Studying the role of persistence in the emergence of antibiotic resistance through experimental evolution

Etthel Windels
Persbericht

Bacteriële persistentie: een nieuwe weg naar superbacteriën

In de context van de oprukkende antibioticumcrisis wordt de bestrijding van infecties steeds uitdagender. Paradoxaal genoeg heeft jarenlang overvloedig antibioticumgebruik onze bacteriële vijanden alleen maar sterker gemaakt. Resistentie is hun meest bekende verdedigingsstrategie, maar daarmee is niet alles gezegd. Haast alle bacteriële populaties bevatten een piepkleine fractie inactieve cellen die tijdelijk ongevoelig zijn voor antibiotica. Dat deze gespecialiseerde cellen dringend meer aandacht verdienen, blijkt nu uit het feit dat ze belangrijke voorlopers vormen van resistente bacteriën.

De ontdekking van antibiotica enkele decennia geleden leverde ons levensreddende middelen op in de strijd tegen infectieziektes. Dankzij de ongeëvenaarde effectiviteit waarmee ze bacteriën verdelgen, geloofden we jarenlang dat die strijd definitief gestreden was. Helaas is niets minder waar. Onder meer de opkomst van resistente bacteriën, die geen hinder ondervinden van het gebruikte antibioticum, tempert de aanvankelijke euforie. Hoewel de aanpak van antibioticumresistentie pas sinds kort op de beleidsagenda staat, doken de eerste problemen reeds op bij de start van het antibioticumtijdperk. Het fenomeen werd echter sterk in de hand gewerkt door het jarenlang overmatig gebruik en misbruik in de geneeskunde en de veehouderij. Het gevolg is een alarmerende opkomst van dodelijke ‘superbacteriën’ die we met geen enkel antibioticum kunnen bestrijden en die zich met een onvoorziene snelheid verspreiden. Door de afnemende werking van de bestaande medicatie beginnen we stilaan te beseffen dat banale infecties of simpele verwondingen opnieuw onze dood kunnen betekenen.

Maar het gevaar schuilt ook in een andere hoek. Het vaak beperkte succes van antibioticumbehandelingen bij chronische infecties heeft immers ook een veel minder bekende oorzaak, die als ‘persistentie’ bestempeld wordt. Hoewel medicatie vaak wel werkzaam is tegenover het overgrote deel van een bacteriële populatie, is een heel klein percentage van zogenaamde ‘persistorcellen’ toch in staat om aan de werking te ontsnappen en zo een volledige uitroeiing van de populatie te vermijden. Dat doen ze door tijdens de behandeling in een soort overlevingsmodus te gaan: hun metabole activiteit daalt aanzienlijk en ze vermenigvuldigen zich niet langer, ten voordele van een sterke tolerantie tegenover tal van antibiotica. Na het stopzetten van de kuur ontwaakt die kleine fractie persistorcellen en vermenigvuldigen ze zich, met hernieuwde symptomen tot gevolg. En alsof dat nog niet zorgwekkend genoeg is, toonde recent onderzoek aan dat frequente antibioticumbehandelingen het aantal persistorcellen in een bacteriële populatie kunnen doen toenemen.

Hoe kunnen we de huidige antibioticumcrisis opnieuw onder controle krijgen? We moeten goed beseffen dat antibiotica een verantwoord en oordeelkundig gebruik vereisen, willen we de effectiviteit ervan niet verder teniet doen. Inzicht in het ontstaan en de mechanismen van zowel resistentie als persistentie vormen dan ook een belangrijk vertrekpunt bij de keuze van een geschikte antibioticumbehandeling. De resultaten van dit eindwerk tonen aan dat persistentie een rol speelt in het ontstaan van resistentie en dus een nieuwe factor is waar antibioticumtherapieën in de toekomst rekening mee moeten houden.

Aan de hand van wiskundige modellen konden we reeds voorspellen dat er sneller resistentie zou ontstaan in een bacteriële populatie die meer persistorcellen bevat. Door antibioticumbehandelingen na te bootsen in het lab, toonden we aan dat die voorspellingen ook overeenkomen met de realiteit. De onderliggende verklaring voor dit fenomeen is tweevoudig. Enerzijds blijven persistorcellen vaak over na een antibioticumbehandeling en vormen ze zo een bron van resistente bacteriën. Anderzijds ontdekten we dat de SOS respons een extra link vormt tussen persistentie en resistentie. De SOS respons is een mechanisme dat het ontstaan van resistentie stimuleert en dat geactiveerd wordt wanneer het DNA van bacteriën beschadigd is, bijvoorbeeld door bepaalde antibiotica. In persistorcellen is de SOS respons actiever, waardoor ze meer kans maken om resistent te worden. Als we erin slagen om dit mechanisme te onderdrukken tijdens een antibioticumbehandeling, dan kunnen we hiermee het ontstaan van resistentie doen vertragen.

Ons onderzoek wijst persistentie dus aan als een nieuwe risicofactor in het ontstaan van antibioticumresistentie. Dit betekent dat persistentie een nog belangrijkere rol speelt in de huidige antibioticumcrisis dan aanvankelijk verwacht. Enige nuance is echter op zijn plaats. Voorlopig toonden we enkel een effect aan wanneer we bacteriën lieten groeien op een vaste voedingsbodem, een situatie die gelijkenissen vertoont met bepaalde chronische weefselinfecties. Verder onderzoek is nodig om te weten te komen of hetzelfde zich voordoet bij andere infectievormen.

Willen we ziekmakende bacteriën alsnog te slim af zijn, dan moeten we onze strategie dus grondig herzien. Het onderdrukken van persistentie moet hierin dringend een nieuw aandachtspunt vormen, niet alleen om chronische infecties te bestrijden, maar ook om het ontstaan van antibioticumresistentie af te remmen.

Bibliografie

[1] Acosta, M. B., Ferreira, R. C., Padilla, G., Ferreira, L. C. and Costa, S. O. (2000). Altered expression of oligopeptide-binding protein (OppA) and aminoglycoside resistance in laboratory and clinical Escherichia coli strains. Journal of Medical Microbiology, 49(5):409–413.

[2] Adams, K. N., Takaki, K., Connolly, L. E., Wiedenhoft, H., Winglee, K., Humbert, O., Edelstein, P. H., Cosma, C. L. and Ramakrishnan, L. (2011). Drug tolerance in replicating mycobacteria mediated by a macrophage-induced efflux mechanism. Cell, 145(1):39–53.

[3] Ades, S. E. (2008). Regulation by destruction: design of the σE-envelope stress response. Current Opinion in Microbiology, 11(6):535–540.

[4] Akiyama, Y., Kanehara, K. and Ito, K. (2004). RseP (YaeL), an Escherichia coli RIP protease, cleaves transmembrane sequences. The EMBO Journal, 23(22):4434–4442.

[5] Alekshun, M. N. and Levy, S. B. (2007). Molecular mechanisms of antibacterial multidrug resistance. Cell, 128(6):1037–1050.

[6] Altuvia, S., Weinstein-Fischer, D., Zhang, A., Postow, L. and Storz, G. (1997). A small, stable RNA induced by oxidative stress: Role as a pleiotropic regulator and antimutator. Cell, 90(1):43–53.

[7] Amato, S. M., Orman, M. A. and Brynildsen, M. P. (2013). Metabolic control of persister formation in Escherichia coli. Molecular Cell, 50(4):475–487.

[8] Andersson, D. I. and Hughes, D. (2010). Antibiotic resistance and its cost: is it possible to reverse resistance? Nature Reviews Microbiology, 8(4):260–271.

[9] Andersson, D. I. and Hughes, D. (2014). Microbiological effects of sublethal levels of antibiotics. Nature Reviews Microbiology, 12(7):465–478.

[10] Andrews, J. M. (2001). Determination of minimum inhibitory concentrations. Journal of Antimicrobial Chemotherapy, 48:5–16.

