Skip to main content

Advertisement

ADVERTISEMENT

Empirical Studies

Recurring and Antimicrobial-Resistant Infections: Considering the Potential Role of Biofilms in Clinical Practice

April 2007

  A biofilm colony is a complex, structured, interdependent community of micro-organisms enclosed in a self-produced polymeric matrix (the biofilm, frequently referred to as glycocalyx or slime). Biofilm is adherent to inert and living surfaces that have sufficient moisture and/or nutrients to sustain its survival.1,2 Like other infections, the biofilm colony may be a single species or a mixture of species of bacteria and/or fungi and may be different strains of the same species.3-7   Biofilm formation is not uncommon or limited to a small number of micro-organisms or tissues.1,8,9 The Centers for Disease Control and Prevention (CDC) has suggested that biofilms account for 80% of human infections.10 Alam et al11 screened 111 cultures of pus, exudate, joint aspirate, and blood obtained aseptically from cases of osteomyelitis and septic arthritis. Glycocalyx was found in 76.3% of isolates of Staphylococcus aureus, 57.1% of S. epidermidis, 50% of Pseudomonas aeruginosa, and 75% of Escherichia coli. Gristina et al12 examined tissues from biomaterials and prosthesis-related infection in 25 surgical patients in a general hospital setting and found that 76% of the causative bacteria grew in biofilms; 17 of these infections were associated with orthopaedic prostheses, 59% of which were in biofilms. However, organisms in biofilms do not always produce an infectious disease and are not always harmful. Current knowledge of biofilm development, resistance of micro-organisms to antibiotics and biocides, and issues related to culturing micro-organisms in biofilms is summarized to help clinicians improve clinical outcomes in soft tissue and bone infections and the treatment of wounds. A glossary of relevant terms (see “Glossary of Terms”) has been provided.

Biofilm Development

  Biofilms form on wet or moist surfaces. Biofilm formation begins immediately upon micro-organism contact with biologic tissue or a medical device4,13-21 (see Figure 1).

  Biologic tissue. Initially, organic molecules in tissue fluids form a layer on the tissues or medical device called a conditioning film. The different strains or species of micro-organisms in the immediate vicinity co-aggregate22; subsequently, cell-to-cell adhesion to the conditioning film occurs.13,23-25 With continued adhesion of micro-organisms, a multilayered colony of cells is formed. The colony of micro-organisms anchors firmly and is surrounded with a self-produced, glue-like slime matrix, comprised primarily of exopolysaccharide and some lipids, proteins, and nucleic acids.26-33 Once the colony is anchored, the process becomes irreversible.

  The biofilm has a rough, irregular surface that contains many individual colonies of non-uniform, mushroom-shaped or finger-like columns surrounded by fluid-filled channels in which nutrients, enzymes, and waste products circulate.34 Biofilms produced under different conditions differ in their cellular morphology and matrix content.35 The biofilm’s strength of attachment depends on its adhesion to the conditioning film.13,36 Human blood has been shown to enhance development of Gram-positive and Gram-negative bacterial biofilms37; heparin has been shown to promote S. aureus biofilm formation.38

  A mature biofilm may take a few hours or several weeks to fully develop. In one study,39 a methicillin-resistant S. aureus biofilm was found to be six cells thick and covered 10% of the surface of a silastic catheter after 2 hours. The biofilm may entrap minerals – mineral build-up is associated with catheter blockage.40

  Biofilm colonies form on traumatized or compromised living tissues and nonviable necrotic tissues such as burns,8 wounds and skin ulcers,41,42 and exposed or damaged tendon.43 Impetigo and furuncles have been identified as ideal surfaces for biofilm formation by contaminating and colonizing bacteria.44 Acute and chronic otitis media, chronic tonsillitis, osteomyelitis, bone fragments or sequestra, and exposed bone are susceptible tissues for biofilm formation.42,45,46 Biofilms are found on heart valves, dental enamel, intestinal mucosa,1 between toes and in armpits,47 and are associated with rheumatoid arthritis48 and genitourinary disease.49,50

  Medical devices. Because metal plates or screws, artificial joints, indwelling catheters, internal fixation devices, sutures,8,51 and other internal medical devices may be wet or moist surfaces, they are subject to biofilm formation. Biofilms also form on bone cement, even in the presence of antibiotics, including gentamicin52-55; they use these surfaces as a foundation to spread into adjacent tissues and form new colonies of biofilm-protected micro-organisms.12

  The effect of strains. Biofilm formation is strain dependent.56,57 Different strains of the same bacteria may develop biofilms of different consistencies. Generally, older biofilms are more developed, provide more protection, and their micro-organisms are more resistant to biocides, antibacterials, and conventional culturing than younger colonies.13,58

  Gene expression. Gene expression plays an important role in biofilm production, pathogenicity, and virulence.59-62 High population or critical cell population density-dependent processes known as quorum sensing provide bacteria the ability to communicate, coordinate, and respond to gene expression via signaling molecules once a threshold population density has been reached.63,64 Bacteria produce and detect signaling molecules (autoinducers) and use them to communicate cell-to-cell and coordinate gene expression (behavior). When bacteria multiply to a high population density (a quorum) and produce enough autoinducer (reach and sense a threshold level), the group responds by coordinating its gene expression with a population-wide gene expression, coordinating the behavior of the entire community of bacteria.65 For example, bacteria may colonize a wound, but they do not become a pathogenic-causing disease until they reach a certain population density such as 105 bacteria per gram of tissue. Micro-organisms use quorum sensing for diverse genetic cellular processes, including biofilm formation.57,64,66

  Survival. Micro-organisms within the biofilm may be in a low metabolic state but remain fit and able to survive under stress on a minimal amount of nutrients. An anerobic layer of bacteria may develop beneath the aerobic layer where oxygen levels are low enough to sustain their survival.

