1932

Abstract

Persisters are nongrowing, transiently antibiotic-tolerant bacteria within a clonal population of otherwise susceptible cells. Their formation is triggered by environmental cues and involves the main bacterial stress response pathways that allow persisters to survive many harsh conditions, including antibiotic exposure. During infection, bacterial pathogens are exposed to a vast array of stresses in the host and form nongrowing persisters that survive both antibiotics and host immune responses, thereby most likely contributing to the relapse of many infections. While antibiotic persisters have been extensively studied over the last decade, the bulk of the work has focused on how these bacteria survive exposure to drugs in vitro. The ability of persisters to survive their interaction with a host is important yet underinvestigated. In order to tackle the problem of persistence of infections that contribute to the worldwide antibiotic resistance crisis, efforts should be made by scientific communities to understand and merge these two fields of research: antibiotic persisters and host-pathogen interactions. Here we give an overview of the history of the field of antibiotic persistence, report evidence for the importance of persisters in infection, and highlight studies that bridge the two areas.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-micro-020518-115650
2019-09-08
2024-03-29
Loading full text...

Full text loading...

/deliver/fulltext/micro/73/1/annurev-micro-020518-115650.html?itemId=/content/journals/10.1146/annurev-micro-020518-115650&mimeType=html&fmt=ahah

Literature Cited

  1. 1. 
    Allison KR, Brynildsen MP, Collins JJ 2011. Metabolite-enabled eradication of bacterial persisters by aminoglycosides. Nature 473:7346216–20
    [Google Scholar]
  2. 2. 
    Amato SM, Fazen CH, Henry TC, Mok WWK, Orman MA et al. 2014. The role of metabolism in bacterial persistence. Front. Microbiol. 5:70
    [Google Scholar]
  3. 3. 
    Amato SM, Orman MA, Brynildsen MP 2013. Metabolic control of persister formation in Escherichia coli. Mol. Cell 50:4475–87
    [Google Scholar]
  4. 4. 
    Bahar AA, Liu Z, Totsingan F, Buitrago C, Kallenbach N, Ren D 2015. Synthetic dendrimeric peptide active against biofilm and persister cells of Pseudomonas aeruginosa. Appl. Microbiol. Biotechnol 99:198125–35
    [Google Scholar]
  5. 5. 
    Baharoglu Z, Mazel D. 2014. SOS, the formidable strategy of bacteria against aggressions. FEMS Microbiol. Rev. 38:61126–45
    [Google Scholar]
  6. 6. 
    Balaban NQ, Helaine S, Lewis K, Ackermann M, Aldridge B et al. 2019. Definitions and guidelines for research on antibiotic persistence. Nat. Rev. Microbiol 17:441–48Consensus paper on definitions of persisters and persistence and agreed guidelines on how to quantify the level of persisters in a population.
    [Google Scholar]
  7. 7. 
    Balaban NQ, Merrin J, Chait R, Kowalik L, Leibler S 2004. Bacterial persistence as a phenotypic switch. Science 305:56901622–25First visualization of persisters and establishment of a mathematical model to describe the persistence switch.
    [Google Scholar]
  8. 8. 
    Beaber JW, Hochhut B, Waldor MK 2004. SOS response promotes horizontal dissemination of antibiotic resistance genes. Nature 427:696972–74
    [Google Scholar]
  9. 9. 
    Bigger J. 1944. Treatment of staphylococcal infections with penicillin by intermittent sterilisation. Lancet 244:6320497–500
    [Google Scholar]
  10. 10. 
    Brielle R, Pinel-Marie M-L, Felden B 2016. Linking bacterial type I toxins with their actions. Curr. Opin. Microbiol. 30:114–21
    [Google Scholar]
  11. 11. 
    Brochado AR, Telzerow A, Bobonis J, Banzhaf M, Mateus A et al. 2018. Species-specific activity of antibacterial drug combinations. Nature 559:7713259–63
    [Google Scholar]
  12. 12. 
    Brown DR. 2019. Nitrogen starvation induces persister cell formation in Escherichia coli. J. Bacteriol 201:3e00622–18
    [Google Scholar]
  13. 13. 
    Cabral D, Wurster J, Belenky P, Cabral DJ, Wurster JI, Belenky P 2018. Antibiotic persistence as a metabolic adaptation: stress, metabolism, the host, and new directions. Pharmaceuticals 11:114
    [Google Scholar]
  14. 14. 