[11] Ankomah, P. and Levin, B. R. (2014). Exploring the collaboration between antibiotics and the immune response in the treatment of acute, self-limiting infections. Proceedings of the National Academy of Sciences of the United States of America, 111(23):8331–8338.

[12] Baba, T., Ara, T., Hasegawa, M., Takai, Y., Okumura, Y., Baba, M., Datsenko, K. A., Tomita, M., Wanner, B. L. and Mori, H. (2006). Construction of Escherichia coli K-12 in-frame, single-gene knockout mutants: the Keio collection. Molecular Systems Biology, 2:2006.0008.

[13] Baek, S., Li, A. H. and Sassetti, C. M. (2011). Metabolic regulation of mycobacterial growth and antibiotic sensitivity. PLoS Biology, 9(5):e1001065.

[14] Baharoglu, Z. and Mazel, D. (2014). SOS, the formidable strategy of bacteria against aggressions. FEMS Microbiology Reviews, 38(6):1126–1145.

[15] Balaban, N. Q. (2011). Persistence: mechanisms for triggering and enhancing phenotypic variability. Current Opinion in Genetics and Development, 21(6):768–775.

[16] Balaban, N. Q., Gerdes, K., Lewis, K. and McKinney, J. D. (2013). A problem of persistence: still more questions than answers? Nature Reviews Microbiology, 11(8):587–91.

[17] Balaban, N. Q., Merrin, J., Chait, R., Kowalik, L. and Leibler, S. (2004). Bacterial persistence as a phenotypic switch. Science, 305:1622–1625.

[18] Barrick, J. E. and Lenski, R. E. (2013). Genome dynamics during experimental evolution. Nature Reviews Genetics, 14(12):827–839.

[19] Battesti, A., Majdalani, N. and Gottesman, S. (2011). The RpoS-mediated general stress response in Escherichia coli. Annual Review of Microbiology, 65:189–213.

[20] Baysarowich, J., Koteva, K., Hughes, D. W., Ejim, L., Griffiths, E., Zhang, K., Junop, M. and Wright, G. D. (2008). Rifamycin antibiotic resistance by ADP-ribosylation: Structure and diversity of Arr. Proceedings of the National Academy of Sciences of the United States of America, 105(12):4886–4891.

[21] Beaber, J. W., Hochhut, B. and Waldor, M. K. (2004). SOS response promotes horizontal dissemination of antibiotic resistance genes. Nature, 427:72–74.

[22] Becker, G., Klauck, E. and Hengge-Aronis, R. (1999). Regulation of RpoS proteolysis in Escherichiacoli: the response regulator RssB is a recognition factor that interacts with the turnover element in RpoS. Proceedings of the National Academy of Sciences of the United States of America, 96(11):6439– 6444.

[23] Bhargava, N., Sharma, P. and Capalash, N. (2014). Pyocyanin stimulates quorum sensing-mediated tolerance to oxidative stress and increases persister cell populations in Acinetobacter baumannii. Infection and Immunity, 82(8):3417–3425.

[24] Bigger, J. W. (1944). Treatment of staphylococcal infections with penicillin by intermittent sterilisation. The Lancet, ii(8):497–500.

[25] Blango, M. G. and Mulvey, M. A. (2010). Persistence of uropathogenic Escherichia coli in the face of multiple antibiotics. Antimicrobial Agents and Chemotherapy, 54(5):1855–1863.

[26] Blattner, F., Plunkett G, I., Bloch, C., Perna, N., Burland, V., Riley, M., Collado-Vides, J., Glasner, J., Rode, C., Mayhew, G., Gregor, J., Davis, N., Kirkpatrick, H., Goeden, M., Rose, D., Mau, B. and Shao, Y. (1997). The complete genome sequence of Escherichia coli K-12. Science, 277:1453–1462.

[27] Breidenstein, E. B. M., de la Fuente-Núñez, C. and Hancock, R. E. W. (2011). Pseudomonas aeruginosa: All roads lead to resistance. Trends in Microbiology, 19(8):419–426.

[28] Brennan, R. G. and Link, T. M. (2007). Hfq structure, function and ligand binding. Current Opinion in Microbiology, 10(2):125–133.

[29] Butala, M., Žgur-Bertok, D. and Busby, S. J. W. (2009). The bacterial LexA transcriptional repressor. Cellular and Molecular Life Sciences, 66(1):82–93.

[30] Campbell, E. A., Korzheva, N., Mustaev, A., Murakami, K., Nair, S., Goldfarb, A. and Darst, S. A. (2001). Structural mechanism for rifampicin inhibition of bacterial RNA polymerase. Cell, 104(6):901–912.

[31] Caroff, N., Espaze, E., Gautreau, D., Richet, H. and Reynaud, A. (2000). Analysis of the effects of -42 and -32 ampC promoter mutations in clinical isolates of Escherichia coli hyperproducing AmpC. Journal of Antimicrobial Chemotherapy, 45(6):783–788.

[32] Cataudella, I., Sneppen, K., Gerdes, K. and Mitarai, N. (2013). Conditional cooperativity of toxin - antitoxin regulation can mediate bistability between growth and dormancy. PLoS Computational Biology, 9(8):e1003174.

[33] Chiang, S. M. and Schellhorn, H. E. (2012). Regulators of oxidative stress response genes in Escherichia coli and their functional conservation in bacteria. Archives of Biochemistry and Biophysics, 525(2):161–169.

[34] Chou, H., Chiu, H., Delaney, N. F., Segrè, D. and Marx, C. J. (2011). Diminishing returns epistasis among beneficial mutations decelerates adaptation. Science, 332(6034):1190–1192.

[35] Cirz, R. T., Chin, J. K., Andes, D. R., de Crécy-Lagard, V., Craig, W. A. and Romesberg, F. E. (2005). Inhibition of mutation and combating the evolution of antibiotic resistance. PLoS Biology, 3(6):1024–1033.

[36] Cisneros-Farrar, F. and Parsons, L. C. (2007). Antimicrobials: Classifications and uses in critical care. Critical Care Nursing Clinics of North America, 19(1):43–51.

[37] Claudi, B., Spröte, P., Chirkova, A., Personnic, N., Zankl, J. and Schürmann, N. (2014). Phenotypic variation of Salmonella in host tissues delays eradication by antimicrobial chemotherapy. Cell, 158:722–733.

[38] Cochran, J. W. and Byrne, R. W. (1974). Isolation and properties of a ribosome-bound factor required for ppGpp and pppGpp synthesis in Escherichia coli. The Journal of Biological Chemistry, 249(2):353–360.

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

[40] Corvec, S., Caroff, N., Espaze, E., Marraillac, J. and Reynaud, A. (2002). -11 Mutation in the ampC promoter increasing resistance to β-lactams in a clinical Escherichia coli strain. Antimicrobial Agents and Chemotherapy, 46(10):3265–3267.

[41] Dalton, T., Cegielski, P., Akksilp, S., Asencios, L., Caoili, J. C., Cho, S. N., Erokhin, V. V., Ershova, J., Gler, M. T., Kazennyy, B. Y., Kim, H. J., Kliiman, K., Kurbatova, E., Kvasnovsky, C., Leimane, V., Van Der Walt, M., Via, L. E., Volchenkov, G. V., Yagui, M. A. and Kang, H. (2012). Prevalence of and risk factors for resistance to second-line drugs in people with multidrug-resistant tuberculosis in eight countries: A prospective cohort study. The Lancet, 380(12):1406–1417.

[42] Datsenko, K. A. and Wanner, B. L. (2000). One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proceedings of the National Academy of Sciences of the United States of America, 97(12):6640–6645.

[43] Davies, J. and Davies, D. (2010). Origins and evolution of antibiotic resistance. Microbiology and Molecular Biology Reviews, 74(3):417–433.

[44] De Groote, V. N., Fauvart, M., Kint, C. I., Verstraeten, N., Jans, A., Cornelis, P. and Michiels, J. (2011). Pseudomonas aeruginosa fosfomycin resistance mechanisms affect non-inherited fluoroquinolone tolerance. Journal of Medical Microbiology, 60:329–336.

