Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-c47g7 Total loading time: 0 Render date: 2024-04-24T15:57:53.498Z Has data issue: false hasContentIssue false

8 - Phase variation and antigenic variation

Published online by Cambridge University Press:  06 August 2009

Richard Villemur
Affiliation:
INRS-Institut
Eric Déziel
Affiliation:
INRS-Institut
Peter Mullany
Affiliation:
University College London
Get access

Summary

Bacteria owe their ability to thrive in diverse and ever-changing conditions to their extraordinary faculty of adaptation. They have developed many strategies to adjust to new environments. These mechanisms include random modifications within their genome, such as point mutation, duplication, deletion, insertion, and acquisition of new DNA (e.g., lateral gene transfer). These multiple events generate a heterogeneous microbial population containing numerous novel phenotypes. Whenever the environment changes, a subpopulation more apt to survive in these new conditions emerges, thus allowing bacteria to thrive, for example, by acquiring resistance to antibiotics. However, as the intensity, duration, and nature of stress are extremely variable, the optimal response to new environmental conditions may be unpredictable. The means by which bacteria either respond to stress, such as exposure to toxic/inhibitory compounds or starvation, or avoid detection by the host's immune system are crucial for their survival. The spontaneous mutation rate is usually insufficient for allowing an efficient adaptation to these changes. However, certain bacterial populations contain hypermutator strains exhibiting highly increased rates of spontaneous mutations, therefore promoting adaptation to changing environments (Taddei et al., 1997). Nevertheless, this benefit may disappear once adaptation is achieved because the evolved genotype may have accumulated irreversible mutations that are detrimental in other conditions (Giraud et al., 2001a, 2001b).

Alternatively, bacteria have developed adaptation strategies based on DNA rearrangement events restricted to specific genomic regions. These defined loci allow bacteria to generate an array of phenotypic variants, whereas minimizing detrimental effects of random mutations on fitness.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2005

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Anriany, Y. A., Weiner, R. M., Johnson, J. A., Rezende, C. E., and Joseph, S. W. (2001). Salmonella enterica serovar Typhimurium DT104 displays a rugose phenotype. Appl. Environ. Microbiol. 67, 4048–4056CrossRefGoogle ScholarPubMed
Arciola, C. R., Baldassarri, L., and Montanaro, L. (2001). Presence of icaA and icaD genes and slime production in a collection of staphylococcal strains from catheter-associated infections. J. Clin. Microbiol. 39, 2151–2156CrossRefGoogle Scholar
Aslund, F., and Beckwith, J. (1999). Bridge over troubled waters: Sensing stress by disulfide bond formation. Cell 96, 751–753CrossRefGoogle ScholarPubMed
Bahl, H., Scholz, H., Bayan, N., Chami, M., Leblon, G., Gulik-Krzywicki, T., Shechter, E.. (1997). Molecular biology of S-layers. FEMS Microbiol. Rev. 20, 47–98CrossRefGoogle ScholarPubMed