    Claudi B, Spröte P, Chirkova A, Personnic N, Zankl J et al. 2014. Phenotypic variation of Salmonella in host tissues delays eradication by antimicrobial chemotherapy. Cell 158:4722–33
    [Google Scholar]
  15. 15. 
    Codd A, Teuscher F, Kyle DE, Cheng Q, Gatton ML 2011. Artemisinin-induced parasite dormancy: a plausible mechanism for treatment failure. Malar. J. 10:56
    [Google Scholar]
  16. 16. 
    Cohen NR, Lobritz MA, Collins JJ 2013. Microbial persistence and the road to drug resistance. Cell Host Microbe 13:6632–42
    [Google Scholar]
  17. 17. 
    Conlon BP, Nakayasu ES, Fleck LE, LaFleur MD, Isabella VM et al. 2013. Activated ClpP kills persisters and eradicates a chronic biofilm infection. Nature 503:7476365–70
    [Google Scholar]
  18. 18. 
    Conlon BP, Rowe SE, Gandt AB, Nuxoll AS, Donegan NP et al. 2016. Persister formation in Staphylococcus aureus is associated with ATP depletion. Nat. Microbiol. 1:516051
    [Google Scholar]
  19. 19. 
    Cui P, Niu H, Shi W, Zhang S, Zhang H et al. 2016. Disruption of membrane by colistin kills uropathogenic Escherichia coli persisters and enhances killing of other antibiotics. Antimicrob. Agents Chemother. 60:116867–71
    [Google Scholar]
  20. 20. 
    Defraine V, Fauvart M, Michiels J 2018. Fighting bacterial persistence: Current and emerging anti-persister strategies and therapeutics. Drug Resist. Updates 38:12–26
    [Google Scholar]
  21. 21. 
    Dhar N, McKinney JD. 2010. Mycobacterium tuberculosis persistence mutants identified by screening in isoniazid-treated mice. PNAS 107:2712275–80
    [Google Scholar]
  22. 22. 
    Diard M, Bakkeren E, Cornuault JK, Moor K, Hausmann A et al. 2017. Inflammation boosts bacteriophage transfer between Salmonella spp. Science 355:63301211–15
    [Google Scholar]
  23. 23. 
    Dörr T, Lewis K, Vulić M 2009. SOS response induces persistence to fluoroquinolones in Escherichia coli. PLOS Genet 5:12e1000760
    [Google Scholar]
  24. 24. 
    Dörr T, Vulić M, Lewis K 2010. Ciprofloxacin causes persister formation by inducing the TisB toxin in Escherichia coli. PLOS Biol 8:2e1000317
    [Google Scholar]
  25. 25. 
    Dufour D, Mankovskaia A, Chan Y, Motavaze K, Gong S-G, Lévesque CM 2018. A tripartite toxin-antitoxin module induced by quorum sensing is associated with the persistence phenotype in Streptococcus mutans. Mol. Oral Microbiol 33:6420–29
    [Google Scholar]
  26. 26. 
    Feng J, Shi W, Zhang S, Sullivan D, Auwaerter PG, Zhang Y 2016. A drug combination screen identifies drugs active against amoxicillin-induced round bodies of in vitro Borrelia burgdorferi persisters from an FDA drug library. Front. Microbiol. 7:743
    [Google Scholar]
  27. 27. 
    Fisher RA, Gollan B, Helaine S 2017. Persistent bacterial infections and persister cells. Nat. Rev. Microbiol. 15:8453–64
    [Google Scholar]
  28. 28. 
    Ford AM, Mansur MB, Furness CL, van Delft FW, Okamura J et al. 2015. Protracted dormancy of pre-leukemic stem cells. Leukemia 29:112202–7
    [Google Scholar]
  29. 29. 
    Ford CB, Lin PL, Chase MR, Shah RR, Iartchouk O et al. 2011. Use of whole genome sequencing to estimate the mutation rate of Mycobacterium tuberculosis during latent infection. Nat. Genet. 43:5482–86
    [Google Scholar]
  30. 30. 
    Fridman O, Goldberg A, Ronin I, Shoresh N, Balaban NQ 2014. Optimization of lag time underlies antibiotic tolerance in evolved bacterial populations. Nature 513:7518418–21ScanLag was used here to record bacterial regrowth after rounds of antibiotic treatments.