[45] Ding, X., Baca-DeLancey, R. R. and Rather, P. N. (2001). Role of SspA in the density-dependent expression of the transcriptional activator AarP in Providencia stuartii. FEMS Microbiology Letters, 196:25–29.

[46] Dörr, T., Lewis, K. and Vulic, M. (2009). SOS response induces persistence to fluoroquinolones in Escherichia coli. PLoS Genetics, 5(12):1–9.

[47] Dörr, T., Vulic, M. and Lewis, K. (2010). Ciprofloxacin causes persister formation by inducing the TisB toxin in Escherichia coli. PLoS Biology, 8(2):1–8.

[48] Doughtery, T. J. and Pucci, M. J. (2012). Antibiotic discovery and development. Springer.

[49] Drlica, K. (2003). The mutant selection window and antimicrobial resistance. Journal of Antimicrobial Chemotherapy, 52(1):11–17.

[50] Drlica, K., Malik, M., Kerns, R. J. and Zhao, X. (2008). Quinolone-mediated bacterial death. Antimicrobial Agents and Chemotherapy, 52(2):385–392.

[51] Durfee, T., Hansen, A., Zhi, H., Blattner, F. R. and Jin, D. J. (2008). Transcription profiling of the stringent response in Escherichia coli. Journal of Bacteriology, 190(3):1084–1096.

[52] Dykhuizen, D. E. and Hartl, D. L. (1983). Selection in chemostats. Microbiology Reviews, 47(2):150– 168.

[53] Elowitz, M. B., Siggia, E. D., Levine, A. J. and Swain, P. S. (2002). Stochastic gene expression in a single cell. Science, 297:1183–1187.

[54] Eng, R. H. K., Padberg, F. T., Smith, S. M., Tan, E. N. and Cherubin, C. E. (1991). Bactericidal effects of antibiotics on slowly growing and nongrowing bacteria. Antimicrobial Agents Chemotherapy, 35(9):1824–1828.

[55] Fang, F. C. (2013). Antibiotic and ROS linkage questioned. Nature Biotechnology, 31(5):415–416.

[56] Fasani, R. A. and Savageau, M. A. (2013). Molecular mechanisms of multiple toxin-antitoxin systems are coordinated to govern the persister phenotype. Proceedings of the National Academy of Sciences of the United States of America, 110(27):E2528–E2537.

[57] Fauvart, M., De Groote, V. N. and Michiels, J. (2011). Role of persister cells in chronic infections: clinical relevance and perspectives on anti-persister therapies. Journal of Medical Microbiology, 60:699–709.

[58] Feng, J., Kessler, D. A., Ben-jacob, E. and Levine, H. (2013). Growth feedback as a basis for

persister bistability. Proceedings of the National Academy of Sciences, 111(1):544–549.

[59] Ferenci, T. (2008). Bacterial physiology, regulation and mutational adaptation in a chemostat environment. Advances in Microbial Physiology, 53:169–230.

[60] Frank, E. G., Ennis, D. G., Gonzalez, M., Levine, A. S. and Woodgate, R. (1996). Regulation of SOS mutagenesis by proteolysis. Proceedings of the National Academy of Sciences of the United States of America, 93(19):10291–10296.

[61] Fridman, O., Goldberg, A., Ronin, I., Shoresh, N. and Balaban, N. Q. (2014). Optimization of lag time underlies antibiotic tolerance in evolved bacterial populations. Nature, 513:418–421.

[62] Friedman, S., Lu, T. and Drlica, K. (2001). Mutation in the DNA gyrase A gene of Escherichia coli that expands the quinolone resistance-determining region. Antimicrobial Agents and Chemotherapy, 45(8):2378-2380.

[63] Fu, H., Yuan, J. and Gao, H. (2015). Microbial oxidative stress response: Novel insights from environmental facultative anaerobic bacteria. Archives of Biochemistry and Biophysics, 584:28–35.

[64] Fung, D. K. C., Chan, E. W. C., Chin, M. L. and Chan, R. C. Y. (2010). Delineation of a bacterial starvation stress response network which can mediate antibiotic tolerance development. Antimicrobial Agents and Chemotherapy, 54(3):1082–1093.

[65] Gardete, S. and Tomasz, A. (2014). Mechanisms of vancomycin resistance in Staphylococcus aureus. The Journal of Clinical Investigation, 124(7):2836–2840.

[66] Gardner, A., West, S. A. and Griffin, A. S. (2007). Is bacterial persistence a social trait? PLoS ONE, 2(8):e752.

[67] Gefen, O. and Balaban, N. Q. (2009). The importance of being persistent: heterogeneity of bacterial populations under antibiotic stress. FEMS Microbiology Reviews, 33:704–717.

[68] Geli, P., Laxminarayan, R., Dunne, M. and Smith, D. L. (2012). “One-Size-Fits-All”? Optimizing treatment duration for bacterial infections. PLoS ONE, 7(1):e29838.

[69] Gerdes, K. and Maisonneuve, E. (2012). Bacterial persistence and toxin-antitoxin loci. Annual Review of Microbiology, 66(1):103–123.

[70] Germain, E., Castro-Roa, D., Zenkin, N. and Gerdes, K. (2013). Molecular mechanism of bacterial persistence by HipA. Molecular Cell, 52(2):248–254.

[71] Gerrish, P. J. and Lenski, R. E. (1998). The fate of competing beneficial mutations in an asexual population. Genetica, 102/103:127–144.

[72] Gibson, J. L., Lombardo, M. J., Thornton, P. C., Hu, K. H., Galhardo, R. S., Beadle, B., Habib, A., Magner, D. B., Frost, L. S., Herman, C., Hastings, P. J. and Rosenberg, S. M. (2010). The σE-stress response is required for stress-induced mutation and amplification in Escherichia coli. Molecular Microbiology, 77(2):415–430.

[73] Gonzalez, M., Frank, E. G., Levine, a. S. and Woodgate, R. (1998). Lon-mediated proteolysis of the Escherichia coli UmuD mutagenesis protein: in vitro degradation and identification of residues required for proteolysis. Genes & Development, 12(24):3889–3899.

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

[75] Gullberg, E., Cao, S., Berg, O. G., Ilbäck, C., Sandegren, L., Hughes, D. and Andersson, D. I. (2011). Selection of resistant bacteria at very low antibiotic concentrations. PLoS Pathogens, 7(7):e1002158.

[76] Gutierrez, A., Laureti, L., Crussard, S., Abida, H., Rodríguez-Rojas, A., Blázquez, J., Baharoglu, Z., Mazel, D., Darfeuille, F., Vogel, J. and Matic, I. (2013). β-lactam antibiotics promote bacterial mutagenesis via an RpoS-mediated reduction in replication fidelity. Nature Communications, 4:1–9.

[77] Hall, B. M., Ma, C.-X., Liang, P. and Singh, K. K. (2009). Fluctuation AnaLysis CalculatOR: a web tool for the determination of mutation rate using Luria-Delbruck fluctuation analysis. Bioinformatics, 25(12):1564–1565.

[78] Hansen, S., Lewis, K. and Vulic, M. (2008). Role of global regulators and nucleotide metabolism in antibiotic tolerance in Escherichia coli. Antimicrobial Agents and Chemotherapy, 52(8):2718–2726.

[79] Helaine, S., Cheverton, A. M., Watson, K. G., Faure, L. M., Matthews, S. A. and Holden, D. W. (2014). Internalization of Salmonella by macrophages induces formation of nonreplicating persisters. Science, 343:204–208.

[80] Henderson-Begg, S. K., Livermode, D. and Hall, L. (2006). Effect of subinhibitory concentrations of antibiotics on mutation frequency in Streptococcus pneumoniae. Journal of Antimicrobial Chemotherapy, 57(5):849–854.

[81] Hirsch, M. and Elliott, T. (2002). Role of ppGpp in rpoS stationary-phase regulation in Escherichia coli. Journal of Bacteriology, 184(18):5077–5087.