Ball, C. A., Osuna, R., Ferguson, K. C., and Johnson, R. C. (1992). Dramatic changes in Fis levels upon nutrient upshift in Escherichia coli. J. Bacteriol. 174, 8043–8056CrossRefGoogle ScholarPubMed
Banerjee, A., Wang, R., Supernavage, S. L., Ghosh, S. K., Parker, J., Ganesh, N. F., Wang, P. G.. (2002). Implications of phase variation of a gene (pgtA) encoding a pilin galactosyl transferase in gonococcal pathogenesis. J. Exp. Med. 196, 147–162CrossRefGoogle ScholarPubMed
Banerjee, A., Wang, R., Uljon, S. N., Rice, P. A., Gotschlich, E. C., and Stein, D. C. (1998). Identification of the gene (lgtG) encoding the lipooligosaccharide beta chain synthesizing glucosyl transferase from Neisseria gonorrhoeae. Proc. Natl. Acad. Sci. U S A 95, 10872–10877CrossRefGoogle ScholarPubMed
Barbour, A. G. (1990). Antigenic variation of a relapsing fever Borrelia species. Annu. Rev. Microbiol. 44, 155–171CrossRefGoogle ScholarPubMed
Barbour, A. G. (1993). Linear DNA of Borrelia species and antigenic variation. Trends Microbiol. 1, 236–239CrossRefGoogle ScholarPubMed
Barbour, A. G., Burman, N., Carter, C. J., Kitten, T., and Bergstrom, S. (1991a). Variable antigen genes of the relapsing fever agent Borrelia hermsii are activated by promoter addition. Mol. Microbiol. 5, 489–493CrossRefGoogle Scholar
Barbour, A. G., Carter, C. J., Burman, N., Freitag, C. S., Garon, C. F., and Bergstrom, S. (1991b). Tandem insertion sequence-like elements define the expression site for variable antigen genes of Borrelia hermsii. Infect. Immun. 59, 390–397Google Scholar
Barbour, A. G., Carter, C. J., and Sohaskey, C. D. (2000). Surface protein variation by expression site switching in the relapsing fever agent Borrelia hermsii. Infect. Immun. 68, 7114–7121CrossRefGoogle ScholarPubMed
Barbour, A. G., and Restrepo, B. I. (2000). Antigenic variation in vector-borne pathogens. Emerg. Infect. Dis. 6, 449–457CrossRefGoogle ScholarPubMed
Bartlett, D. H., and Silverman, M. (1989). Nucleotide sequence of IS492, a novel insertion sequence causing variation in extracellular polysaccharide production in the marine bacterium Pseudomonas atlantica. J. Bacteriol. 171, 1763–1766CrossRefGoogle ScholarPubMed
Bartlett, D. H., Wright, M. E., and Silverman, M. (1988). Variable expression of extracellular polysaccharide in the marine bacterium Pseudomonas atlantica is controlled by genome rearrangement. Proc. Natl. Acad. Sci. U S A 85, 3923–3927CrossRefGoogle ScholarPubMed
Bayliss, C. D., Field, D., and Moxon, E. R. (2001). The simple sequence contingency loci of Haemophilus influenzae and Neisseria meningitidis. J. Clin. Invest. 107, 657–662CrossRefGoogle ScholarPubMed
Bayliss, C. D., Ven, T., and Moxon, E. R. (2002). Mutations in polI but not mutSLH destabilize Haemophilus influenzae tetranucleotide repeats. EMBO J. 21, 1465–1476CrossRefGoogle Scholar
Belland, R. J. (1991). H-DNA formation by the coding repeat elements of neisserial opa genes. Mol. Microbiol. 5, 2351–2360CrossRefGoogle ScholarPubMed
Bhugra, B., Voelker, L. L., Zou, N., Yu, H., and Dybvig, K. (1995). Mechanism of antigenic variation in Mycoplasma pulmonis: Interwoven, site-specific DNA inversions. Mol. Microbiol. 18, 703–714CrossRefGoogle ScholarPubMed
Blomfield, I. C. (2001). The regulation of pap and type 1 fimbriation in Escherichia coli. Adv. Microb. Physiol. 45, 1–49CrossRefGoogle ScholarPubMed