    [Google Scholar]
  31. 31. 
    Furman R, Danhart EM, NandyMazumdar M, Yuan C, Foster MP, Artsimovitch I 2015. pH dependence of the stress regulator DksA. PLOS ONE 10:3e0120746
    [Google Scholar]
  32. 32. 
    Ghosh A, Baltekin Ö, Wäneskog M, Elkhalifa D, Hammarlöf DL et al. 2018. Contact‐dependent growth inhibition induces high levels of antibiotic‐tolerant persister cells in clonal bacterial populations. EMBO J 37:9e98026
    [Google Scholar]
  33. 33. 
    Goormaghtigh F, Fraikin N, Putrinš M, Hallaert T, Hauryliuk V et al. 2018. Reassessing the role of type II toxin-antitoxin systems in formation of Escherichia coli type II persister cells. mBio 9:3e00640–18
    [Google Scholar]
  34. 34. 
    Grant SS, Kaufmann BB, Chand NS, Haseley N, Hung DT 2012. Eradication of bacterial persisters with antibiotic-generated hydroxyl radicals. PNAS 109:3012147–52
    [Google Scholar]
  35. 35. 
    Griffin AJ, Li L-X, Voedisch S, Pabst O, McSorley SJ 2011. Dissemination of persistent intestinal bacteria via the mesenteric lymph nodes causes typhoid relapse. Infect. Immun. 79:41479–88Characterization of MLNs as persister reservoir in a murine model of systemic Salmonella infection.
    [Google Scholar]
  36. 36. 
    Gutierrez A, Laureti L, Crussard S, Abida H, Rodríguez-Rojas A et al. 2013. β-Lactam antibiotics promote bacterial mutagenesis via an RpoS-mediated reduction in replication fidelity. Nat. Commun. 4:11610
    [Google Scholar]
  37. 37. 
    Gutiérrez D, Ruas-Madiedo P, Martínez B, Rodríguez A, García P 2014. Effective removal of staphylococcal biofilms by the endolysin LysH5. PLOS ONE 9:e107307
    [Google Scholar]
  38. 38. 
    Hall AM, Gollan B, Helaine S 2017. Toxin-antitoxin systems: reversible toxicity. Curr. Opin. Microbiol. 36:102–10
    [Google Scholar]
  39. 39. 
    Hangauer MJ, Viswanathan VS, Ryan MJ, Bole D, Eaton JK et al. 2017. Drug-tolerant persister cancer cells are vulnerable to GPX4 inhibition. Nature 551:7679247–50
    [Google Scholar]
  40. 40. 
    Hansen S, Lewis K, Vulić M 2008. Role of global regulators and nucleotide metabolism in antibiotic tolerance in Escherichia coli. Antimicrob. Agents Chemother 52:82718–26
    [Google Scholar]
  41. 41. 
    Harms A, Brodersen DE, Mitarai N, Gerdes K 2018. Toxins, targets, and triggers: an overview of toxin-antitoxin biology. Mol. Cell 70:5768–84
    [Google Scholar]
  42. 42. 
    Harms A, Fino C, Sørensen MA, Semsey S, Gerdes K 2017. Prophages and growth dynamics confound experimental results with antibiotic-tolerant persister cells. mBio 8:6e01964–17
    [Google Scholar]
  43. 43. 
    Harms A, Maisonneuve E, Gerdes K 2016. Mechanisms of bacterial persistence during stress and antibiotic exposure. Science 354:6318aaf4268
    [Google Scholar]
  44. 44. 
    Harrison JJ, Ceri H, Roper NJ, Badry EA, Sproule KM, Turner RJ 2005. Persister cells mediate tolerance to metal oxyanions in Escherichia coli. Microbiology 151:103181–95
    [Google Scholar]
  45. 45. 
    Harrison JJ, Wade WD, Akierman S, Vacchi-Suzzi C, Stremick CA et al. 2009. The chromosomal toxin gene yafQ is a determinant of multidrug tolerance for Escherichia coli growing in a biofilm. Antimicrob. Agents Chemother. 53:62253–58
    [Google Scholar]
  46. 46. 