[82] Hofsteenge, N., van Nimwegen, E. and Silander, O. K. (2013). Quantitative analysis of persister fractions suggests different mechanisms of formation among environmental isolates of E. coli. BMC microbiology, 13(1):25.

[83] Holden, B. D. W. (2015). Persisters unmasked. Science, 347:30–32.

[84] Hong, S. H., Wang, X., O’Connor, H. F., Benedik, M. J. and Wood, T. K. (2012). Bacterial persistence increases as environmental fitness decreases. Microbial Biotechnology, 5(4):509–522.

[85] Hu, Y. and Coates, A. R. M. (2005). Transposon mutagenesis identifies genes which control antimicrobial drug tolerance in stationary-phase Escherichia coli. FEMS Microbiology Letters, 243(1):117– 124.

[86] Jacoby, G. (2005). Mechanisms of resistance to quinolones. Clinical Infectious Diseases, 41:S120–126.

[87] James, T. W. (1961). Continuous culture of microorganisms. Annual Review of Microbiology, 15:27– 46.

[88] Jansen, G., Barbosa, C. and Schulenburg, H. (2013). Experimental evolution as an efficient tool to dissect adaptive paths to antibiotic resistance. Drug Resistance Updates, 16(6):96–107.

[89] Jaurin, B., Grundström, T. and Normark, S. (1982). Sequence elements determining ampC promoter strength in E. coli. The EMBO Journal, 1(7):875–881.

[90] Jayaraman, R. (2008). Bacterial persistence: some new insights into an old phenomenon. Journal of Biosciences, 33:795–805.

[91] Johnson, P. J. T. and Levin, B. R. (2013). Pharmacodynamics, population dynamics, and the evolution of persistence in Staphylococcus aureus. PLoS Genetics, 9(1):e1003123.

[92] Jørgensen, K. M., Wassermann, T., Jensen, P. Ø., Hengzuang, W., Molin, S., Høiby, N. and Ciofu, O. (2013). Sublethal ciprofloxacin treatment leads to rapid development of high-level ciprofloxacin resistance during long term experimental evolution of Pseudomonas aeruginosa. Antimicrobial Agents and Chemotherapy, 57(9):4215–4121.

[93] Kaiser, P., Regoes, R. R., Dolowschiak, T., Wotzka, S. Y., Lengefeld, J., Slack, E., Grant, A. J., Ackermann, M. and Hardt, W. (2014). Cecum lymph node dendritic cells harbor slow-growing bacteria phenotypically tolerant to antibiotic treatment. PLoS Biology, 12(2):e1001793.

[94] Kaspy, I., Rotem, E., Weiss, N., Ronin, I., Balaban, N. Q. and Glaser, G. (2013). HipA-mediated antibiotic persistence via phosphorylation of the glutamyl-tRNA-synthetase. Nature Communications, 4:3001.

[95] Kawecki, T. J., Lenski, R. E., Ebert, D., Hollis, B., Olivieri, I. and Whitlock, M. C. (2012). Experimental evolution. Trends in Ecology and Evolution, 27(10):547–560.

[96] Kayama, S., Murakami, K., Ono, T., Ushimaru, M., Yamamoto, A., Hirota, K. and Miyake, Y. (2009). The role of rpoS gene and quorum-sensing system in ofloxacin tolerance in Pseudomonas aeruginosa. FEMS Microbiology Letters, 298(2):184–192.

[97] Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock, S., Buxton, S., Cooper, A., Markowitz, S., Duran, C., Thierer, T., Ashton, B., Meintjes, P. and Drummond, A. (2012). Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics, 28(12):1647–1649.

[98] Keren, I., Minami, S., Rubin, E. and Lewis, K. C. (2011). Characterization and transcriptome analysis of Mycobacterium tuberculosis persisters. mBio, 2(3):e00100–11.

[99] Keren, I., Kaldalu, N., Spoering, A., Wang, Y. and Lewis, K. (2004). Persister cells and tolerance to antimicrobials. FEMS Microbiology Letters, 230(1):13–18.

[100] Keren, I., Shah, D., Spoering, A., Kaldalu, N. and Lewis, K. (2004). Specialized persister cells and the mechanism of multidrug tolerance in Escherichia coli. Journal of Bacteriology, 186(24):8172– 8180.

[101] Keren, I., Wu, Y., Inocencio, J., Mulcahy, L. R. and Lewis, K. (2013). Killing by bactericidal antibiotics does not depend on reactive oxygen species. Science, 339:1213–1216.

[102] Kim, D. Y. (2015). Two stress sensor proteins for the expression of σE-regulon: DegS and RseB. Journal of Microbiology, 53(5):306–310.

[103] Kim, J.-S., Heo, P., Yang, T.-J., Lee, K.-S., Jin, Y.-S., Kim, S.-K., Shin, D. and Kweon, D.-H. (2011). Bacterial persisters tolerate antibiotics by not producing hydroxyl radicals. Biochemical and Biophysical Research Communications, 413(1):105–110.

[104] Kim, Y. and Wood, T. K. (2010). Toxins Hha and CspD and small RNA regulator Hfq are involved in persister cell formation through MqsR in Escherichia coli. Biochemical and Biophysical Research Communications, 391(1):209–213.

[105] Kint, C. I., Verstraeten, N., Fauvart, M. and Michiels, J. (2012). New-found fundamentals of bacterial persistence. Trends in Microbiology, 20(12):577–585.

[106] Kivisaar, M. (2003). Stationary phase mutagenesis: Mechanisms that accelerate adaptation of microbial populations under environmental stress. Environmental Microbiology, 5(10):814–827.

[107] Klemm, E. J., Gkrania-Klotsas, E., Hadfield, J., Forbester, J. L., Harris, S. R., Hale, C., Heath, J. N., Wileman, T., Clare, S., Kane, L., Goulding, D., Otto, T. D., Kay, S., Doffinger, R., Cooke, F. J., Carmichael, A., Lever, A. M. L., Parkhill, J., MacLennan, C. A., Kumararatne, D., Dougan, G. and Kingsley, R. A. (2016). Emergence of host-adapted Salmonella Enteritidis through rapid evolution in an immunocompromised host. Nature Microbiology, 1:15023.

[108] Koh, R. S. and Dunlop, M. J. (2012). Modeling suggests that gene circuit architecture controls phenotypic variability in a bacterial persistence network. BMC Systems Biology, 6(1):47.

[109] Kohanski, M. A., DePristo, M. A. and Collins, J. J. (2010). Sublethal antibiotic treatment leads to multidrug resistance via radical-induced mutagenesis. Molecular Cell, 37(3):311–320.

[110] Kohanski, M. A., Dwyer, D. J. and Collins, J. J. (2010). How antibiotics kill bacteria: from targets to networks. Nature Reviews Microbiology, 8(6):423–435.

[111] Kohanski, M. A., Dwyer, D. J., Hayete, B., Lawrence, C. A. and Collins, J. J. (2007). A common mechanism of cellular death induced by bactericidal antibiotics. Cell, 130(5):797–810.

[112] Korch, S. B., Henderson, T. A. and Hill, T. M. (2003). Characterization of the hipA7 allele of Escherichia coli and evidence that high persistence is governed by (p)ppGpp synthesis. Molecular Microbiology, 50(4):1199–1213.

[113] Korch, S. B. and Hill, T. M. (2006). Ectopic overexpression of wild-type and mutant hipA genes in Escherichia coli: Effects on macromolecular synthesis and persister formation. Journal of Bacteriology, 188(11):3826–3836.

[114] Kussell, E., Kishony, R., Balaban, N. Q. and Leibler, S. (2005). Bacterial persistence: A model of survival in changing environments. Genetics, 169:1807–1814.

[115] Kussell, E. and Leibler, S. (2005). Phenotypic diversity, population growth, and information in fluctuating environments. Science, 309:2075–2078.

[116] Kwan, B. W., Valenta, J. A., Benedik, M. J. and Wood, T. K. (2013). Arrested protein synthesis increases persister-like cell formation. Antimicrobial Agents and Chemotherapy, 57(3):1468–1473.