  Infection. Bacteria in biofilms are commonly responsible for recurring infections after repeated trials of antibiotics.1,67 The micro-organisms may be slow to produce clinical symptoms and may remain dormant for weeks or years before causing local or systemic signs and symptoms of infection. Because the biofilm is a dynamic environment,30,32,68,69 the integrity of part or all of the biofilm may fail at any time. The micro-organisms no longer within the biofilm may quickly multiply and disperse, causing rapid increases in bacterial counts and random showers of bacteria.70,71

  Detachment or separation and dispersal of bacteria from the biofilm, such as after an injury, can act as a nidus of infection.72,73 The micro-organisms that are now planktonic, or free-floating, may cause a bacteremia or a chronic recurring infection such as osteomyelitis, cellulitis, sinusitis, or urinary tract infections. Biofilm colonies are often polymicrobial – the same bacteria may not always cause the recurrent infection.

Biofilm Resistance

  Biofilms are complex and depending on a variety of factors, including bacterial strain and dosage of the antibiotic, may decrease or may have little or no impact on the effectiveness of antimicrobials.74-83 Appropriate antibiotics act on the bacteria outside the biofilm while the same species of bacteria inside the biofilm are effectively protected from most antimicrobials and the host’s defense mechanisms.

  MBC concentration. Bacteria inside the biofilm have a much higher minimum bactericidal concentration (MBC) than the same strain of bacteria outside the biofilm.72,77,84,85 Bacteria within a biofilm may require levels of antibiotics up to 5,000 times the MBC to kill all of their biofilm-protected bacteria compared to MBC levels that kill free-floating cells of the same strain.73 In a microbiological survey of automated water systems, Dreeszen86 reported bacteria in biofilms 3,000 times more resistant to free chlorine than the planktonic bacteria. This may overstate the clinical problem since most human infections are treated successfully with antimicrobials. However, achievable safe therapeutic levels of most antibiotics have been shown to be ineffective in killing most biofilm-protected bacteria.8,81,87

  The mechanism of resistance is complex and remains unclear but is known to be multifactorial and varies with the species and the strain of the micro-organism. Resistance appears to depend on multicellular synergistic behavior,88 the physiological characteristics of the biofilm itself, the age of the biofilm, altered physiology of cells within the biofilm, and phenotypic changes in the cells within the biofilm.4,9,89-91

  Synergistic behavior. Rather than conforming to individual, single-species bacterial behavior, micro-organisms in a biofilm, whether or not mixed species, communicate between and among the same and different species by cell-to-cell signaling, interacting with each other and their environment, sharing resources, and exhibiting multicellular synergistic behavior with common goals similar to a community.92-96 One of the goals of these communities is to lessen the effects of biocides and antimicrobials. This behavior cannot be achieved by individual single-species micro-organisms.32,92-94,97-101

  Charge and age. Other physiological characteristics of the biofilm contribute to resistance. The biofilm has a negative charge and may restrict diffusion of positive-charged antibiotics such as aminoglycosides. In general, older biofilms are more developed, thicker, and more viscous.58,102-104 The biofilm may act as a mechanical barrier, resisting penetration or slowing the rate of diffusion of the biocide or antimicrobial; hence, diminishing its effect.105 Only those bacteria found in the outer layer of the biofilm may be exposed to the antimicrobials and antigens. In addition, antibiotic penetration is agent- and organism-specific.

  Altered physiology. The biofilm may dilute, bind, or entrap all or part of the antimicrobial, preventing therapeutic levels of antibiotics, antibodies, or phagocytes from reaching the bacteria within the biofilm,102,106-113 limiting therapy effectiveness. Slow diffusion through the biofilm allows more time for the bacteria within the biofilm to provide a community-wide defense response to weaken the effect of the incoming antimicrobials.114

  When the antimicrobials are able to penetrate the biofilm and kill the bacteria within, a few resistant strains may survive. These surviving bacteria (persisters) may be intracellular and are resistant to further antibiotic treatment.115,116 When antibiotic therapy is discontinued, the persisters reform the biofilm.

  Growth rate. Bacteria in biofilms are associated with slow growth rates,117 especially bacteria in the deeper layers of the biofilm where nutrients may have difficulty penetrating the biofilm and waste product excretion may be slow.1,77,118-120 Many antibiotics (eg, cephalosporins) are less effective against slowly multiplying bacteria.

  Colony size. Biofilm colonies are generally too large for phagocytes to engulf them. The biofilm may resist penetration by the phagocytes, reducing their effectiveness. Chemotactic activities of phagocytes may be inhibited by the biofilm.8,52,113,121,122

  Gene expression. Gene expression plays an important role in pathogenicity and virulence of micro-organisms in the biofilm.123 The phenotypic changes that occur in the micro-organisms within the biofilm make them resistant to antimicrobial treatment and to the host’s immune response.89,124,125 Micro-organisms in biofilms often will express more virulent phenotypes than the same planktonic strain.59,123,126,127 The bacterial cell has a small number of target sites for antibiotics. It is theorized that some cells in the biofilm, using different genes than the planktonic bacteria, phenotypically alter these target sites to protect themselves.124,128-131 Quorum-sensing and sigma factor systems that signal bacteria to change their biochemistry regulate many of these gene expressions. Biocides have a much larger number of target sites, making it difficult for micro-organisms to develop resistance to such compounds; however, bacteria in the biofilm can and do resist biocides.

Biofilm Cultures

  Micro-organisms grow almost everywhere but fewer than 10% can be grown outside their environment in the laboratory.132-137 Accurate identification and determination of micro-organism antibiotic sensitivities are important for the selection of appropriate therapy. However, micro-organisms within biofilm resist conventional culturing methods.8,73,138,139 The same strain of micro-organism outside the biofilm and micro-organisms released when the biofilm is disrupted lose their protection and may be cultured using conventional culturing methods.