    Helaine S, Cheverton AM, Watson KG, Faure LM, Matthews SA, Holden DW 2014. Internalization of Salmonella by macrophages induces formation of nonreplicating persisters. Science 343:6167204–8
    [Google Scholar]
  47. 47. 
    Hobby GL, Meyer K, Chaffee E 1942. Observations on the mechanism of action of penicillin. Exp. Biol. Med. 50:2281–85
    [Google Scholar]
  48. 48. 
    Hogg T, Mechold U, Malke H, Cashel M, Hilgenfeld R et al. 2004. Conformational antagonism between opposing active sites in a bifunctional RelA/SpoT homolog modulates (p)ppGpp metabolism during the stringent response. Cell 117:457–68
    [Google Scholar]
  49. 49. 
    Hu Y, Coates ARM. 2013. Enhancement by novel anti-methicillin-resistant Staphylococcus aureus compound HT61 of the activity of neomycin, gentamicin, mupirocin and chlorhexidine: in vitro and in vivo studies. J. Antimicrob. Chemother. 68:374–84
    [Google Scholar]
  50. 50. 
    Huang L, Kushner NL, Theriault ME, Pisu D, Tan S et al. 2018. The deconstructed granuloma: a complex high-throughput drug screening platform for the discovery of host-directed therapeutics against tuberculosis. Front. Cell. Infect. Microbiol. 8:275
    [Google Scholar]
  51. 51. 
    Jain P, Weinrick BC, Kalivoda EJ, Yang H, Munsamy V et al. 2016. Dual-reporter mycobacteriophages (Φ2 DRMs) reveal preexisting Mycobacterium tuberculosis persistent cells in human sputum. mBio 7:5e01023–16
    [Google Scholar]
  52. 52. 
    Jensen A, Fagö-Olsen H, Sørensen CH, Kilian M 2013. Molecular mapping to species level of the tonsillar crypt microbiota associated with health and recurrent tonsillitis. PLOS ONE 8:2e56418
    [Google Scholar]
  53. 53. 
    Johnson PJT, Levin BR. 2013. Pharmacodynamics, population dynamics, and the evolution of persistence in Staphylococcus aureus. PLOS Genet 9:1e1003123
    [Google Scholar]
  54. 54. 
    Jolly MK, Kulkarni P, Weninger K, Orban J, Levine H 2018. Phenotypic plasticity, bet-hedging, and androgen independence in prostate cancer: role of non-genetic heterogeneity. Front. Oncol. 8:50
    [Google Scholar]
  55. 55. 
    Kaiser P, Regoes RR, Dolowschiak T, Wotzka SY, Lengefeld J et al. 2014. Cecum lymph node dendritic cells harbor slow-growing bacteria phenotypically tolerant to antibiotic treatment. PLOS Biol 12:2e1001793
    [Google Scholar]
  56. 56. 
    Keren I, Kaldalu N, Spoering A, Wang Y, Lewis K 2004. Persister cells and tolerance to antimicrobials. FEMS Microbiol. Lett. 230:113–18
    [Google Scholar]
  57. 57. 
    Keren I, Minami S, Rubin E, Lewis K 2011. Characterization and transcriptome analysis of Mycobacterium tuberculosis persisters. mBio 2:3e00100–11
    [Google Scholar]
  58. 58. 
    Kim J-S, Heo P, Yang T-J, Lee K-S, Cho D-H et al. 2011. Selective killing of bacterial persisters by a single chemical compound without affecting normal antibiotic-sensitive cells. Antimicrob. Agents Chemother. 55:115380–83
    [Google Scholar]
  59. 59. 
    Kim J-S, Liu L, Fitzsimmons LF, Wang Y, Crawford MA et al. 2018. DksA-DnaJ redox interactions provide a signal for the activation of bacterial RNA polymerase. PNAS 115:50E11780–89
    [Google Scholar]
  60. 60. 
    Kim Y, Wood TK. 2010. Toxins Hha and CspD and small RNA regulator Hfq are involved in persister cell formation through MqsR in Escherichia coli. Biochem. Biophys. Res. Commun 391:1209–13
    [Google Scholar]
  61. 61. 
    Kim Y-R, Yang C-S. 2017. Host-directed therapeutics as a novel approach for tuberculosis treatment. J. Microbiol. Biotechnol. 27:91549–58
    [Google Scholar]
  62. 62. 