[117] Lachmann, M. and Jablonka, E. (1996). The inheritance of phenotypes: an adaptation to fluctuating environments. Journal of Theoretical Biology, 181(1):1–9.

[118] LaFleur, M. D., Qi, Q. and Lewis, K. (2010). Patients with long-term oral carriage harbor highpersister mutants of Candida albicans. Antimicrobial Agents and Chemotherapy, 54(1):39–44.

[119] Lange, R. and Hengge-Aronis, R. (1994). The cellular concentration of the σS-subunit of RNA polymerase in Escherichia coli is controlled at the levels of transcription, translation, and protein stability. Genes & Development, 8(13):1600–1612.

[120] Lázár, V., Pal Singh, G., Spohn, R., Nagy, I., Horváth, B., Hrtyan, M., Busa-Fekete, R., Bogos, B., Méhi, O., Csörgo, B., Pósfai, G., Fekete, G., Szappanos, B., Kégl, B., Papp, B. and Pál, C. (2013). Bacterial evolution of antibiotic hypersensitivity. Molecular Systems Biology, 9:700.

[121] Lee, H. H., Molla, M. N., Cantor, C. R. and Collins, J. J. (2010). Bacterial charity work leads to population-wide resistance. Nature, 467(7311):82–85.

[122] Lenski, R. E., Rose, M. R., Simpson, S. C. and Tadler, S. C. (1991). Long-term experimental evolution in Escherichia coli. I. Adaptation and divergence during 2,000 generations. The American Naturalist, 138(6):1315–1341.

[123] Lenski, R. E. (2011). Evolution in action: a 50,000-generation salute to Charles Darwin. Microbe, 6(1):30–33.

[124] Leung, V. and Lévesque, C. M. (2012). A stress-inducible quorum-sensing peptide mediates the formation of persister cells with noninherited multidrug tolerance. Journal of Bacteriology, 194(9):2265– 2274.

[125] Levin, B. R., Concepción-Acevedo, J. and Udekwu, K. I. (2014). Persistence: A copacetic and parsimonious hypothesis for the existence of non-inherited resistance to antibiotics. Current Opinionin Microbiology, 21:18–21.

[126] Levin, B. R. and Rozen, D. E. (2006). Non-inherited antibiotic resistance. Nature Reviews Microbiology, 4:556–562.

[127] Levy, S. B. and Marshall, B. (2004). Antibacterial resistance worldwide: causes, challenges and responses. Nature Medicine Supplement, 10(12):S122–S129.

[128] Lewin, C. S., Howard, B. M. A., Ratcliffe, N. T. and Smith, J. T. (1989). 4-Quinolones and the SOS response. Journal of Medical Microbiology, 29(2):139–144.

[129] Lewis, K. (2007). Persister cells, dormancy and infectious disease. Nature Reviews Microbiology, 5:48–56.

[130] Lewis, K. (2010). Persister cells. Annual Review of Microbiology, 64:357–372.

[131] Li, X.-Z. and Nikaido, H. (2009). Efflux-mediated drug resistance in bacteria: an update. Drugs, 69(12):1555–1623.

[132] Li, Y. and Zhang, Y. (2007). PhoU is a persistence switch involved in persister formation and tolerance to multiple antibiotics and stresses in Escherichia coli. Antimicrobial Agents and Chemotherapy, 51(6):2092–2099.

[133] Lieberman, T. D., Michel, J.-B., Aingaran, M., Potter-Bynoe, G., Roux, D., Davis, M. R., Skurnik, D., Leiby, N., LiPuma, J. J., Goldberg, J. B., McAdam, A. J., Priebe, G. P. and Kishony, R. (2011). Parallel bacterial evolution within multiple patients identifies candidate pathogenicity genes. Nature Genetics, 43(12):1275–1280.

[134] Lin, D., Gibson, I. B., Moore, J. M., Thornton, P. C., Leal, S. M. and Hastings, P. J. (2011). Global chromosomal structural instability in a subpopulation of starving Escherichia coli cells. PLoS Genetics, 7(8):e1002223.

[135] Liu, A., Fong, A., Becket, E., Yuan, J., Tamae, C., Medrano, L., Maiz, M., Wahba, C., Lee, C., Lee, K., Tran, K. P., Yang, H., Hoffman, R. M., Salih, A. and Miller, J. H. (2011). Selective advantage of resistant strains at trace levels of antibiotics: a simple and ultrasensitive color test for detection of antibiotics and genotoxic agents. Antimicrobial Agents and Chemotherapy, 55(3):1204–1210.

[136] Liu, H. and Tomasz, A. (1985). Penicillin tolerance in multiply drug-resistant neutral isolates of Streptococcus pneumoniae. Journal of Infectious Diseases, 152(2):365–372.

[137] Liu, Y. and Imlay, J. A. (2013). Cell death from antibiotics without the involvement of reactive oxygen species. Science, 339:1210–1213.

[138] Loewe, L., Textor, V. and Scherer, S. (2003). High deleterious genomic mutation rate in stationary phase of Escherichia coli. Science, 302:1558–1560.

[139] López, E., Elez, M., Matic, I. and Blázquez, J. (2007). Antibiotic-mediated recombination: ciprofloxacin stimulates SOS-independent recombination of divergent sequences in Escherichia coli. Molecular Microbiology, 64(1):83–93.

[140] Lou, C., Li, Z. and Ouyang, Q. (2008). A molecular model for persister in E. coli. Journal of Theoretical Biology, 255:205–209.

[141] Louie, A., Brown, D. L., Liu, W., Kulawy, R. W., Deziel, M. R. and Drusano, G. L. (2007). In vitro infection model characterizing the effect of efflux pump inhibition on prevention of resistance to levofloxacin and ciprofloxacin in Streptococcus pneumoniae. Antimicrobial Agents and Chemotherapy, 51(11):3988–4000.

[142] Luria, S. and Delbrück, M. (1943). Mutations of bacteria from virus sensitivity to virus resistance. Genetics, 28(6):491–511.

[143] Maddamsetti, R., Lenski, R. E. and Barrick, J. E. (2015). Adaptation, clonal interference, and frequency-dependent interactions in a long-term evolution experiment with Escherichia coli. Genetics, 200(2):619–631.

[144] Magnusson, L. U., Farewell, A. and Nyström, T. (2005). ppGpp: a global regulator in Escherichia coli. Trends in Microbiology, 13(5):236–242.

[145] Maisonneuve, E., Castro-Camargo, M. and Gerdes, K. (2013). (p)ppGpp controls bacterial persistence

by stochastic induction of toxin-antitoxin activity. Cell, 154(5):1140–1150.

[146] Maisonneuve, E. and Gerdes, K. (2014). Molecular mechanisms underlying bacterial persisters. Cell, 157(3):539–548.

[147] Maisonneuve, E., Shakespeare, L. J., Jørgensen, M. G. and Gerdes, K. (2011). Bacterial persistence by RNA endonucleases. Proceedings of the National Academy of Sciences of the United States of America, 108(32):13206–13211.

[148] Mayers, D. L. (2009). Antimicrobial Drug Resistance, volume 1. Humana Press.

[149] McClure, W. R. and Cech, C. L. (1978). On the mechanism of rifampicin inhibition of RNA synthesis. Journal of Biological Chemistry, 253(24):8949–8956.

[150] McDermott, W. (1958). Microbial Persistence. The Yale Journal of Biology and Medicine, 30:257– 291.

[151] McKenzie, G. J., Harris, R. S., Lee, P. L. and Rosenberg, S. M. (2000). The SOS response regulates adaptive mutation. Proceedings of the National Academy of Sciences of the United States of America, 97(12):6646–6651.

[152] Mechler, L., Herbig, A., Paprotka, K., Fraunholz, M., Nieselt, K. and Bertram, R. (2015). A novel point mutation promotes growth phase-dependent daptomycin tolerance in Staphylococcus aureus. Antimicrobial Agents and Chemotherapy, 59:5366–5376.

