift

A shift in metabolism from one carbon source to another.

Stochastic gene expression

The random (or noisy) fluctuations in the transcription of a particular gene.

Stringent response

A global change in gene expression and protein regulation following amino acid starvation signalled by the alarmones guanosine tetraphosphate and guanosine pentaphosphate (collectively referred to as (p)ppGpp) in bacteria and plants, directing resources away from growth and towards amino acid synthesis to promote survival.

Post-segregational killing

A mechanism of plasmid maintenance that is used by some low-copy-number plasmids through the action of toxin–antitoxin modules, whereby any progeny bacterial daughter cells produced after division that have not received a copy of the plasmid will be killed through unregulated toxin activity.

SOS response

A global change in gene expression and protein regulation due to the degradation of the LexA repressor caused by the detection of DNA damage in a bacterial cell by the RecA recombinase.

Alarmones

Intracellular signalling molecules that are produced in response to stress.

Proton motive force

The movement of protons down an electrochemical gradient across a membrane to drive ATP synthesis and motility in bacteria.

DNA gyrase

A topoisomerase enzyme that decreases the supercoiling of DNA during replication and transcription through cleaving, rotating and re-ligating the DNA double-strand.

Trans-translation

A quality control mechanism in protein synthesis that uses transfer-messenger RNA (tmRNA) to rescue a ribosome that has stalled during translation.

Isoniazid

An antibiotic that is commonly used for the treatment of tuberculosis. Isoniazid is a pro-drug that is converted into its active form by the catalase enzyme KatG inside bacterial cells, it is then able to inhibit the cytochrome P450 system, leading to the production of lethal free radicals.

Antimicrobial peptides

(AMPs). Small peptides synthesized by plants and animals that have antimicrobial properties and often target the bacterial membrane.

PhoPQ two-component system

Two-component systems comprise a sensor and a regulator that are able to detect environmental changes and mediate transcriptional changes in response. The PhoPQ system regulates the glycerophospholipid and protein content of the outer membrane in response to pH.

Conditional cooperativity

The autoregulation of certain toxin–antitoxin modules, whereby the antitoxin and toxin are able to form a complex only at a certain stoichiometric ratio, which can then effectively repress transcription of the module.

Transfer-messenger RNA

(tmRNA). A specialized form of RNA that mimics a tRNA with an mRNA-like coding element that is used during trans-translation to continue stalled protein synthesis and target the resultant aberrant protein for degradation.

Deampylation

The removal of an adenylyl group.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fisher, R., Gollan, B. & Helaine, S. Persistent bacterial infections and persister cells. Nat Rev Microbiol 15, 453–464 (2017). https://doi.org/10.1038/nrmicro.2017.42

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrmicro.2017.42

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