    Kubistova L, Dvoracek L, Tkadlec J, Melter O, Licha I 2018. Environmental stress affects the formation of Staphylococcus aureus persisters tolerant to antibiotics. Microb. Drug Resist. 24:5547–55
    [Google Scholar]
  63. 63. 
    Kuczyńska-Wiśnik D, Stojowska K, Matuszewska E, Leszczyńska D, Algara MM et al. 2015. Lack of intracellular trehalose affects formation of Escherichia coli persister cells. Microbiology 161:4786–96
    [Google Scholar]
  64. 64. 
    Kwan BW, Chowdhury N, Wood TK 2015. Combatting bacterial infections by killing persister cells with mitomycin C. Environ. Microbiol. 17:114406–14
    [Google Scholar]
  65. 65. 
    Kwan BW, Valenta JA, Benedik MJ, Wood TK 2013. Arrested protein synthesis increases persister-like cell formation. Antimicrob. Agents Chemother. 57:31468–73
    [Google Scholar]
  66. 66. 
    Lafleur MD, Qi Q, Lewis K 2010. Patients with long-term oral carriage harbor high-persister mutants of Candida albicans. Antimicrob. Agents Chemother 54:139–44
    [Google Scholar]
  67. 67. 
    Layton JC, Foster PL. 2003. Error-prone DNA polymerase IV is controlled by the stress-response sigma factor, RpoS, in Escherichia coli. Mol. Microbiol 50:2549–61
    [Google Scholar]
  68. 68. 
    Leszczynska D, Matuszewska E, Kuczynska-Wisnik D, Furmanek-Blaszk B, Laskowska E 2013. The formation of persister cells in stationary-phase cultures of Escherichia coli is associated with the aggregation of endogenous proteins. PLOS ONE 8:1e54737
    [Google Scholar]
  69. 69. 
    Leung V, Lévesque CM. 2012. A stress-inducible quorum-sensing peptide mediates the formation of persister cells with noninherited multidrug tolerance. J. Bacteriol. 194:92265–74
    [Google Scholar]
  70. 70. 
    Levin-Reisman I, Gefen O, Fridman O, Ronin I, Shwa D et al. 2010. Automated imaging with ScanLag reveals previously undetectable bacterial growth phenotypes. Nat. Methods 7:9737–39
    [Google Scholar]
  71. 71. 
    Levin-Reisman I, Ronin I, Gefen O, Braniss I, Shoresh N, Balaban NQ 2017. Antibiotic tolerance facilitates the evolution of resistance. Science 355:6327826–30
    [Google Scholar]
  72. 72. 
    Li T, Yin N, Liu H, Pei J, Lai L 2016. Novel inhibitors of toxin HipA reduce multidrug tolerant persisters. ACS Med. Chem. Lett. 7:449–53
    [Google Scholar]
  73. 73. 
    Lin PL, Flynn JL. 2010. Understanding latent tuberculosis: a moving target. J. Immunol. 185:115–22
    [Google Scholar]
  74. 74. 
    Liu Y, Tan S, Huang L, Abramovitch RB, Rohde KH et al. 2016. Immune activation of the host cell induces drug tolerance in Mycobacterium tuberculosis both in vitro and in vivo. J. Exp. Med. 213:809–25
    [Google Scholar]
  75. 75. 
    Lou C, Li Z, Ouyang Q 2008. A molecular model for persister in E. coli. J. Theor. Biol 255:2205–9
    [Google Scholar]
  76. 76. 
    Lu TK, Collins JJ. 2009. Engineered bacteriophage targeting gene networks as adjuvants for antibiotic therapy. PNAS 106:124629–34
    [Google Scholar]
  77. 77. 
    Manina G, Dhar N, McKinney JD 2015. Stress and host immunity amplify Mycobacterium tuberculosis phenotypic heterogeneity and induce nongrowing metabolically active forms. Cell Host Microbe 17:132–46
    [Google Scholar]
  78. 78. 
    Mariathasan S, Tan M. 2017. Antibody-antibiotic conjugates : a novel therapeutic platform against bacterial infections. Trends Mol. Med. 23:2135–49
    [Google Scholar]
  79. 79. 