[153] Medaney, F., Dimitriu, T., Ellis, R. J. and Raymond, B. (2015). Live to cheat another day: bacterial dormancy facilitates the social exploitation of β-lactamases. The ISME Journal, 10(3):778–787.

[154] Meredith, H. R., Srimani, J. K., Lee, A. J., Lopatkin, A. J. and You, L. (2015). Collective antibiotic tolerance: mechanisms, dynamics and intervention. Nature Chemical Biology, 11(3):182–188.

[155] Michel, B. (2005). After 30 years of study, the bacterial SOS response still surprises us. PLoS Biology, 3(7):1174–1176.

[156] Michiels, J. and Fauvart, M. (2016). Bacterial persistence - Methods and protocols. Springer Science+ Business Media, New York.

[157] Mika, F. and Hengge, R. (2005). A two-component phosphotransfer network involving ArcB, ArcA, and RssB coordinates synthesis and proteolysis of σS (RpoS) in E. coli. Genes & Development, 19(22):2770–2781.

[158] Möker, N., Dean, C. R. and Tao, J. (2010). Pseudomonas aeruginosa increases formation of multidrug-tolerant persister cells in response to quorum-sensing signaling molecules. Journal of Bacteriology, 192(7):1946–1955.

[159] Montero, M., Rahimpour, M., Viale, A. M., Almagro, G., Eydallin, G., Sevilla, Á., Cánovas, M., Bernal, C., Lozano, A. B., Muñoz, F. J., Baroja-Fernández, E., Bahaji, A., Mori, H., Codoñer, F. M. and Pozueta-Romero, J. (2014). Systematic production of inactivating and non-inactivating suppressor mutations at the relA locus that compensate the detrimental effects of complete spoT loss and affect glycogen content in Escherichia coli. PLoS ONE, 9(9):e106938.

[160] Moyed, H. S. and Bertrand, K. P. (1983). hipA, a newly recognized gene of Escherichia coli K-12 that affects frequency of persistence after inhibition of murein synthesis. Microbiology, 155(2):768–775.

[161] Mulcahy, L. R., Burns, J. L., Lory, S. and Lewis, K. (2010). Emergence of Pseudomonas aeruginosa strains producing high levels of persister cells in patients with cystic fibrosis. Journal of Bacteriology, 192(23):6191–6199.

[162] Murakami, K., Ono, T., Viducic, D., Kayama, S., Mori, M., Hirota, K., Nemoto, K. and Miyake, Y. (2005). Role for rpoS gene of Pseudomonas aeruginosa in antibiotic tolerance. FEMS Microbiology Letters, 242(1):161–167.

[163] Nair, C. G., Chao, C., Ryall, B. and Williams, H. D. (2013). Sub-lethal concentrations of antibiotics increase mutation frequency in the cystic fibrosis pathogen Pseudomonas aeruginosa. Letters in Applied Microbiology, 56(2):149–154.

[164] Navarro Llorens, J. M., Tormo, A. and Martínez-García, E. (2010). Stationary phase in gramnegative bacteria. FEMS Microbiology Reviews, 34(4):476–495.

[165] Neu, H. and Reeves, D. (1986). Ciprofloxacin, volume 1. Springer.

[166] Newmark, K. G., O’Reilly, E. K., Pohlhaus, J. R. and Kreuzer, K. N. (2005). Genetic analysis of the requirements for SOS induction by nalidixic acid in Escherichia coli. Gene, 356:69–76.

[167] Nguyen, D., Joshi-Datar, A., Lepine, F., Bauerle, E., Olakanmi, O., Beer, K., McKay, G., Siehnel, R., Schafhauser, J., Wang, Y., Britigan, B. E. and Singh, P. K. (2011). Active starvation responses mediate antibiotic tolerance in biofilms and nutrient-limited bacteria. Science, 334(6058):982–6.

[168] Nilsson, O. (2012). Vancomycin resistant enterococci in farm animals – occurrence and importance. Infection Ecology & Epidemiology, 2:16959.

[169] Novak, R., Henriques, B., Charpentier, E., Normark, S. and Tuomanen, E. (1999). Emergence of vancomycin tolerance in Streptococcus pneumoniae. Nature, 399:590–593.

[170] Ochman, H. and Selander, R. (1984). Standard reference strains of Escherichia coli from natural populations. Journal of Bacteriology, 157(2):690–693.

[171] Orman, M. A. and Brynildsen, M. P. (2013). Dormancy is not necessary or sufficient for bacterial persistence. Antimicrobial Agents and Chemotherapy, 57(7):3230–3239.

[172] Oz, T., Guvenek, A., Yildiz, S., Karaboga, E., Tamer, Y. T., Mumcuyan, N., Ozan, V. B., Senturk, G. H., Cokol, M., Yeh, P. and Toprak, E. (2014). Strength of selection pressure is an important parameter contributing to the complexity of antibiotic resistance evolution. Molecular Biology and Evolution, 31(9):2387–2401.

[173] Pál, C., Papp, B. and Lázár, V. (2015). Collateral sensitivity of antibiotic-resistant microbes. Trends in Microbiology, 23(7):401–407.

[174] Palmer, K. L., Kos, V. N. and Gilmore, M. S. (2010). Horizontal gene transfer and the genomics of enterococcal antibiotic resistance. Current Opinion in Microbiology, 13(5):632–639.

[175] Patra, P. and Klumpp, S. (2013). Population dynamics of bacterial persistence. PLoS ONE, 8(5):e62814.

[176] Pearl, S., Gabay, C., Kishony, R., Oppenheim, A. and Balaban, N. Q. (2008). Nongenetic individuality in the host-phage interaction. PLoS Biology, 6(5):0957–0964.

[177] Pedersen, K., Christensen, S. K. and Gerdes, K. (2002). Rapid induction and reversal of a bacteriostatic condition by controlled expression of toxins and antitoxins. Molecular Microbiology, 45(2):501–510.

[178] Pennington, J. M. and Rosenberg, S. M. (2007). Spontaneous DNA breakage in single living Escherichia coli cells. Nature Genetics, 39(6):797–802.

[179] Piddock, L. J. and Walters, R. N. (1992). Bactericidal activities of five quinolones for Escherichia coli strains with mutations in genes encoding the SOS response or cell division. Antimicrobial Agents and Chemotherapy, 36(4):819–825.

[180] Pomposiello, P. J. and Demple, B. (2001). Redox-operated genetic switches: The SoxR and OxyR transcription factors. Trends in Biotechnology, 19(3):109–114.

[181] Ponder, R. G., Fonville, N. C. and Rosenberg, S. M. (2005). A switch from high-fidelity to error-prone DNA double-strand break repair underlies stress-induced mutation. Molecular Cell, 19(6):791–804.

[182] Poole, K. (2012). Stress responses as determinants of antimicrobial resistance in Gram-negative bacteria. Trends in Microbiology, 20(5):227–234.

[183] Raghavan, A. and Chatterji, D. (1998). Guanosine tetraphosphate-induced dissociation of open complexes at the Escherichia coli ribosomal protein promoters rplJ and rpsA P1: nanosecond depolarization spectroscopic studies. Biophysical Chemistry, 75(1):21–32.

[184] Ram, Y. and Hadany, L. (2014). Stress-induced mutagenesis and complex adaptation. Proceedings of the Royal Society B: Biological Sciences, 281(1792):45.

[185] Ramirez, M., Rajaram, S., Steininger, R. J., Osipchuk, D., Roth, M. A., Morinishi, L. S., Evans, L., Ji, W., Hsu, C.-H., Thurley, K., Wei, S., Zhou, A., Koduru, P. R., Posner, B. A., Wu, L. F. and Altschuler, S. J. (2016). Diverse drug-resistance mechanisms can emerge from drug-tolerant cancer persister cells. Nature Communications, 7:1–8.

[186] Rao, S. P. S., Alonso, S., Rand, L., Dick, T. and Pethe, K. (2008). The protonmotive force is required for maintaining ATP homeostasis and viability of hypoxic, nonreplicating Mycobacterium tuberculosis. Proceedings of the National Academy of Sciences of the United States of America, 105(33):11945–11950.