  Biofilm colonies are dynamic and frequently changing.30 A biofilm colony containing many bacteria may detach intact, sometimes separated by an injury, and grow as a single colony on or in culture media.74 Because these colonies may be slow growing,117 the cultures may be discarded before the organisms have been identified. When biofilm colonies are cultured, contaminating or colonizing bacteria may dominate the culture and the pathogenic organism, if grown at all, may never be identified.12

  The micro-organisms that separate from the biofilm and are identified by conventional culturing methods may not be representative of the types, numbers, or pathogenicity of bacteria within the biofilm.8,140 Instead of identifying the pathogenic bacteria in the biofilm causing the infection, conventional cultures may identify the non-pathogenic planktonic bacteria; therefore, failure to disrupt the biofilm may result in a falsely sterile culture or the recovery of the non-pathogenic planktonic bacteria colonizing or contaminating the tissues. Sensitivity reports for these bacteria may lead to a false clinical interpretation. Consequently, long-established conventional methods of collecting micro-organisms from bone, blood, joint effusion, swabs, or soft tissue samples in sufficient numbers to be identified in cultures may be thwarted by the properties of the biofilm.12

  Ultrasonic oscillation (sonication) of biofilm culture specimens using low-energy, high frequency sound waves has been shown to disrupt the biofilm.104,141,142 Once the biofilm is disrupted or the micro-organisms leave the biofilm, the micro-organisms revert back to their original phenotype and can be identified using conventional culture techniques. Another method to identify bacteria is a culture-based independent approach to detect and identify micro-organisms in a biofilm-protected environment.135,136,143,144 These non-conventional methods are not readily available to the clinician.145,146

Treatment

  The treatment of biofilm-related infections is complex and beyond the scope of this paper. In addition, definitive treatment for biofilm infections for the most part remains in the sphere of research studies that have yet to reach clinical practice. Much of the data come from animal or laboratory models or industrial use and vary widely in design. Because research studies are performed under ideal conditions that are much different from the conditions encountered in the human body, they are not necessarily clinically useful. Furthermore, the biofilm and the bacteria in the biofilm are dynamic and constantly changing. Because of the complexity of this issue, the treatment of biofilm infections remains poorly understood and under investigation.

Discussion

  Infection is rare considering that on a daily basis the human body coexists in a symbiotic relationship with 1014 micro-organisms from a countless number of vectors.147 The development of an infectious disease and the virulence of the micro-organisms involves a multitude of interrelated micro-organism and host factors that vary among species. Clinicians who understand infectious disease development recognize the importance of biofilms in this complexity.

  An infectious disease does not occur every time a pathogen colonizes the body. Before causing disease, the bacteria must be able to adhere to and colonize the host’s tissues and reproduce successfully while remaining fit and overcoming the host’s defenses. A minimum number of bacteria must be present to express a coordinated sequence of genetic events resulting in infectious disease. DNA comprises instructions that dictate how bacterial pathogens evade antimicrobials and the immune system, change their virulence, and cause infection – the formation of a biofilm is one method micro-organisms use to subvert antimicrobials and the host’s immune system and enable survival in the human body.

  Biofilms attach to wet or moist surfaces – a successful strategy micro-organisms have developed for their survival. Biofilm disease is difficult to eradicate, is a source of many recalcitrant infections, and resembles a multicellular organism or a community structure with many common goals for survival. In addition to those addressed in healthcare, bacterial cells in food, water, and industrial and environmental ecosystems are predominantly organized in specialized biofilms that have significantly different phenotypic properties from free-floating bacteria of the same species. The micro-organisms in the biofilm may remain dormant for years and may periodically shed bacteria; this phenomenon and an injury that dislodges the bacteria may release enough micro-organisms to cause an infection.

  Treatment of an infectious disease has typically depended on the microbiology laboratory to describe planktonic, freely suspended, rapidly growing micro-organisms based on their growth characteristics in culture media and their sensitivities to antibiotics. However, the laboratory environment is not representative of how micro-organisms appear and respond in their natural environment in a host; hence, conventional cultures often do not reveal all the organisms present. In addition, hundreds of bacteria in a biofilm may grow as a single colony.

  Biofilms comprise slow-growing, difficult- or impossible-to-culture micro-organisms. Easily grown, rapidly multiplying, contaminating, and colonizing micro-organisms may overshadow biofilm micro-organisms, making them difficult to recognize or easily overlooked. Thus, the quantity or variety of organisms present may be underestimated. Sonication of the biofilm will disrupt the integrity of the biofilm and release the micro-organisms, providing an opportunity for culture and treatment by conventional methods.

  Treatment of biofilm micro-organisms can be difficult and frustrating. Biofilm infections tend to persist on medical devices, dead bone, and necrotic tissues despite antibiotics, antiseptics, and the host’s immune response. If no biofilm is present, these sites may act as sites of adhesion, the early stage of biofilm development. Biofilm resistance to antimicrobials or the failure to kill or suppress micro-organisms protected in the biofilm may be an underlying dynamic in chronic osteomyelitis, chronic wounds, and the resistance of exposed tendon to coverage with granulation tissue or skin grafting. As a result, biofilms may help explain why some wound care treatments are successful and others are not. Much research is needed in this area.

Conclusion

  Biofilms help explain why infections respond well to removal of a medical device; why chronic quiescent osteomyelitis, cellulitis in the calf, and other recurring infections become active infections; why wounds should be healing but are not; why infected bone and wounds respond well to debridement; and why skin ulcers, wounds, burns, and other necrotic tissues respond well to frequent maintenance debridement

.   Biofilms, persisters, and other unculturable (and therefore, unrecognized) pathogens underscore the inadequacy of sampling and culturing methods presently in use. Experiences with biofilms and persisters suggest the need for a culture-independent approach for pathogen identification. Current culture methods to demonstrate presence of infectious disease and antibiotic treatment based on sensitivities from these cultures may soon become obsolete. Along with other unrecognized pathogens, biofilms provide an opportunity to reconsider commonly held beliefs and assumptions regarding infection and offer new possibilities for diagnosis and treatment. The ramifications of biofilms can be widespread. To this end, clinicians should not only maintain a healthy skepticism regarding seemingly unexplainable phenomena, but also consider all possibilities, no matter how unorthodox.

1. Costerton JW, Stewart PS, Greenberg EP. Bacterial biofilms: a common cause of persistent infections. Science. 1999;284(5418):1318-1322.

2. Singh PK, Schaefer AL, Parsek MR, et al. Quorum-sensing signals indicate that cystic fibrosis lungs are infected with bacterial biofilms. Nature. 2000;407(6805):762-764.

3. Baillie GS, Douglas LJ. Candida biofilms and their susceptibility to antifungal agents. Methods Enzymol. 1999;310:644-656.

4. Kuchma SL, O’Toole GA. Surface-induced and biofilm-induced changes in gene expression. Curr Opin Biotechnol. 2000;11(5):429-433.

5. Adam B, Baillie GS, Douglas LJ. Mixed species biofilms of Candida albicans and Staphylococcus epidermidis. J Med Microbiol. 2002;51(4):344-349.

6. Pratt LA, Kolter R. Genetic analyses of bacterial biofilm formation. Curr Opin Microbiol. 1999;2(6):598-603.

7. Lamfon H, Porter SR, McCullough M, Pratten J. Susceptibility of Candida albicans biofilms grown in a constant depth film fermentor to chlorhexidine, fluconazole and miconazole: a longitudinal study. J Antimicrob Chemother. 2004;53(2):383-385.

8. Gristina AG, Costerton JW. Bacterial adherence and the glycocalyx and their role in musculoskeletal infection. Orthoped Clin N Am. 1984;15(3):517-535.

9. Stickler D. Biofilms. Curr Opin Microbiol. 1999;2(3):270-275.

10. Guide N. NIH Guide: Research on microbial biofilms. Research on microbial biofilms cited. Available from: http://grants.nih.gov/grants/guide/pa-files/PA-98-070.html. Accessed May 14, 1998.

11. Alam SI, Khan KA, Ahmad A. Glycocalyx-positive bacteria isolated from chronic osteomyelitis and septic arthritis. Ceylon Med J. 1990;35(1):21-23.

12. Gristina AG, Costerton JW. Bacterial adherence to biomaterials and tissue. The significance of its role in clinical sepsis. J Bone Joint Surg Am. 1985;67(2 suppl A):264-273.

13. Bos R, Van der Mei HC, Busscher H. Physico-chemistry of initial microbial adhesive interactions – its mechanisms and methods for study. FEMS Microbiol Rev. 1999;23(2):179-230.

14. Merritt K, Chang CC. Factors influencing bacterial adherence to biomaterials. J Biomater Appl. 1991;5(3):185-203.

15. Mittleman M. Biological fouling of purified-water systems: Part 1. Bacterial growth and replication. Microcontamination. 1985;3(10):51-55.

16. Cramton SE, Gerke C, Schnell NF, Nichols WW, Gotz F. The intercellular adhesion (ica) locus is present in Staphylococcus aureus and is required for biofilm formation. Infect Immun. 1999;67(10):5427-5433.

17. Le Thi TT, Prigent-Combaret C, Dorel C, Lejeune P. First stages of biofilm formation: characterization and quantification of bacterial functions involved in colonization process. Methods Enzymol. 2001;336:152-159.

18. Stoodley P, Boyle JD, DeBeer D, Lappin-Scott HM. Evolving perspectives of biofilm structure. Biofouling. 1999;14(1):75-90.

19. Neu TR. Significance of bacterial surface active compounds in interaction with interfaces. Microbiol Rev. 1996;60(1):151-166.

20. Jenkinson H, Lappin-Scott HM. Biofilms adhere to stay. Trends Microbiol. 2001;9(1):9-10.

21. Palmer RJ Jr, White DC. Developmental biology of biofilms: implication for treatment and control. Trends Microbiol. 1997;5(11):435-440.

22. Rickard AH, Gilbert P, High NJ, Kolenbrander PE, Handley PS. Bacterial coaggregation: an integral process in the development of multi-species biofilms. Trends Microbiol. 2003;11(2):94-100.

23. Mims C. The Pathogenesis of Infectious Diseases, 3rd ed. San Diego, Calif: Academic Press;1987.

24. Patti JM, Allen BL, McGavin MJ, Hook M. MSCRAMM-mediated adherence of microorganisms to host tissue. Ann Rev Microbiol. 1994;48:585-617.

25. Kolenbrander P, et al. Biofilm matrix polymers – role in adhesion. In Newman H, Wilson W, eds. Dental Plaque Revisited. London: BioLine Publications;1999.

26. Allison D, Sutherland I, Neu T, eds. EPS: what's in an acronym? In: McBain A, ed. Biofilm Communities: Order from Chaos? Cardiff, UK: Bioline;2003.

27. Costerton JW, Irvin I, Cheng KJ. The bacterial glycocalyx in nature and disease. Ann Rev Microbiol. 1981;35:299-324.

28. Costerton JW, Cheng KJ, Geesey GG, et al. Bacterial biofilms in nature and disease. Ann Rev Microbiol. 1987;41:435-464.

29. Jahn A, Griebe T, Nielson PH. Composition of Pseudomonas putida biofilms: accumulation of protein in the biofilm matrix. Biofouling. 2000;14:49-57.