    Marques CNH, Morozov A, Planzos P, Zelaya HM 2014. The fatty acid signaling molecule cis-2-decenoic acid increases metabolic activity and reverts persister cells to an antimicrobial-susceptible state. Appl. Environ. Microbiol. 80:226976–91
    [Google Scholar]
  80. 80. 
    Martins D, McKay G, Sampathkumar G, Khakimova M, English AM, Nguyen D 2018. Superoxide dismutase activity confers (p)ppGpp-mediated antibiotic tolerance to stationary-phase Pseudomonas aeruginosa. PNAS 115:399797–802
    [Google Scholar]
  81. 81. 
    Mok WWK, Brynildsen MP. 2018. Timing of DNA damage responses impacts persistence to fluoroquinolones. PNAS 115:27E6301–9
    [Google Scholar]
  82. 82. 
    Möker N, Dean CR, Tao J 2010. Pseudomonas aeruginosa increases formation of multidrug-tolerant persister cells in response to quorum-sensing signaling molecules. J. Bacteriol. 192:71946–55
    [Google Scholar]
  83. 83. 
    Monack DM. 2013. Helicobacter and Salmonella persistent infection strategies. Cold Spring Harb. Perspect. Med. 3:12a010348
    [Google Scholar]
  84. 84. 
    Mouton JM, Helaine S, Holden DW, Sampson SL 2016. Elucidating population-wide mycobacterial replication dynamics at the single-cell level. Microbiology 162:6966–78
    [Google Scholar]
  85. 85. 
    Moyed HS, Bertrand KP. 1983. hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. J. Bacteriol. 155:2768–75
    [Google Scholar]
  86. 86. 
    Mulcahy LR, Burns JL, Lory S, Lewis K 2010. Emergence of Pseudomonas aeruginosa strains producing high levels of persister cells in patients with cystic fibrosis. J. Bacteriol. 192:236191–99Reveals the selection of high persister mutants over years of antibiotic treatments.
    [Google Scholar]
  87. 87. 
    Nakazawa S, Maoka T, Uemura H, Ito Y, Kanbara H 2002. Malaria parasites giving rise to recrudescence in vitro. Antimicrob. Agents Chemother. 46:4958–65
    [Google Scholar]
  88. 88. 
    Narayanaswamy VP, Keagy LL, Duris K, Wiesmann W, Loughran AJ et al. 2018. Novel glycopolymer eradicates antibiotic- and CCCP-induced persister cells in Pseudomonas aeruginosa. Front. Microbiol 9:1724
    [Google Scholar]
  89. 89. 
    Newsom SW. 1970. Staphylococcal persisters grown from empyema fluid on L-form medium. J. Med. Microbiol. 3:4669–73
    [Google Scholar]
  90. 90. 
    Nguyen D, Joshi-Datar A, Lepine F, Bauerle E, Olakanmi O et al. 2011. Active starvation responses mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. Science 334:6058982–86
    [Google Scholar]
  91. 91. 
    O'Brien VP, Hannan TJ, Yu L, Livny J, Roberson EDO et al. 2016. A mucosal imprint left by prior Escherichia coli bladder infection sensitizes to recurrent disease. Nat. Microbiol. 2:116196
    [Google Scholar]
  92. 92. 
    Okoro CK, Kingsley RA, Quail MA, Kankwatira AM, Feasey NA et al. 2012. High-resolution single nucleotide polymorphism analysis distinguishes recrudescence and reinfection in recurrent invasive nontyphoidal Salmonella Typhimurium disease. Clin. Infect. Dis. 54:7955–63
    [Google Scholar]
  93. 93. 
    Orman MA, Brynildsen MP. 2013. Dormancy is not necessary or sufficient for bacterial persistence. Antimicrob. Agents Chemother. 57:73230–39
    [Google Scholar]
  94. 94. 
    Orman MA, Brynildsen MP. 2016. Persister formation in Escherichia coli can be inhibited by treatment with nitric oxide. Free Radic. Biol. Med. 93:145–54
    [Google Scholar]
  95. 95. 
    Pan J, Xie X, Tian W, Bahar AA, Lin N et al. 2013. (Z)-4-Bromo-5-(bromomethylene)-3-methylfuran-2(5H)-one sensitizes Escherichia coli persister cells to antibiotics. Appl. Microbiol. Biotechnol. 97:209145–54
    [Google Scholar]
  96. 96. 