[187] Ratcliff, W. C. and Travisano, M. (2014). Experimental evolution of multicellular complexity in Saccharomyces cerevisiae. BioScience, 64(5):383–393.

[188] Regoes, R. R., Wiuff, C., Zappala, R. M., Garner, K. N., Baquero, F. and Levin, B. R. (2004). Pharmacodynamic functions: A multiparameter approach to the design of antibiotic treatment regimens. Antimicrobial Agents and Chemotherapy, 48(10):3670–3676.

[189] Renggli, S., Keck, W., Jenal, U. and Ritz, D. (2013). Role of autofluorescence in flow cytometric analysis of Escherichia coli treated with bactericidal antibiotics. Journal of Bacteriology, 195(18):4067– 4073.

[190] Rhodius, V. A., Suh, W. C., Nonaka, G., West, J. and Gross, C. A. (2006). Conserved and variable functions of the σS-stress response in related genomes. PLoS Biology, 4(1):0043–0059.

[191] Rodriguez, M. and Costa, S. (1999). Spontaneous kanamycin-resistant Escherichia coli mutant with altered periplasmic oligopeptide permease protein (OppA) and impermeability to aminoglycosides. Revista de Microbiologia, 30:153–156.

[192] Rosche, W. A. and Foster, P. L. (2000). Determining mutation rates in bacterial populations. Methods, 20:4–17.

[193] Rosenberg, S. M., Shee, C., Frisch, R. L. and Hastings, P. J. (2012). Stress-induced mutation via DNA breaks in Escherichia coli: A molecular mechanism with implications for evolution and medicine. Bioessays, 34:885–892.

[194] Rotem, E., Loinger, A., Ronin, I., Levin-Reisman, I., Gabay, C., Shoresh, N., Biham, O. and Balaban, N. Q. (2010). Regulation of phenotypic variability by a threshold-based mechanism underlies bacterial persistence. Proceedings of the National Academy of Sciences, 107(28):12541–12546.

[195] Rowley, G., Spector, M., Kormanec, J. and Roberts, M. (2006). Pushing the envelope: extracytoplasmic stress responses in bacterial pathogens. Nature Reviews Microbiology, 4(5):383–394.

[196] Schumacher, M. A., Piro, K. M., Xu, W., Hansen, S., Lewis, K. and Brennan, R. G. (2009). Molecular mechanisms of HipA-mediated multidrug tolerance and its neutralization by HipB. Science, 323:396– 401.

[197] Schweder, T., Lee, K. H., Lomovskaya, O. and Matin, A. (1996). Regulation of Escherichia coli starvation sigma factor (σS) by ClpXP protease. Journal of Bacteriology, 178(2):470–476.

[198] Shah, D., Zhang, Z., Khodursky, A. B., Kaldalu, N., Kurg, K. and Lewis, K. (2006). Persisters: a distinct physiological state of E. coli. BMC Microbiology, 6:53.

[199] Shan, Y., Lazinski, D., Rowe, S., Camilli, A. and Lewis, K. (2015). Genetic basis of persister tolerance to aminoglycosides in Escherichia coli. mBio, 6(2):e00078–15.

[200] Sharma, S. V., Lee, D. Y., Li, B., Quinlan, M. P., Takahashi, F., Maheswaran, S., McDermott, U., Azizian, N., Zou, L., Fischbach, M. A., Wong, K. K., Brandstetter, K., Wittner, B., Ramaswamy, S., Classon, M. and Settleman, J. (2010). A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell, 141(1):69–80.

[201] Simner, P. J., Zhanel, G. G., Pitout, J., Tailor, F., McCracken, M., Mulvey, M. R., Lagacé- Wiens, P. R., Adam, H. J. and Hoban, D. J. (2011). Prevalence and characterization of extendedspectrum β-lactamase– and AmpC β-lactamase–producing Escherichia coli: results of the CANWARD 2007–2009 study. Diagnostic Microbiology and Infectious Disease, 69(3):326–334.

[202] Smith, P. A. and Romesberg, F. E. (2007). Combating bacteria and drug resistance by inhibiting mechanisms of persistence and adaptation. Nature Chemical Biology, 3(9):549–556.

[203] Speer, B. S., Shoemaker, N. B. and Salyers, A. A. (1992). Bacterial resistance to tetracycline: Mechanisms, transfer, and clinical significance. Clinical Microbiology Reviews, 5(4):387–399.

[204] Spellberg, B. and Doi, Y. (2015). The rise of fluoroquinolone-resistant Escherichia coli in the community: Scarier than we thought. Journal of Infectious Diseases, 212:1853–1855.

[205] Spoering, A. L. and Lewis, K. (2001). Biofilms and planktonic cells of Pseudomonas aeruginosa have similar resistance to killing by antimicrobials. Journal of Bacteriology, 183(23):6746–6751.

[206] Stepanyan, K., Wenseleers, T., Duéñez-Guzmán, E. A., Muratori, F., Van den Bergh, B., Verstraeten, N., De Meester, L., Verstrepen, K. J., Fauvart, M. and Michiels, J. (2015). Fitness tradeoffs explain low levels of persister cells in the opportunistic pathogen Pseudomonas aeruginosa. Molecular Ecology, 24:1572–1583.

[207] Stewart, B. and Rozen, D. E. (2012). Genetic variation for antibiotic persistence in Escherichia coli. Evolution, 66(3):933–939.

[208] Strateva, T. and Yordanov, D. (2009). Pseudomonas aeruginosa - A phenomenon of bacterial resistance. Journal of Medical Microbiology, 58(9):1133–1148.

[209] Tam, V. H., Louie, A., Deziel, M. R., Liu, W. and Drusano, G. L. (2007). The relationship between quinolone exposures and resistance amplification is characterized by an inverted U: a new paradigm for optimizing pharmacodynamics to counterselect resistance. Antimicrobial Agents Chemotherapy, 51(2):744–777.

[210] Tan, C., Smith, R. P., Srimani, J. K., Riccione, K. A., Prasada, S., Kuehn, M. and You, L. (2012). The inoculum effect and band-pass bacterial response to periodic antibiotic treatment. Molecular Systems Biology, 8:617.

[211] Taylor, M. J., Tanna, S. and Sahota, T. (2010). Pharmacokinetics in drug discovery. Journal of Pharmaceutical Sciences, 99(10):4215–4227.

[212] Temkin, E., Adler, A., Lerner, A. and Carmeli, Y. (2014). Carbapenem-resistant Enterobacteriaceae: Biology, epidemiology, and management. Annals of the New York Academy of Sciences, 1323:22–42.

[213] Thattai, M. and Van Oudenaarden, A. (2004). Stochastic gene expression in fluctuating environments. Genetics, 167:523–530.

[214] Toprak, E., Veres, A., Michel, J.-B., Chait, R., Hartl, D. L. and Kishony, R. (2011). Evolutionary paths to antibiotic resistance under dynamically sustained drug selection. Nature Genetics, 44(1):101–105.

[215] Toprak, E., Veres, A., Yildiz, S., Pedraza, J. M., Chait, R., Paulsson, J. and Kishony, R. (2013). Building a morbidostat: an automated continuous-culture device for studying bacterial drug resistance under dynamically sustained drug inhibition. Nature Protocols, 8(3):555–567.

[216] Tracz, D. M., Boyd, D. A., Bryden, L., Hizon, R., Giercke, S., Caeseele, P. V. and Mulvey, M. R. (2005). Increase in ampC promoter strength due to mutations and deletion of the attenuator in a clinical isolate of cefoxitin-resistant Escherichia coli as determined by RT-PCR. Journal of Antimicrobial Chemotherapy, 55:768–772.

[217] Tsutsui, A., Suzuki, S., Yamane, K., Matsui, M., Konda, T., Marui, E., Takahashi, K. and Arakawa, Y. (2011). Genotypes and infection sites in an outbreak of multidrug-resistant Pseudomonas aeruginosa. Journal of Hospital Infection, 78(4):317–322.