30. Sutherland I. The biofilm matrix –an immobilized but dynamic microbial environment. Trends Microbiol. 2001;9(5):222-227.

31. Sutherland IW. Biofilm exopolysaccharides: a strong and sticky framework. Trends Microbiol. 2001;9(5):222-227.

32. Flemming HC, Wingender J. Relevance of microbial extracellular polymeric substances (EPSs) – Part I: structural and ecological aspects. Water Sci Technol. 2001;43(6):1-8.

33. Wingender J, Flemming HC, Neu TR. What are bacterial extracellular polymeric substances? In: Wingender J, Neu T, Flemming H, eds. Microbial Extracellular Polymeric Substances – Characterization, Structure and Function. Berlin, Germany: Springer;1999:1-19.

34. Marshall KC. Microbial adhesion in biotechnological processes. Curr Opin Biotechnol. 1994;5(3):296-301.

35. Kumamoto C. Candida biofilms. Curr Opin Microbiol. 2002;5(6):608-611.

36. Korstgens,V, Flemming HC, Wingender J, Borchard W. Influence of calcium ions on the mechanical properties of a model biofilm of mucoid Pseudomonas aeruginosa. Water Sci Technol. 2001;43(6):49-57.

37. Murga R, Miller JM, Donlan RM. Biofilm formation by Gram-negative bacteria on central venous catheter connectors: effect of conditioning films in a laboratory model. J Clin Microbiol. 2001;39(6):2294-2297.

38. Shanks RM, Donegan NP, Graber ML, et al. Heparin stimulates Staphylococcus aureus biofilm formation. Infect Immun. 2005;73(8):4596-4606.

39. Jones SM, Morgan M, Humphrey TJ, Lappin-Scott H. Effect of vancomycin and rifampicin on methicillin-resistant Staphylococcus aureus biofilms. Lancet. 2001;357(9249):40-41.

40. Tunney M, Jones DS, Gorman S. Biofilm and biofilm-related encrustation of urinary tract devices. In: Doyle R, ed. Methods in Enzymology. San Diego, Calif: Academic Press;1999.

41. Serralta V, Harrison-Balestra C, Cazzaniga AL, Davis SC, Mertz PM. Lifestyles of bacteria in wounds: presence of biofilms? WOUNDS. 2001;13(1):29-34.

42. Gristina AG, Naylor PT, Myrvik QN. Musculoskeletal infection, microbial adhesion, and antibiotic resistance. Infect Dis Clin N Am. 1990;4(3):391-408.

43. Webb LX, Hobgood CD, Meredith JW, Gristina AG. Adhesive bacterial colonization of exposed traumatized tendon. Orthop Rev. 1987;16(5):304-309.

44. Akiyama H, Yamasaki O, Kanzaki H, Tada J, Arata J. Effects of sucrose and silver on Staphylococcus aureus biofilms. J Antimicrob Chemother. 1998;42(5):629-634.

45. Post JC, Stoodley P, Hall-Stoodley L, Ehrlich GD. The role of biofilms in otolaryngologic infections. Curr Opin Otolaryngol Head Neck Surg. 2004;12(3):185-190.

46. Lambe DW Jr, Ferguson KP, Mayberry-Carson KJ, Tober-Meyer B, Costerson JW. Foreign-body associated experimental osteomyelitis induced with Bacteroides fragilis and Staphylococcus epidermidis in rabbits. Clin Orthopaed. 1991;266:285-294.

47. Bruinsma G. Detergents in biofilm detachment. Biomed Engineer.

48. Gristina A, Rovere G, Shoji H. Spontaneous septic arthritis complicating rheumatoid arthritis. JBJS. 1974;56A:30-35.

49. Marrie TJ, Lam J, Costerton JW. Bacterial adhesion to uroepithelial cells: a morphologic study. J Infect Dis. 1980;142(2):239-246.

50. Eden CS, Eriksson B, Hanson LA. Adhesion of Escherichia coli to human uroepithelial cells in vitro. J Immun. 1977;18(3):767-774.

51. Katz S, Izhar M, Mirelman D. Bacterial adherence to surgical sutures. A possible factor in suture-induced infection. Ann Surg. 1981;194(1):35-41.

52. Kendall RW, et al. Persistence of bacteria on antibiotic loaded acrylic depots. Clin Orthopaed Related Res. 1996;329:273-280.

53. Gristina AG, Oga M, Webb LX, Hobgood CD. Adherent bacterial colonization in the pathogenesis of osteomyelitis. Science. 1985;228(4702):990-993.

54. van de Belt H, Neut D, Schenk W, et al. Staphylococcus aureus biofilm formation on different gentamicin-loaded polymethylmethacrylate bone cements. Biomaterials. 2001;22(12):1607-1611.

55. van de Belt H, Neut D, Schenk W, et al. Gentamicin release from polymethylmethacrylate bone cements and Staphylococcus aureus biofilm formation. Acta Orthop Scand. 2000;71(6):625-629.

56. Gracia E, Fernandez A, Conchello P, et al. Adherence of Staphylococcus aureus slime-producing strain variants to biomaterials used in orthopaedic surgery. Int Orthop. 1997;21(1):46-51.

57. Vuong C, Saenz HL, Gotz F, Otto M. Impact of the agr quorum-sensing system on adherence to polystyrene in Staphylococcus aureus. J Infect Dis. 2000;182(6):1688-1693.

58. Buxton TB, Horner J, Hinton A, Rissing JP. In vivo glycocalyx expression by Staphylococcus aureus phage type 52/52A/80 in S. aureus osteomyelitis. J Infect Dis. 1987;156(6):942-946.

59. Arciola C, Baldassarri L, Montanaro L. Presence of icaA and icaD genes and slime production in a collection of staphylococcal strains from catheter-associated infections. J Clin Microbiol. 2001;39(6):2151-2156.

60. Stickler DJ, Morris NS, McLean RJ, Fuqua C. Biofilms on indwelling urethral catheters produce quorum sensing signal molecules in situ and in vitro. App Environ Microbiol. 1998;64(9):3486-3490.

61. Whiteley M, Parsek MR, Greenberg EP. Regulation of quorum sensing by RpoS in Pseudomonas aeruginosa. J Bacteriol. 2000;182(15):4356-4360.

62. Strauss EJ, Falkow S. Microbial pathogenesis: genomics and beyond. Science. 1997;276(5313):707-712.

63. Swift S, Throup JP, Williams P, Salmond GP, Steward GS. Quorum sensing: a population-density component in the determination of bacterial phenotype. Trends Biochem Sci. 1996;21(6):214-219.

64. Parsek MR, Greenberg EP. Acyl-homoserine lactone quorum sensing in gram-negative bacteria: a signaling mechanism involved in associations with higher organisms. Proc Natl Acad Sci USA. 2000;97(16):8789-8793.

65. Davies DG, Parsek MR, Pearson JP, et al. The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science. 1998;280(5361):295-298.

66. Bollinger N, Hassett DJ, Iglewski BH, Costerton JW, McDermott TR. Gene expression in Pseudomonas aeruginosa: evidence of iron override effects on quorum sensing and biofilm specific gene regulation. J Bacteriol. 2001;183(6):1990-1996.

67. Hoyle BD, Costerton JW. Bacterial resistance to antibiotics: the role of biofilms. Prog Drug Res. 1991;37:91-105.

68. Rijnaarts HH, Norde W, Bouwer EJ, Lyklema J. Bacterial adhesion under static and dynamic conditions. Appl Environ Microbiol. 1993;59(10):3255-3265.

69. Jackson DW, Suzuki K, Oakford L, et al. Biofilm formation and dispersal under the influence of the global regulator CsrA of Escherichia coli. J Bacteriol. 2002;184(1):290-301.

70. Mayette D. The existence and significance of biofilms in water. Water Review. 1992:1-3.

71. Patterson MK, Husted GR, Rutkowski A, Mayette DC. Isolation, identification and microscopic properties of biofilms in high-purity water distribution systems. Ultrapure Water. 1991;8(4):18-24.

72. Costerton JW, Khoury AE, WArd KH, Anwar H. Practical measures to control device-related bacterial infections. Int J Artif Organs. 1993;16(11):765-770.

73. Khoury A, Lam K, Ellis BD, Costerton JW. Prevention and control of bacterial infections associated with medical devices. Am Soc Artif Intern Organs J. 1991;38(3):M174-M178.

74. Dougherty SH, Simmons RL. Endogenous factors contributing to prosthetic device infections. Infect Dis Clin North Am. 1989;3(2):199-209.

75. Gristina AG. Biomaterial-centered infection: microbial adhesion versus tissue integration. Science. 1987;237(H822):1588-1595.

76. Dunne W Jr, Mason E, Kaplan S. Diffusion of rifampin and vancomycin through a Staphylococcus epidermidis biofilm. Antimicrob Agents Chemother. 1993;37(12):2522-2526. 7

7. Darouiche RO, Dhir A, Miller AJ. Vancomycin penetration into biofilm covering infected prostheses and effect on bacteria. J Infect Dis. 1994;170(3):720-723.

78. Chandy JP, Angles ML. Determination of nutrients limiting biofilm formation and the subsequent impact on disinfectant decay. Water Res. 2001;35(11):2677-2682.

79. Yasuda H, Ajiki Y, Aoyama J, Yokota T. Interaction between human polymorphonuclear leucocytes and bacteria released from in-vitro bacterial biofilm models. J Med Microbiol. 1994;41(5):359-367.

80. Stewart PS. A review of experimental measurements of effective diffusive permeabilities and effective diffusion coefficients in biofilms. Biotechnol Bioeng. 1998;59(3):261-272.

81. Wilcox MH, Kite P, Mills K, Sugden S. In situ measurement of linezolid and vancomycin concentrations in intravascular catheter-associated biofilm. J Antimicrob Chemother. 2001;47(2):171-175.

82. Gander S. Bacterial biofilms: resistance to antimicrobial agents. J Antimicrob Chemother. 1996;37(6):1047-1050.

83. Zheng Z, Stewart PS. Penetration of rifampin through Staphylococcus epidermidis biofilms. Antimicrob Agents Chemother. 2002;46(3):900-903.

84. Nickel JC, Ruseka I, Wright JB, Costerton JW. Tobramycin resistance of cells of Pseudomonas aeruginosa growing as a biofilm on urinary catheter material. Antimicrob Agents Chemother. 1985;27(4):619-624.

85. Sepandj F, Ceri H, Gibb A, Read R, Olson M. Minimum inhibitory concentration (MIC) versus minimum biofilm eliminating concentration (MBEC) in evaluation of antibiotic sensitivity of Gram-negative bacilli causing peritonitis. Perit Dial Int. 2004;24(1):65-67.

86. Dreeszen P. The key to understanding and controlling bacterial growth in automated drinking water systems. Available at: www.edstrom.com/DocLib/biofilm.pdf. Accessed January 31, 2005.

87. Kumon H. Management of biofilm infections in the urinary tract. World J Surg. 2000;24(10):1193-1196.

88. Stewart PS. Multicellular resistance: biofilms. Trends Microbiol. 2001;9(5):204.

89. Schwank S, Rajacic Z, Zimmerli W, Blaser J. Impact of bacterial biofilm formation on in vitro and in vivo activities of antibiotics. Antimicrob Agents Chemother. 1998;42(4):895-898.

90. Mah T, O’Toole GA. Mechanisms of biofilm resistance to antimicrobial agents. Trends Microbiol. 2001;9(1):34-39.

91. Sheldon A Jr. Antibiotic resistance: a survival strategy. Clin Lab Sci. 2005;18(3):170-180.

92. Wimpenny J. An overview of biofilms as functional communities. In: Allison D, Gilbert P, Lappin-Scott HM, Wilson M, eds. SGM Symposium. Cambridge, UK: Cambridge University Press;2000:1-24.

93. Marsh P, Bradshaw D. Microbial community interactions in biofilms, In: Allison D, Gilbert P, Lappin-Scott HM, Wilson M, eds. SGM Symposium. Cambridge, UK: Cambridge University Press;2000:167-198.

94. Gilbert P, Maira-Litran T, McBain AJ, Rickard AH, Whyte FW. The physiology and collective recalcitrance of microbial biofilm communities. Adv Microb Physiol. 2002;46:202-256.

95. Miller MB, Bassler BL. Quorum sensing in bacteria. Ann Rev Microbiol. 2001;55(1):165-199.

96. Riedel K, Hentzer M, Geisenberger O, et al. N-acylhomoserine-lactone-mediated communication between Pseudomonas aeruginosa and Burkholderia cepacia in mixed biofilms. Microbiology. 2001;147(12):3249-3262.

97. Davey ME, O’Toole GA. Microbial biofilms: from ecology to molecular genetics. Microbiol Mol Biol Rev. 2000;64(4):847-867.

98. Moller S, Sternberg C, Andersen JB, et al. Gene expression in mixed culture biofilms: evidence of metabolic interactions between community members. Appl Environ Microbiol. 1998;64(2):721-732.

99. Nielsen AT, Tolker-Nielsen T, Barken KB, Molin S. Role of commensal relationships on the spatial structure of a surface-attached microbial consortium. Environ Microbiol. 2000;2(1):59-68.

100. Cowan S, Gilbert E, Liepmann D, Keasling JD. Commensal interactions in a dual species biofilm exposed to mixed organic compounds. Appl Environ Microbiol. 2000;66(10):4481-4485.

101. Lawrence JR, Korber DR, Wolfaardt GM, Caldwell DE. Behavioral strategies of surface-colonizing bacteria. Adv Microb Ecol. 1995;14:1-75.

102. Anwar H, Strap JL, Costerton JW. Establishment of aging biofilms: possible mechanism of bacterial resistance to antimicrobial therapy. Antimicrob Agents Chemother. 1992;36(7):1347-1351.

103. Brown ML, Aldrich HC, Gauthier JJ. Relationship between glycocalyx and povidone-iodine resistance in Pseudomonas aeruginosa (ATCC 27853) biofilms. Appl Environ Microbiol. 1995;61(1):187-193.

104. Amorena B, Gracia E, Monzon M, et al. Antibiotic susceptibility assay for Staphylococcus aureus in biofilms developed in vitro. J Antimicrob Chemother. 1999;44(1):43-55.

105. Evans K, Passador L, Srikumar R, et al. Influence of the MexAB-OprM Multidrug Efflux System on quorum sensing in Pseudomonas aeruginosa. J Bacteriol. 1998;180(20):5443-5447.

106. Anderl JN, Franklin MJ, Stewart PS. Role of antibiotic penetration limitation in Klebsiella pneumoniae biofilm resistance to ampicillin and ciprofloxacin. Antimicrob Agents Chemother. 2000;44(7):1818-1824.

107. Stewart P. Theoretical aspects of antibiotic diffusion into microbial biofilms. Antimicrob Agents Chemother. 1996;40(11):2517-2522.

108. Shigeta M, Tanaka G, Komatsuzawa H, et al. Permeation of antimicrobial agents through Pseudomonas aeruginosa biofilms: a simple method. Chemotherapy. 1997;43(5):340-345.

109. Vrany JS, Stewart PS, Suci P. Comparison of recalcitrance to ciprofloxacin and levofloxacin exhibited by Pseudomonas aeruginosa biofilms displaying rapid-transport characteristics. Antimicrob Agents Chemother. 1997;41(6):1352.

110. Brooun A, Liu S, Lewis K. A dose-response study of antibiotic resistance in Pseudomonas aeruginosa biofilms. Antimicrob Agents Chemother. 2000;44(3):640-646.

111. Cochran WL, McFeters GA, Stewart PS. Reduced susceptibility of thin Pseudomonas aeruginosa biofilms to hydrogen peroxide and monochloramine. J Appl Microbiol. 2000;88:22-30.

112. Ishida H, Ishida Y, Kurosaka Y, et al. In vitro and in vivo activities of levofloxacin against biofilm-producing Pseudomonas aeruginosa. Antimicrob Agents Chemother. 1998;42(7):1641-1645.

113. Veringa EM, Ferguson DA Jr, Lambe DW Jr, Verhoef J. The role of glycocalyx in surface phagocytosis of Bacteroides spp. in the presence and absence of clindamycin. J Antimicrob Chemother. 1989;23(5):711-720.

114. Szomolay B, Klapper I, Dockery J, Stewart PS. Adaptive responses to antimicrobial agents in biofilms. Environ Microbiol. 2005;7(8):1186-1191.

115. Lewis K. Programmed death in bacteria. Microbiol Mol Biol. 2000;64(3):503-514.

116. Lewis K. Riddle of biofilm resistance. Antimicrob Agents Chemother. 2001;45(4): 999-1007.

117. Sternberg C, Christensen BB, Johansen T, et al. Distribution of bacterial growth activity in flow-chamber biofilms. Appl Environ Microbiol. 1999;65(9):4108-4117.

118. Evans DJ, Allison DG, Brown MR, Gilbert P. Susceptibility of Escherichia coli and Pseudomonas aeruginosa biofilms toward ciprofloxacin: effect of specific growth rate. J Antimicrob Chemother. 1991;27:177-184.

119. Duguid IG, Evans E, Brown MR, Gilbert P. Growth-rate independent killing by ciprofloxacin of biofilm derived Staphylococcus epidermidis: evidence for cell-cycle dependency. J Antimicrob Chemother. 1992;30(6):791-802.

120. Evans DJ, Brown MR, Allison DG, Gilbert P. Susceptibility of bacterial biofilms to tobramycin: role of specific growth rate and phase in the division cycle. J Antimicrob Chemother. 1990;25(4):585-591.

121. Zimmerli W, Lew PD, Waldvogel FA. Pathogenesis of foreign body infection. Evidence of a local granulocyte defect. J Clin Invest. 1984;73(4):1191-1200.

122. Vaudaux P, Zulian G, Huggler E, Waldvogel FA. Attachment of Staphylococcus aureus to polymethylmethacrylate increases it resistance to phagocytosis in foreign body infection. Infect Immun. 1985;50(2):472-477.

123. Gelosia A, Baldassarri L, Deighton M, van Nguyen T. Phenotypic and genotypic markers of Staphylococcus epidermidis virulence. Clin Microbiol Infect. 2001;7(4):193-199.

124. Becker P, Hufnagle W, Peters G, Herrmann M. Detection of differential gene expression in biofilm-forming versus planktonic populations of Staphylococcus aureus using micro-representational – difference analysis. Appl Environ Microbiol. 2001;67(7):2958-2965.

125. Foley I, Gilbert P. In-vitro studies of the activity of glycopeptide combinations against Enterococcus faecalis biofilms. J Antimicrob Chemother. 1997;40(5):667-672.

126. de Kievit T, Seed PC, Nezezon J, Passador L, Iglewski BH. RsaL, a novel repressor of virulence gene expression in Pseudomonas aeruginosa. J Bacteriol. 1999;181(7):2175-2184.

127. Braga PC, Sasso MD, Sala MT. Sub-MIC concentrations of cefodizime interfere with various factors affecting bacterial virulence. Antimicrob Chemotherap. 2000;45(1):15-25.

128. Gilbert P, Brown MRW. Phenotypic plasticity and mechanisms of protection of bacterial biofilms from antimicrobial agents. In: Lappin-Scott H, Costerton JW, eds. Microbial Biofilms. Cambridge, UK: Cambridge University Press;1995:118-132.

129. Hamilton-Miller JM, Shah S. Activity of quinupristin/dalfopristin against Staphylococcus epidermidis in biofilms: a comparison with ciprofloxacin. J Antimicrob Chemother. 1997;39(suppl A):S103-S108.

130. Konig C, Schwank S, Blaser J. Factors compromising antibiotic activity against biofilms of Staphylococcus epidermidis. Eur J Clin Microbiol Infec Dis. 1991;20(1):20-26.

131. Coughlan A. Slime city. New Scientist. 1996;15(2045):32-36.

132. Kaeberlein T, Lewis K, Epstein SS. Isolating “uncultivable” microorganisms in pure culture in a simulated natural environment. Science. 2002;296(5570):1127-1129.

133. Pace N. A molecular view of microbial diversity and the biosphere. Science. 1997;276(5313):734-740.

134. Amann R, Ludwig W, Schleifer KH. Phylogenetic identification and in situ detection of individual microbial cells without cultivation. Microbiol Rev. 1995;59(1):143-169.

135. Relman DA. The identification of uncultured microbial pathogens. J Infect Dis. 1993;168(1):1-8.

136. Relman DA. The search for unrecognized pathogens. Science. 1999;284(5418):1308-1310.

137. Whitman WB, Coleman DC, Wiebe WJ. Prokaryotes: the unseen majority. Proc Natl Acad Sci USA. 1998;95(12):6578-6583.

138. Boyd A, Chakrabarty AM. Pseudomonas aeruginosa biofilms: role of the alginate exopolysaccharide. J Ind Microbiol. 1995;15(3):162-168.

139. Hussain M, Wilcox MH, White PJ. The slime of coagulase-negative staphylococci: biochemistry and relation to adherence. FEMS Microbiol Rev. 1993;10(3-4):191-207.

140. Brown MR, Barker J. Unexplored reservoirs of pathogenic bacteria: protozoa and biofilms. Trends Microbiol. 1999;7(1):46-50.

141. Tollefson DF, Bandyk DF, Kaebnick HW, Seabrook GR, Towne JB. Surface biofilm disruption. Enhanced recovery of microorganisms from vascular prostheses. Arch Surg. 1987;122(1):38-43.

142. Pajkos A, Deva AK, Vickery K, et al. Detection of subclinical infection in significant breast implant capsules. Plast Reconstr Surg. 2003;111(5):1605-1611.

143. Relman DA, Falkow S. Identification of uncultured microorganism: expanding the spectrum of characterized microbial pathogens. Infect Agents Dis. 1992;1(5):245-253.

144. Loesche WJ, Lopatin DE, Stoll J, van Poperin N, Hujoel PP. Comparison of various detection methods for periodontopathic bacteria: can culture be considered the primary reference standard. J Clin Microbiol. 1992;30(2):418-426.

145. Amann R, Kuhl M. In situ methods for assessment of microorganisms and their activities. Curr Opin Microbiol. 1998;1(3):352-358.

146. Arciola CR, Collamati S, Donati E, Montanaro L. A rapid PCR method for the detection of slime-producing strains of Staphylococcus epidermidis and S. aureus in periprosthesis infections. Diagn Mol Pathol. 2001;10(2):130-137.

147.Mietzner BA, Mietzner TA, eds. Microbial Pathogenesis: A Principles-Oriented Approach. Madison, Conn: Fence Creek Publishing;1999.

Advertisement

Advertisement

Advertisement