    Patel A, Malinovska L, Saha S, Wang J, Alberti S et al. 2017. ATP as a biological hydrotrope. Science 356:6339753–56
    [Google Scholar]
  97. 97. 
    Pearl Mizrahi S, Gefen O, Simon I, Balaban NQ 2016. Persistence to anti-cancer treatments in the stationary to proliferating transition. Cell Cycle 15:243442–53
    [Google Scholar]
  98. 98. 
    Phillips I, Culebras E, Moreno F, Baquero F 1987. Induction of the SOS response by new 4-quinolones. J. Antimicrob. Chemother. 20:5631–38
    [Google Scholar]
  99. 99. 
    Potrykus K, Cashel M. 2008. (p)ppGpp: still magical. ? Annu. Rev. Microbiol. 62:135–51
    [Google Scholar]
  100. 100. 
    Pu Y, Li Y, Jin X, Tian T, Ma Q et al. 2019. ATP-dependent dynamic protein aggregation regulates bacterial dormancy depth critical for antibiotic tolerance. Mol. Cell 73:1143–56.e4
    [Google Scholar]
  101. 101. 
    Que Y-A, Hazan R, Strobel B, Maura D, He J et al. 2013. A quorum sensing small volatile molecule promotes antibiotic tolerance in bacteria. PLOS ONE 8:12e80140
    [Google Scholar]
  102. 102. 
    Radzikowski JL, Vedelaar S, Siegel D, Ortega ÁD, Schmidt A, Heinemann M 2016. Bacterial persistence is an active σS stress response to metabolic flux limitation. Mol. Syst. Biol. 12:9882
    [Google Scholar]
  103. 103. 
    Ragheb MN, Thomason MK, Hsu C, Nugent P, Gage J et al. 2019. Inhibiting the evolution of antibiotic resistance. Mol. Cell 73:1157–165.e5
    [Google Scholar]
  104. 104. 
    Ren H, He X, Zou X, Wang G, Li S, Wu Y 2015. Gradual increase in antibiotic concentration affects persistence of Klebsiella pneumoniae. J. Antimicrob. Chemother 70:123267–72
    [Google Scholar]
  105. 105. 
    Ronneau S, Helaine S. 2019. Clarifying the link between toxin-antitoxin modules and bacterial persistence. J. Mol. Biol. In press
    [Google Scholar]
  106. 106. 
    Roostalu J, Jõers A, Luidalepp H, Kaldalu N, Tenson T 2008. Cell division in Escherichia coli cultures monitored at single cell resolution. BMC Microbiol 8:168
    [Google Scholar]
  107. 107. 
    Rotem E, Loinger A, Ronin I, Levin-Reisman I, Gabay C et al. 2010. Regulation of phenotypic variability by a threshold-based mechanism underlies bacterial persistence. PNAS 107:2812541–46
    [Google Scholar]
  108. 108. 
    Schuch R, Lee HM, Schneider BC, Sauve KL, Law C et al. 2014. Combination therapy with lysin CF-301 and antibiotic is superior to antibiotic alone for treating methicillin-resistant Staphylococcus aureus-induced murine bacteremia. J. Infect. Dis. 209:1469–78
    [Google Scholar]
  109. 109. 
    Schumacher MA, Balani P, Min J, Chinnam NB, Hansen S et al. 2015. HipBA-promoter structures reveal the basis of heritable multidrug tolerance. Nature 524:756359–64Biological insights on the role of HipA in persister formation.
    [Google Scholar]
  110. 110. 
    Schumacher MA, Piro KM, Xu W, Hansen S, Lewis K, Brennan RG 2009. Molecular mechanisms of HipA-mediated multidrug tolerance and its neutralization by HipB. Science 323:5912396–401
    [Google Scholar]
  111. 111. 
    Shan Y, Brown Gandt A, Rowe SE, Deisinger JP, Conlon BP, Lewis K 2017. ATP-dependent persister formation in Escherichia coli. mBio 8:1e02267–16
    [Google Scholar]
  112. 112. 
    Singh A, Weinberger LS. 2009. Stochastic gene expression as a molecular switch for viral latency. Curr. Opin. Microbiol. 12:460–66
    [Google Scholar]
  113. 113. 
    Spoering AL, Lewis K. 2001. Biofilms and planktonic cells of Pseudomonas aeruginosa have similar resistance to killing by antimicrobials. J. Bacteriol. 183:236746–51
    [Google Scholar]
  114. 114. 
    Stapels DAC, Hill PWS, Westermann AJ, Fisher RA, Thurston TL et al. 2018. Salmonella persisters undermine host immune defenses during antibiotic treatment. Science 362:64191156–60
    [Google Scholar]
  115. 115. 
    Tuomanen E, Cozens R, Tosch W, Zak O, Tomasz A 1986. The rate of killing of Escherichia coli by beta-lactam antibiotics is strictly proportional to the rate of bacterial growth. J. Gen. Microbiol. 132:51297–304
    [Google Scholar]
  116. 116. 
    Van den Bergh B, Fauvart M, Michiels J 2017. Formation, physiology, ecology, evolution and clinical importance of bacterial persisters. FEMS Microbiol. Rev. 41:3219–51
    [Google Scholar]
  117. 117. 
    Van den Bergh B, Michiels JE, Wenseleers T, Windels EM, Boer PV et al. 2016. Frequency of antibiotic application drives rapid evolutionary adaptation of Escherichia coli persistence. Nat. Microbiol. 1:516020
    [Google Scholar]
  118. 118. 
    Vega NM, Allison KR, Khalil AS, Collins JJ 2012. Signaling-mediated bacterial persister formation. Nat. Chem. Biol. 8:5431–33
    [Google Scholar]
  119. 119. 
    Vega NM, Allison KR, Samuels AN, Klempner MS, Collins JJ 2013. Salmonella typhimurium intercepts Escherichia coli signaling to enhance antibiotic tolerance. PNAS 110:3514420–25
    [Google Scholar]
  120. 120. 
    Verstraeten N, Knapen W, Fauvart M, Michiels J 2016. A historical perspective on bacterial persistence. Methods Mol. Biol. 1333:3–13
    [Google Scholar]
  121. 121. 
    Verstraeten N, Knapen WJ, Kint CI, Liebens V, Van den Bergh B et al. 2015. Obg and membrane depolarization are part of a microbial bet-hedging strategy that leads to antibiotic tolerance. Mol. Cell 59:19–21
    [Google Scholar]
  122. 122. 
    Völzing KG, Brynildsen MP. 2015. Stationary-phase persisters to ofloxacin sustain DNA damage and require repair systems only during recovery. mBio 6:5e00731–15
    [Google Scholar]
  123. 123. 
    Wang T, El Meouche I, Dunlop MJ 2017. Bacterial persistence induced by salicylate via reactive oxygen species. Sci. Rep. 7:143839
    [Google Scholar]
  124. 124. 
    Westermann AJ, Förstner KU, Amman F, Barquist L, Chao Y et al. 2016. Dual RNA-seq unveils noncoding RNA functions in host-pathogen interactions. Nature 529:7587496–501
    [Google Scholar]
  125. 125. 
    Willett JLE, Ruhe ZC, Goulding CW, Low DA, Hayes CS 2015. Contact-dependent growth inhibition (CDI) and CdiB/CdiA two-partner secretion proteins. J. Mol. Biol. 427:233754–65
    [Google Scholar]
  126. 126. 
    Windels EM, Michiels JE, Fauvart M, Wenseleers T, Van den Bergh B, Michiels J 2019. Bacterial persistence promotes the evolution of antibiotic resistance by increasing survival and mutation rates. ISME J 13:1239–51
    [Google Scholar]
  127. 127. 
    Wu Y, Vulić M, Keren I, Lewis K 2012. Role of oxidative stress in persister tolerance. Antimicrob. Agents Chemother. 56:94922–26
    [Google Scholar]
  128. 128. 
    Wuyts J, Van Dijck P, Holtappels M 2018. Fungal persister cells: the basis for recalcitrant infections?. PLOS Pathog 14:e1007301
    [Google Scholar]
  129. 129. 
    Zhou C, Lehar S, Gutierrez J, Rosenberger CM, Ljumanovic N et al. 2016. Pharmacokinetics and pharmacodynamics of DSTA4637A: a novel THIOMAB™ antibody antibiotic conjugate against Staphylococcus aureus in mice. mAbs 8:1612–19
    [Google Scholar]
/content/journals/10.1146/annurev-micro-020518-115650
Loading
/content/journals/10.1146/annurev-micro-020518-115650
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error