[218] Tuomanen, E., Cozens, R., Tosch, W., Zak, O. and Tomasz, A. (1986). The rate of killing of Escherichia coli by β-lactam antibiotics is strictly proportional to the rate of bacterial growth. Journal of General Microbiology, 132(5):1297–1304.

[219] Tupin, A., Gualtieri, M., Roquet-Banères, F., Morichaud, Z., Brodolin, K. and Leonetti, J. P. (2010). Resistance to rifampicin: At the crossroads between ecological, genomic and medical concerns. International Journal of Antimicrobial Agents, 35(6):519–523.

[220] Úbeda, C., Maiques, E., Knecht, E., Lasa, Í., Novick, R. P. and Penadés, J. R. (2005). Antibioticinduced SOS response promotes horizontal dissemination of pathogenicity island-encoded virulence factors in staphylococci. Molecular Microbiology, 56(3):836–844.

[221] Van den Bergh, B., Michiels, J. E., Wenseleers, T., Windels, E. M., Vanden Boer, P., Kestemont, D., De Meester, L., Verstrepen, K. J., Verstraeten, N., Fauvart, M. and Michiels, J. (2016). Frequency of antibiotic application drives rapid evolutionary adaptation of Escherichia coli persistence. Nature Microbiology, 1(3):16020.

[222] Van Melderen, L. and Aertsen, A. (2009). Regulation and quality control by Lon-dependent proteolysis. Research in Microbiology, 160(9):645–651.

[223] Vàzquez-Laslop, N., Lee, H. and Neyfakh, A. A. (2006). Increased persistence in Escherichia coli caused by controlled expression of toxins or other unrelated proteins. Journal of Bacteriology, 188(10):3494–3497.

[224] Verstraeten, N., Knapen, W. J., Kint, C. I., Liebens, V., Van den Bergh, B., Dewachter, L., Michiels, J. E., Fu, Q., David, C. C., Fierro, A. C., Marchal, K., Beirlant, J., Versées, W., Hofkens, J., Jansen, M., Fauvart, M. and Michiels, J. (2015). Obg and membrane depolarization are part of a microbial bet-hedging strategy that leads to antibiotic tolerance. Molecular Cell, 59:1–13.

[225] Viducic, D., Ono, T., Murakami, K., Susilowati, H., Kayama, S., Hirota, K. and Miyake, Y. (2006). Functional analysis of spoT, relA and dksA genes on quinolone tolerance in Pseudomonas aeruginosa under nongrowing condition.

[226] Vogwill, T., Comfort, A. C., Furió, V. and MaClean, R. C. (2016). Persistence and resistance as complementary bacterial adaptations to antibiotics. Journal of Evolutionary Biology, n/a–n/a.

[227] Völzing, K. G. and Brynildsen, M. P. (2015). Stationary-phase persisters to ofloxacin sustain DNA damage and require repair systems only during recovery. mBio, 6(5):e00731–15.

[228] Wahl, L. M., Gerrish, P. J. and Saika-Voivod, I. (2002). Evaluating the impact of population bottlenecks in experimental evolution. Genetics, 162(2):961–971.

[229] Wakamoto, Y., Dhar, N., Chait, R., Schneider, K., Signorino-Gelo, F., Leibler, S. and McKinney, J. (2013). Dynamic persistence of antibiotic-stressed Mycobacteria. Science, 339:91–95.

[230] Walsh, N. P., Alba, B. M., Bose, B., Gross, C. A. and Sauer, R. T. (2003). OMP peptide signals initiate the envelope-stress response by activating DegS protease via relief of inhibition mediated by its PDZ domain. Cell, 113(1):61–71.

[231] Wayne, L. G. and Sohaskey, C. D. (2001). Nonreplicating persistence of Mycobacterium tuberculosis. Annual Review of Microbiology, 55:139–163.

[232] Weber, H., Polen, T., Heuveling, J., Wendisch, V. F. and Hengge, R. (2005). Genome-wide analysis of the general stress response network in Escherichia coli : σS-dependent genes, promoters, and sigma factor selectivity. Journal of Bacteriology, 187(5):1591–1603.

[233] WHO (2014). Antimicrobial resistance: Global report on surveillance 2014. World Health Organization.

[234] Wimberly, H., Shee, C., Thornton, P. C., Sivaramakrishnan, P., Rosenberg, S. M. and Hastings, P. J. (2013). R-loops and nicks initiate DNA breakage and genome instability in non-growing Escherichia coli. Nature Communications, 4:2115.

[235] Wolf, L. N. and Barrick, J. E. (2012). Tracking winners and losers in E. coli evolution experiments. Microbe, 7(3):124–128.

[236] Wong, A., Rodrigue, N. and Kassen, R. (2012). Genomics of adaptation during experimental evolution of the opportunistic pathogen Pseudomonas aeruginosa. PLoS Genetics, 8(9):e1002928.

[237] Wood, T. K., Knabel, S. J. and Kwan, B. W. (2013). Bacterial persister cell formation and dormancy. Applied and Environmental Microbiology, 79(23):7116–7121.

[238] Wu, N., Chen, B., Tian, S. and Chu, Y. (2014). The inoculum effect of antibiotics against CTXM- extended-spectrum β-lactamase-producing Escherichia coli. Annals of Clinical Microbiology and Antimicrobials, 13(1):45.

[239] Wu, N., He, L., Cui, P., Wang, W., Yuan, Y., Liu, S., Xu, T., Zhang, S., Wu, J., Zhang, W. and Zhang, Y. (2015). Ranking of persister genes in the same Escherichia coli genetic background demonstrates varying importance of individual persister genes in tolerance to different antibiotics. Frontiers in Microbiology, 6:1003.

[240] Wu, Y., Vulic, M., Keren, I. and Lewis, K. (2012). Role of oxidative stress in persister tolerance. Antimicrobial Agents and Chemotherapy, 56(9):4922–4926.

[241] Xiao, H., Kalman, M., Ikehara, K., Zemel, S., Glaser, G. and Cashel, M. (1991). Residual guanosine 3’,5’-bispyrophosphate synthetic activity of relA null mutants can be eliminated by spoT null mutations. Journal of Biological Chemistry, 266(9):5980–5990.

[242] Xiaodong, C. (2007). Exact stochastic simulation of coupled chemical reactions with delays. Journal of Chemical Physics, 126(12):124108.

[243] Xie, X. S., Choi, P. J., Li, G.-W., Lee, N. K. and Lia, G. (2008). Single-molecule approach to molecular biology in living bacterial cells. Annual Review of Biophysics, 37:417–44.

[244] Yang, L., Jelsbak, L., Marvig, R. L., Damkiaer, S., Workman, C. T., Rau, M. H., Hansen, S. K., Folkesson, A., Johansen, H. K., Ciofu, O., Hoiby, N., Sommer, M. O. A. and Molin, S. (2011). Evolutionary dynamics of bacteria in a human host environment. Proceedings of the National Academy of Sciences, 108(18):7481–7486.

[245] Zaslaver, A., Bren, A., Ronen, M., Itzkovitz, S., Kikoin, I., Shavit, S., Liebermeister, W., Surette, M. G. and Alon, U. (2006). A comprehensive library of fluorescent transcriptional reporters for Escherichia coli. Nature Methods, 3(8):623–628.

[246] Zhang, Q., Lambert, G., Liao, D., Kim, H., Robin, K., Tung, C.-K., Pourmand, N. and Austin, R. H. (2011). Acceleration of emergence of bacterial antibiotic resistance in connected microenvironments. Science, 333:1764–1767.

[247] Ziv, N., Brandt, N. J. and Gresham, D. (2013). The use of chemostats in microbial systems biology. Journal of Visualized Experiments, 80:e50168.

Universiteit of Hogeschool
Master of science in de bio-ingenieurswetenschappen: cel- & gentechnologie
Publicatiejaar
2016
Promotor(en)
Jan Michiels
Kernwoorden
Share this on: