1887

Abstract

This study investigates the link between adaptation to biocides and antibiotics in . An enrichment continuous culture of NCIMB 10421 (MIC 25 mg BKC l) was operated (=0.04 h, 792 h) with added benzalkonium chloride (BKC). A derivative, PA-29 (696 h), demonstrated a >12-fold decrease in sensitivity to the biocide (MIC >350 mg BKC l). The variant demonstrated a 256-fold increase in resistance to ciprofloxacin, with a mutation in the gene (Thr-83→Ile). Similarly, culturing of the original strain in a continuous-culture system with ciprofloxacin selection pressure led to the evolution of BKC-adapted populations (MIC 100 mg BKC l). Efflux pump activity predominantly contributed to the developed phenotype of PA-29. An amino acid substitution (Val-51→Ala) in , the Mex efflux system regulator gene, was observed for PA-29. Overexpression of both MexAB-OprM and MexCD-OprJ was recorded for PA-29. Similarly, , a repressor of the Mex system, was downregulated. Competition studies were carried out in continuous culture between PA-29 and the original strain (in the presence of subinhibitory concentrations of BKC). The outcome of competition was influenced by the concentration of biocide used and the nature of limiting nutrient. The inclusion of 1 mg BKC l in the medium feed was sufficient to select (=0.011) for the BKC-adapted strain in magnesium-limited culture. Conversely, the presence of 10 mg BKC l in the medium supply was insufficient to select for the same organism (=−0.017) in the glucose-limited culture. These results indicate the importance of environmental conditions on selection and maintenance of biocide adaptation.

Loading

Article metrics loading...

/content/journal/micro/10.1099/mic.0.029751-0
2010-01-01
2024-03-28
Loading full text...

Full text loading...

/deliver/fulltext/micro/156/1/30.html?itemId=/content/journal/micro/10.1099/mic.0.029751-0&mimeType=html&fmt=ahah

References

  1. Akimitsu N., Hamamoto H., Inoue R. I., Shoji M., Takemori K., Hamasaki N., Sekimizu K. 1999; Increase in resistance of methicillin-resistant Staphylococcus aureus to β-lactams caused by mutations conferring resistance to benzalkonium chloride, a disinfectant widely used in hospitals. Antimicrob Agents Chemother 43:3042–3043
    [Google Scholar]
  2. Anderson R. L., Carr J. H., Bond W. W., Favero M. S. 1997; Susceptibility of vancomycin-resistant enterococci to environmental disinfectants. Infect Control Hosp Epidemiol 18:195–199
    [Google Scholar]
  3. Chuanchuen R., Beinlich K., Hoang T. T., Becher A., Karkhoff-Schweizer R. R., Schweizer H. P. 2001; Cross-resistance between triclosan and antibiotics in Pseudomonas aeruginosa is mediated by multidrug efflux pumps: exposure of a susceptible mutant strain to triclosan selects nfxB mutants overexpressing MexCD-OprJ. Antimicrob Agents Chemother 45:428–432
    [Google Scholar]
  4. Chuanchuen R., Karkhoff-Schweizer R. R., Schweizer H. P. 2003; High-level triclosan resistance in Pseudomonas aeruginosa is solely a result of efflux. Am J Infect Control 31:124–127
    [Google Scholar]
  5. Cole E. C., Addison R. M., Rubino J. R., Leese K. E., Dulaney P. D., Newell M. S., Wilkins J., Gaber D. J., Wineinger T., Criger D. A. 2003; Investigation of antibiotic and antibacterial agent cross-resistance in target bacteria from homes of antibacterial product users and nonusers. J Appl Microbiol 95:664–676
    [Google Scholar]
  6. Dean C. R., Visalli M. A., Projan S. J., Sum P. E., Bradford P. A. 2003; Efflux-mediated resistance to tigecycline (GAR-936) in Pseudomonas aeruginosa PAO1. Antimicrob Agents Chemother 47:972–978
    [Google Scholar]
  7. Dykhuizen D. E. 1993; Chemostats used for studying natural selection and adaptive evolution. Methods Enzymol 224:613–631
    [Google Scholar]
  8. Fernandez-Astorga A., Hijarrubia M. J., Hernandez M., Arana I., Sunen E. 1995; Disinfectant tolerance and antibiotic resistance in psychrotrophic Gram-negative bacteria isolated from vegetables. Lett Appl Microbiol 20:308–311
    [Google Scholar]
  9. Fleming G., Patching J. W. 2008; The fermenter in research and development. In Practical Fermentation Technology pp 347–377 Edited by McNeil B., Harvey L. M. Chichester: Wiley;
    [Google Scholar]
  10. Fleming G. T., McCarthy D. M., Colombet N., Patching J. W. 2002; The effect of levofloxacin concentration on the development and maintenance of antibiotic-resistant clones of Escherichia coli in chemostat culture. J Ind Microbiol Biotechnol 29:155–162
    [Google Scholar]
  11. Fraise A. P. 2002; Biocide abuse and antimicrobial resistance – a cause for concern?. J Antimicrob Chemother 49:11–12
    [Google Scholar]
  12. Gilbert P., McBain A. J. 2003; Potential impact of increased use of biocides in consumer products on prevalence of antibiotic resistance. Clin Microbiol Rev 16:189–208
    [Google Scholar]
  13. Goldberg I., Er-el Z. 1981; The chemostat – an efficient technique for medium optimization. Process Biochem 16:2–4
    [Google Scholar]
  14. Higgins P. G., Fluit A. C., Milatovic D., Verhoef J., Schmitz F. J. 2003; Mutations in GyrA, ParC, MexR and NfxB in clinical isolates of Pseudomonas aeruginosa. Int J Antimicrob Agents 21:409–413
    [Google Scholar]
  15. Irizarry L., Merlin T., Rupp J., Griffith J. 1996; Reduced susceptibility of methicillin-resistant Staphylococcus aureus to cetylpyridinium chloride and chlorhexidine. Chemotherapy 42:248–252
    [Google Scholar]
  16. Jalal S., Wretlind B. 1998; Mechanisms of quinolone resistance in clinical strains of Pseudomonas aeruginosa. Microb Drug Resist 4:257–261
    [Google Scholar]
  17. Jones M. V., Herd T. M., Christie H. J. 1989; Resistance of Pseudomonas aeruginosa to amphoteric and quaternary ammonium biocides. Microbios 58:49–61
    [Google Scholar]
  18. Karatzas K. A. G., Webber M. A., Jorgensen F., Woodward M. J., Piddock L. J. V., Humphrey T. J. 2007; Prolonged treatment of Salmonella enterica serovar Typhimurium with commercial disinfectants selects for multiple antibiotic resistance, increased efflux and reduced invasiveness. J Antimicrob Chemother 60:947–955
    [Google Scholar]
  19. Langsrud S., Sundheim G., Borgmann-Strahsen R. 2003; Intrinsic and acquired resistance to quaternary ammonium compounds in food-related Pseudomonas spp. J Appl Microbiol 95:874–882
    [Google Scholar]
  20. Lee J. K., Lee Y. S., Park Y. K., Kim B. S. 2005; Alterations in the GyrA and GyrB subunits of topoisomerase II and the ParC and ParE subunits of topoisomerase IV in ciprofloxacin-resistant clinical isolates of Pseudomonas aeruginosa. Int J Antimicrob Agents 25:290–295
    [Google Scholar]
  21. Livak K. J., Schmittgen T. D. 2001; Analysis of relative gene expression data using real-time quantitative PCR and the method. Methods 25:402–408
    [Google Scholar]
  22. Lomovskaya O., Bostian K. A. 2006; Practical applications and feasibility of efflux pump inhibitors in the clinic – a vision for applied use. Biochem Pharmacol 71:910–918
    [Google Scholar]
  23. Lomovskaya O., Lee A., Hoshino K., Ishida H., Mistry A., Warren M. S., Boyer E., Chamberland S., Lee V. J. 1999; Use of a genetic approach to evaluate the consequences of inhibition of efflux pumps in Pseudomonas aeruginosa. Antimicrob Agents Chemother 43:1340–1346
    [Google Scholar]
  24. Loughlin M. F., Jones M. V., Lambert P. A. 2002; Pseudomonas aeruginosa cells adapted to benzalkonium chloride show resistance to other membrane-active agents but not to clinically relevant antibiotics. J Antimicrob Chemother 49:631–639
    [Google Scholar]
  25. Maniatis T. 1982; Appendix A. Biochemical techniques: liquid media. In Molecular Cloning: a Laboratory Manual p 440 Edited by Maniatis T., Fritsch E. F., Sambrook J. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory;
    [Google Scholar]
  26. Manzoor S. E., Lambert P. A., Griffiths P. A., Gill M. J., Fraise A. P. 1999; Reduced glutaraldehyde susceptibility in Mycobacterium chelonae associated with altered cell wall polysaccharides. J Antimicrob Chemother 43:759–765
    [Google Scholar]
  27. McBain A. J., Gilbert P. 2001; Biocide tolerance and the harbingers of doom. Int Biodeterior Biodegradation 47:55–61
    [Google Scholar]
  28. McMahon M. A. S., Blair I. S., Moore J. E., McDowell D. A. 2007; Habituation to sub-lethal concentrations of tea tree oil ( Melaleuca alternifolia) is associated with reduced susceptibility to antibiotics in human pathogens. J Antimicrob Chemother 59:125–127
    [Google Scholar]
  29. Méchin L., Dubois-Brissonnet F., Heyd B., Leveau J. Y. 1999; Adaptation of Pseudomonas aeruginosa ATCC 15442 to didecyldimethylammonium bromide induces changes in membrane fatty acid composition and in resistance of cells. J Appl Microbiol 86:859–866
    [Google Scholar]
  30. Moken M. C., McMurry L. M., Levy S. B. 1997; Selection of multiple-antibiotic-resistant ( mar) mutants of Escherichia coli by using the disinfectant pine oil: roles of the mar and acrAB loci. Antimicrob Agents Chemother 41:2770–2772
    [Google Scholar]
  31. NCCLS 2003 Methods for dilution antimicrobial susceptibility tests for bacteria that grow aerobically. Approved Standard M7–A6 Wayne, PA: National Committee for Clinical Laboratory Standards;
    [Google Scholar]
  32. Noguchi N., Hase M., Kitta M., Sasatsu M., Deguchi K., Kono M. 1999; Antiseptic susceptibility and distribution of antiseptic-resistance genes in methicillin-resistant Staphylococcus aureus. FEMS Microbiol Lett 172:247–253
    [Google Scholar]
  33. O'Reilly A., Smith P. 1999; Development of methods for predicting the minimum concentrations of oxytetracycline capable of exerting a selection for resistance to this agent. Aquaculture 180:1–11
    [Google Scholar]
  34. Piddock L. J. V. 2006; Multidrug-resistance efflux pumps are not just for resistance. Nat Rev Microbiol 4:629–636
    [Google Scholar]
  35. Poole K. 2005; Efflux-mediated antimicrobial resistance. J Antimicrob Chemother 56:20–51
    [Google Scholar]
  36. Price C. T. D., Singh V. K., Jayaswal R. K., Wilkinson B. J., Gustafson J. E. 2002; Pine oil cleaner-resistant Staphylococcus aureus: reduced susceptibility to vancomycin and oxacillin and involvement of SigB. Appl Environ Microbiol 68:5417–5421
    [Google Scholar]
  37. Russell A. D. 1999; Bacterial resistance to disinfectants: present knowledge and future problems. J Hosp Infect 43 :SupplementS57–S68
    [Google Scholar]
  38. Russell A. D. 2002; Introduction of biocides into clinical practice and the impact on antibiotic-resistant bacteria. Symp Ser Soc Appl Microbiol 31:121S–135S
    [Google Scholar]
  39. Russell A. D., Tattawasart U., Maillard J. Y., Furr J. R. 1998; Possible link between bacterial resistance and use of antibiotics and biocides. Antimicrob Agents Chemother 42:2151
    [Google Scholar]
  40. Rutala W. A., Stiegel M. M., Sarubbi F. A., Weber D. J. 1997; Susceptibility of antibiotic-susceptible and antibiotic-resistant hospital bacteria to disinfectants. Infect Control Hosp Epidemiol 18:417–421
    [Google Scholar]
  41. Savli H., Karadenizli A., Kolayli F., Gundes S., Ozbek U., Vahaboglu H. 2003; Expression stability of six housekeeping genes: a proposal for resistance gene quantification studies of Pseudomonas aeruginosa by real-time quantitative RT-PCR. J Med Microbiol 52:403–408
    [Google Scholar]
  42. Schwartz L. S., Jansen N. B., Ho N. W. Y., Tsao G. T. 1988; Plasmid instability kinetics of the yeast S288C pUCKm8 [Cir(+)] in non-selective and selective media. Biotechnol Bioeng 32:733–740
    [Google Scholar]
  43. Stecchini M. L., Manzano M., Sarais I. 1992; Antibiotic and disinfectant susceptibility in Enterobacteriaceae isolated from minced meat. Int J Food Microbiol 16:79–85
    [Google Scholar]
  44. Suller M. T. E., Russell A. D. 2000; Triclosan and antibiotic resistance in Staphylococcus aureus. J Antimicrob Chemother 46:11–18
    [Google Scholar]
  45. Tabata A., Nagamune H., Maeda T., Murakami K., Miyake Y., Kourai H. 2003; Correlation between resistance of Pseudomonas aeruginosa to quaternary ammonium compounds and expression of outer membrane protein OprR. Antimicrob Agents Chemother 47:2093–2099
    [Google Scholar]
  46. Thorrold C. A., Letsoalo M. E., Duse A. G., Marais E. 2007; Efflux pump activity in fluoroquinolone and tetracycline resistant Salmonella and Escherichia coli implicated in reduced susceptibility to household antimicrobial cleaning agents. Int J Food Microbiol 113:315–320
    [Google Scholar]
  47. To M. S., Favrin S., Romanova N., Griffiths M. W. 2002; Postadaptational resistance to benzalkonium chloride and subsequent physicochemical modifications of Listeria monocytogenes. Appl Environ Microbiol 68:5258–5264
    [Google Scholar]
http://instance.metastore.ingenta.com/content/journal/micro/10.1099/mic.0.029751-0
Loading
/content/journal/micro/10.1099/mic.0.029751-0
Loading

Data & Media loading...

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