Boemare, N. E., and Akhurst, R. J. (1988). Biochemical and physiological characterization of colony form variants in Xenorhabdus spp. (Enterobacteriaceae). J. Gen. Microbiol. 134, 751–761Google Scholar
Boguslavsky, S., Menaker, D., Lysnyansky, I., Liu, T., Levisohn, S., Rosengarten, R., Garcia, M.. (2000). Molecular characterization of the Mycoplasma gallisepticum pvpA gene which encodes a putative variable cytadhesin protein. Infect. Immun. 68, 3956–3964CrossRefGoogle ScholarPubMed
Bonifield, H. R., and Hughes, K. T. (2003). Flagellar phase variation in Salmonella enterica is mediated by a posttranscriptional control mechanism. J. Bacteriol. 185, 3567–3574CrossRefGoogle ScholarPubMed
Boot, H. J., and Pouwels, P. H. (1996). Expression, secretion and antigenic variation of bacterial S-layer proteins. Mol. Microbiol. 21, 1117–1123CrossRefGoogle ScholarPubMed
Carter, C. J., Bergstrom, S., Norris, S. J., and Barbour, A. G. (1994). A family of surface-exposed proteins of 20 kilodaltons in the genus Borrelia. Infect. Immun. 62, 2792–2799Google ScholarPubMed
Cho, S. H., Naber, K., Hacker, J., and Ziebuhr, W. (2002). Detection of the icaADBC gene cluster and biofilm formation in Staphylococcus epidermidis isolates from catheter-related urinary tract infections. Int. J. Antimicrob. Agents 19, 570–575CrossRefGoogle ScholarPubMed
Conlon, K. M., Humphreys, H., and O'Gara, J. P. (2004). Inactivations of rsbU and sarA by IS256 represent novel mechanisms of biofilm phenotypic variation in Staphylococcus epidermidis. J. Bacteriol. 186, 6208–6219CrossRefGoogle ScholarPubMed
Correnti, J., Munster, V., Chan, T., and van der Woude, M. (2002). Dam-dependent phase variation of Ag43 in Escherichia coli is altered in a seqA mutant. Mol. Microbiol. 44, 521–532CrossRefGoogle Scholar
Costerton, J. W., Lewandowski, Z., Caldwell, D. E., Korber, D. R., and Lappin-Scott, H. M. (1995). Microbial biofilms. Annu. Rev. Microbiol. 49, 711–745CrossRefGoogle ScholarPubMed
Danaher, R. J., Levin, J. C., Arking, D., Burch, C. L., Sandlin, R., and Stein, D. C. (1995). Genetic basis of Neisseria gonorrhoeae lipooligosaccharide antigenic variation. J. Bacteriol. 177, 7275–7279CrossRefGoogle ScholarPubMed
Danese, P. N., Pratt, L. A., Dove, S. L., and Kolter, R. (2000). The outer membrane protein, antigen 43, mediates cell-to-cell interactions within Escherichia coli biofilms. Mol. Microbiol. 37, 424–432CrossRefGoogle ScholarPubMed
Bolle, X., Bayliss, C. D., Field, D., Ven, T., Saunders, N. J., Hood, D. W., and Moxon, E. R. (2000). The length of a tetranucleotide repeat tract in Haemophilus influenzae determines the phase variation rate of a gene with homology to type III DNA methyltransferases. Mol. Microbiol. 35, 211–222CrossRefGoogle ScholarPubMed
Deitsch, K. W., Moxon, E. R., and Wellems, T. E. (1997). Shared themes of antigenic variation and virulence in bacterial, protozoal, and fungal infections. Microbiol. Mol. Biol. Rev. 61, 281–293Google ScholarPubMed
Déziel, E., Comeau, Y., and Villemur, R. (2001). Initiation of biofilm formation by Pseudomonas aeruginosa 57RP correlates with emergence of hyperpiliated and highly adherent phenotypic variants deficient in swimming, swarming, and twitching motilities. J. Bacteriol. 183, 1195–1204CrossRefGoogle ScholarPubMed
Donnenberg, M. S., and Welch, R. A. (1996). Virulence determinants of uropathogenic Escherichia coli. In Mobley, H. L. and Warren, J. W. (Eds.), Urinary tract infections: Molecular pathogenesis and clinical management, pp. 135–174. Washington, DC: ASM PressGoogle Scholar
Dorman, C. J., and Higgins, C. F. (1987). Fimbrial phase variation in Escherichia coli: Dependence on integration host factor and homologies with other site-specific recombinases. J. Bacteriol. 169, 3840–3843CrossRefGoogle ScholarPubMed
Dove, S. L., and Dorman, C. J. (1994). The site-specific recombination system regulating expression of the type 1 fimbrial subunit gene of Escherichia coli is sensitive to changes in DNA supercoiling. Mol. Microbiol. 14, 975–988CrossRefGoogle ScholarPubMed
Dove, S. L., and Dorman, C. J. (1996). Multicopy fimB gene expression in Escherichia coli: Binding to inverted repeats in vivo, effect on fimA gene transcription and DNA inversion. Mol. Microbiol. 21, 1161–1173CrossRefGoogle ScholarPubMed
Dove, S. L., Smith, S. G., and Dorman, C. J. (1997). Control of Escherichia coli type 1 fimbrial gene expression in stationary phase: A negative role for RpoS. Mol. Gen. Genet. 254, 13–20CrossRefGoogle ScholarPubMed
Dworkin, J., and Blaser, M. J. (1996). Generation of Campylobacter fetus S-layer protein diversity utilizes a single promoter on an invertible DNA segment. Mol. Microbiol. 19, 1241–1253CrossRefGoogle ScholarPubMed
Dworkin, J., and Blaser, M. J. (1997a). Molecular mechanisms of Campylobacter fetus surface layer protein expression. Mol. Microbiol. 26, 433–440CrossRefGoogle Scholar
Dworkin, J., and Blaser, M. J. (1997b). Nested DNA inversion as a paradigm of programmed gene rearrangement. Proc. Natl. Acad. Sci. U S A 94, 985–990CrossRefGoogle Scholar
Dworkin, J., Shedd, O. L., and Blaser, M. J. (1997). Nested DNA inversion of Campylobacter fetus S-layer genes is recA dependent. J. Bacteriol. 179, 7523–7529CrossRefGoogle Scholar
Dworkin, J., Tummuru, M. K., and Blaser, M. J. (1995a). A lipopolysaccharide-binding domain of the Campylobacter fetus S-layer protein resides within the conserved N terminus of a family of silent and divergent homologs. J. Bacteriol. 177, 1734–1741CrossRefGoogle Scholar
Dworkin, J., Tummuru, M. K., and Blaser, M. J. (1995b). Segmental conservation of sapA sequences in type B Campylobacter fetus cells. J. Biol. Chem. 270, 15093–15101CrossRefGoogle Scholar
Dybvig, K. (1993). DNA rearrangements and phenotypic switching in prokaryotes. Mol. Microbiol. 10, 465–471CrossRefGoogle ScholarPubMed
Eisenstein, B. I., Sweet, D. S., Vaughn, V., and Friedman, D. I. (1987). Integration host factor is required for the DNA inversion that controls phase variation in Escherichia coli. Proc. Natl. Acad. Sci. U S A 84, 6506–6510CrossRefGoogle ScholarPubMed
Faguy, D. M. (2000). The controlled chaos of shifty pathogens. Curr. Biol. 10, R498–501CrossRefGoogle ScholarPubMed
Fenlon, D. R. (1999). Listeria monocytogenes in the natural environment. In Ryser, E. T. and Marth, E. H. (Eds.), Listeria, listeriosis, and food safety, vol. 1, pp. 21–38. New York: Marcel DekkerGoogle Scholar
Flitman-Tene, R., Levisohn, S., Lysnyansky, I., Rapoport, E., and Yogev, D. (2000). A chromosomal region of Mycoplasma agalactiae containing vsp-related genes undergoes in vivo rearrangement in naturally infected animals. FEMS Microbiol. Lett. 191, 205–212CrossRefGoogle ScholarPubMed
Forst, S., Dowds, B., Boemare, N., and Stackebrandt, E. (1997). Xenorhabdus and Photorhabdus spp.: Bugs that kill bugs. Annu. Rev. Microbiol. 51, 47–72CrossRefGoogle ScholarPubMed
Freitag, N. E., Seifert, H. S., and Koomey, M. (1995). Characterization of the pilF-pilD pilus-assembly locus of Neisseria gonorrhoeae. Mol. Microbiol. 16, 575–586CrossRefGoogle ScholarPubMed
Gally, D. L., Bogan, J. A., Eisenstein, B. I., and Blomfield, I. C. (1993). Environmental regulation of the fim switch controlling type 1 fimbrial phase variation in Escherichia coli K-12: Effects of temperature and media. J. Bacteriol. 175, 6186–6193CrossRefGoogle ScholarPubMed
Gally, D. L., Leathart, J., and Blomfield, I. C. (1996). Interaction of FimB and FimE with the fim switch that controls the phase variation of type 1 fimbriae in Escherichia coli K-12. Mol. Microbiol. 21, 725–738CrossRefGoogle ScholarPubMed
Gally, D. L., Rucker, T. J., and Blomfield, I. C. (1994). The leucine-responsive regulatory protein binds to the fim switch to control phase variation of type 1 fimbrial expression in Escherichia coli K-12. J. Bacteriol. 176, 5665–5672CrossRefGoogle ScholarPubMed
Gilbert, P., Maira-Litran, T., McBain, A. J., Rickard, A. H., and Whyte, F. W. (2002). The physiology and collective recalcitrance of microbial biofilm communities. Adv. Microb. Physiol. 46, 202–256Google ScholarPubMed
Giraud, A., Matic, I., Tenaillon, O., Clara, A., Radman, M., Fons, M., and Taddei, F. (2001a). Costs and benefits of high mutation rates: Adaptive evolution of bacteria in the mouse gut. Science 291, 2606–2608CrossRefGoogle Scholar
Giraud, A., Radman, M., Matic, I., and Taddei, F. (2001b). The rise and fall of mutator bacteria. Curr. Opin. Microbiol. 4, 582–585CrossRefGoogle Scholar
Glew, M. D., Marenda, M., Rosengarten, R., and Citti, C. (2002). Surface diversity in Mycoplasma agalactiae is driven by site-specific DNA inversions within the vpma multigene locus. J. Bacteriol. 184, 5987–5998CrossRefGoogle ScholarPubMed
Gotschlich, E. C. (1994). Genetic locus for the biosynthesis of the variable portion of Neisseria gonorrhoeae lipooligosaccharide. J. Exp. Med. 180, 2181–2190CrossRefGoogle ScholarPubMed
Gotz, F. (2002). Staphylococcus and biofilms. Mol. Microbiol. 43, 1367–1378CrossRefGoogle ScholarPubMed
Gunther, N. W., Snyder, J. A., Lockatell, V., Blomfield, I., Johnson, D. E., and Mobley, H. L. (2002). Assessment of virulence of uropathogenic Escherichia coli type 1 fimbrial mutants in which the invertible element is phase-locked on or off. Infect. Immun. 70, 3344–3354CrossRefGoogle ScholarPubMed
Gyohda, A., Funayama, N., and Komano, T. (1997). Analysis of DNA inversions in the shufflon of plasmid R64. J. Bacteriol. 179, 1867–1871CrossRefGoogle ScholarPubMed
Gyohda, A., Furuya, N., Kogure, N., and Komano, T. (2002). Sequence-specific and non-specific binding of the Rci protein to the asymmetric recombination sites of the R64 shufflon. J. Mol. Biol. 318, 975–983CrossRefGoogle ScholarPubMed
Gyohda, A., Furuya, N., Ishiwa, A., Zhu, S., and Komano, T. (2004). Structure and function of the shufflon in plasmid R64. Adv. Biophysics 38, 183–213CrossRefGoogle ScholarPubMed
Gyohda, A., and Komano, T. (2000). Purification and characterization of the R64 shufflon-specific recombinase. J. Bacteriol. 182, 2787–2792CrossRefGoogle ScholarPubMed
Haagmans, W., and Woude, M. (2000). Phase variation of Ag43 in Escherichia coli: Dam-dependent methylation abrogates OxyR binding and OxyR-mediated repression of transcription. Mol. Microbiol. 35, 877–887CrossRefGoogle ScholarPubMed
Haake, D. A. (2000). Spirochaetal lipoproteins and pathogenesis. Microbiology 146, 1491–1504CrossRefGoogle ScholarPubMed
Haas, R., and Meyer, T. F. (1986). The repertoire of silent pilus genes in Neisseria gonorrhoeae: Evidence for gene conversion. Cell 44, 107–115CrossRefGoogle ScholarPubMed
Haas, R., and Meyer, T. F. (1987). Molecular principles of antigenic variation in Neisseria gonorrhoeae. Antonie Van Leeuwenhoek 53, 431–434CrossRefGoogle ScholarPubMed
Haas, R., Schwarz, H., and Meyer, T. F. (1987). Release of soluble pilin antigen coupled with gene conversion in Neisseria gonorrhoeae. Proc. Natl. Acad. Sci. U S A 84, 9079–9083CrossRefGoogle ScholarPubMed
Hal Jones, C., Dodson, K., and Hultgren, S. J. (1996). Structure, function, and assembly of adhesive P pili. In Mobley, H. L. and Warren, J. W. (Eds.), Urinary tract infections: Molecular pathogenesis and clinical management, pp. 175–219. Washington, DC: ASM PressGoogle Scholar
Hammerschmidt, S., Hilse, R., Putten, J. P., Gerardy-Schahn, R., Unkmeir, A., and Frosch, M. (1996). Modulation of cell surface sialic acid expression in Neisseria meningitidis via a transposable genetic element. EMBO J. 15, 192–198Google Scholar
Handke, L. D., Conlon, K. M., Slater, S. R., Elbaruni, S., Fitzpatrick, F., Humphreys, H., Giles, W. P., Rupp, M. E., Fey, P. D., and O'Gara, J. P. (2004). Genetic and phenotypic analysis of biofilm phenotypic variation in multiple Staphylococcus epidermidis isolates. J. Med. Microbiol. 53, 367–374CrossRefGoogle ScholarPubMed
Hasman, H., Chakraborty, T., and Klemm, P. (1999). Antigen-43-mediated autoaggregation of Escherichia coli is blocked by fimbriation. J. Bacteriol. 181, 4834–4841Google ScholarPubMed
Hasman, H., Schembri, M. A., and Klemm, P. (2000). Antigen 43 and type 1 fimbriae determine colony morphology of Escherichia coli K-12. J. Bacteriol. 182, 1089–1095CrossRefGoogle ScholarPubMed
Heichman, K. A., and Johnson, R. C. (1990). The Hin invertasome: Protein-mediated joining of distant recombination sites at the enhancer. Science 249, 511–517CrossRefGoogle ScholarPubMed
Heilmann, C., Schweitzer, O., Gerke, C., Vanittanakom, N., Mack, D., and Gotz, F. (1996). Molecular basis of intercellular adhesion in the biofilm-forming Staphylococcus epidermidis. Mol. Microbiol. 20, 1083–1091CrossRefGoogle ScholarPubMed
Henderson, I. R., Meehan, M., and Owen, P. (1997). Antigen 43, a phase-variable bipartite outer membrane protein, determines colony morphology and autoaggregation in Escherichia coli K- 12. FEMS Microbiol. Lett. 149, 115–120CrossRefGoogle ScholarPubMed
Henderson, I. R., and Owen, P. (1999). The major phase-variable outer membrane protein of Escherichia coli structurally resembles the immunoglobulin A1 protease class of exported protein and is regulated by a novel mechanism involving Dam and OxyR. J. Bacteriol. 181, 2132–2141Google ScholarPubMed
Henderson, I. R., Owen, P., and Nataro, J. P. (1999). Molecular switches – The ON and OFF of bacterial phase variation. Mol. Microbiol. 33, 919–932CrossRefGoogle ScholarPubMed
Hengge-Aronis, R. (2002). Signal transduction and regulatory mechanisms involved in control of the sigma(S) (RpoS) subunit of RNA polymerase. Microbiol. Mol. Biol. Rev. 66, 373–395CrossRefGoogle ScholarPubMed
Hernday, A. D., Braaten, B. A., and Low, D. A. (2003). The mechanism by which DNA adenine methylase and PapI activate the pap epigenetic switch. Mol. Cell 12, 947–957CrossRefGoogle ScholarPubMed
Hernday, A., Braaten, B., and Low, D. (2004). The intricate workings of a bacterial epigenetic switch. Adv. Exp. Med. Biol. 547, 83–89CrossRefGoogle ScholarPubMed
Hernday, A., Krabbe, M., Braaten, B., and Low, D. A. (2002). Self-perpetuating epigenetic pili switches in bacteria. Proc. Nat. Acad. Sc. USA. 99, 16470–16476 Suppl. 4CrossRefGoogle ScholarPubMed
Hill, S. A., and Grant, C. C. (2002). Recombinational error and deletion formation in Neisseria gonorrhoeae: A role for RecJ in the production of pilE (L) deletions. Mol. Genet. Genomics 266, 962–972Google ScholarPubMed
Howell-Adams, B., and Seifert, H. S. (1999). Insertion mutations in pilE differentially alter gonococcal pilin antigenic variation. J. Bacteriol. 181, 6133–6141Google ScholarPubMed
Howell-Adams, B., and Seifert, H. S. (2000). Molecular models accounting for the gene conversion reactions mediating gonococcal pilin antigenic variation. Mol. Microbiol. 37, 1146–1158CrossRefGoogle ScholarPubMed
Howell-Adams, B., Wainwright, L. A., and Seifert, H. S. (1996). The size and position of heterologous insertions in a silent locus differentially affect pilin recombination in Neisseria gonorrhoeae. Mol. Microbiol. 22, 509–522CrossRefGoogle Scholar
Hultgren, S. J., Hal Jones, C., and Normark, S. (1996). Bacterial adhesins and their assembly. In Curtiss, R. III, Ingraham, J. L., Lin, E. C. C., Low, K. B., Magasanik, B., Reznikoff, W. S., Riley, et al M.. (Eds.), Escherichia coli and Salmonella. Cellular and molecular biology, 2nd ed., pp. 2730–2756. Washington, DC: ASM PressGoogle Scholar
Ishiwa, A., and Komano, T. (2000). The lipopolysaccharide of recipient cells is a specific receptor for PilV proteins, selected by shufflon DNA rearrangement, in liquid matings with donors bearing the R64 plasmid. Mol. Gen. Genet. 263, 159–164CrossRefGoogle ScholarPubMed
Jennings, M. P., Virji, M., Evans, D., Foster, V., Srikhanta, Y. N., Steeghs, L., Ley, P.. (1998). Identification of a novel gene involved in pilin glycosylation in Neisseria meningitidis. Mol. Microbiol. 29, 975–984CrossRefGoogle ScholarPubMed
Johnson, R. C., Glasgow, A. C., and Simon, M. I. (1987). Spatial relationship of the Fis binding sites for Hin recombinational enhancer activity. Nature 329, 462–465CrossRefGoogle ScholarPubMed
Jonsson, A. B., Nyberg, G., and Normark, S. (1991). Phase variation of gonococcal pili by frameshift mutation in pilC, a novel gene for pilus assembly. EMBO J. 10, 477–488Google ScholarPubMed
Joyce, S. A., and Clarke, D. J. (2003). A hexA homologue from Photorhabdus regulates pathogenicity, symbiosis and phenotypic variation. Mol. Microbiol. 47, 1445–1457CrossRefGoogle ScholarPubMed
Kanaar, R., Klippel, A., Shekhtman, E., Dungan, J. M., Kahmann, R., and Cozzarelli, N. R. (1990). Processive recombination by the phage Mu Gin system: Implications for the mechanisms of DNA strand exchange, DNA site alignment, and enhancer action. Cell 62, 353–366CrossRefGoogle ScholarPubMed
Karlyshev, A. V., Linton, D., Gregson, N. A., and Wren, B. W. (2002). A novel paralogous gene family involved in phase-variable flagella-mediated motility in Campylobacter jejuni. Microbiology 148, 473–480CrossRefGoogle ScholarPubMed
Kitten, T., and Barbour, A. G. (1990). Juxtaposition of expressed variable antigen genes with a conserved telomere in the bacterium Borrelia hermsii. Proc. Natl. Acad. Sci. U S A 87, 6077–6081CrossRefGoogle ScholarPubMed
Kitten, T., and Barbour, A. G. (1992). The relapsing fever agent Borrelia hermsii has multiple copies of its chromosome and linear plasmids. Genetics 132, 311–324Google ScholarPubMed
Kitten, T., Barrera, A. V., and Barbour, A. G. (1993). Intragenic recombination and a chimeric outer membrane protein in the relapsing fever agent Borrelia hermsii. J. Bacteriol. 175, 2516–2522CrossRefGoogle Scholar
Kjaergaard, K., Schembri, M. A., Ramos, C., Molin, S., and Klemm, P. (2000). Antigen 43 facilitates formation of multispecies biofilms. Environ. Microbiol. 2, 695–702CrossRefGoogle ScholarPubMed
Klemm, P. (1986). Two regulatory fim genes, fimB and fimE, control the phase variation of type 1 fimbriae in Escherichia coli. EMBO J. 5, 1389–1393Google ScholarPubMed
Komano, T., Kim, S. R., Yoshida, T., and Nisioka, T. (1994). DNA rearrangement of the shufflon determines recipient specificity in liquid mating of IncI1 plasmid R64. J. Mol. Biol. 243, 6–9CrossRefGoogle ScholarPubMed
Koomey, M., Gotschlich, E. C., Robbins, K., Bergstrom, S., and Swanson, J. (1987). Effects of recA mutations on pilus antigenic variation and phase transitions in Neisseria gonorrhoeae. Genetics 117, 391–398Google ScholarPubMed
Krasomil-Osterfeld, K. (1995). Influence of osmolarity on phase shift in Photorhabdus luminescens. Appl. Environ. Microbiol. 61, 3748–3749Google Scholar
Kulasekara, H. D., and Blomfield, I. C. (1999). The molecular basis for the specificity of fimE in the phase variation of type 1 fimbriae of Escherichia coli K-12. Mol. Microbiol. 31, 1171–1181CrossRefGoogle ScholarPubMed
Lenz, L. L., and Portnoy, D. A. (2002). Identification of a second Listeria secA gene associated with protein secretion and the rough phenotype. Mol. Microbiol. 45, 1043–1056CrossRefGoogle ScholarPubMed
Li, X., Lockatell, C. V., Johnson, D. E., and Mobley, H. L. (2002). Identification of MrpI as the sole recombinase that regulates the phase variation of MR/P fimbria, a bladder colonization factor of uropathogenic Proteus mirabilis. Mol. Microbiol. 45, 865–874CrossRefGoogle ScholarPubMed
Linton, D., Karlyshev, A. V., and Wren, B. W. (2001). Deciphering Campylobacter jejuni cell surface interactions from the genome sequence. Curr. Opin. Microbiol. 4, 35–40CrossRefGoogle ScholarPubMed
Liu, L., Panangala, V. S., and Dybvig, K. (2002). Trinucleotide GAA repeats dictate pMGA gene expression in Mycoplasma gallisepticum by affecting spacing between flanking regions. J. Bacteriol. 184, 1335–1339CrossRefGoogle ScholarPubMed
Low, D., Braaten, B., and van der Woude, M. (1996). Fimbriae. In Curtiss, R. III, Ingraham, J. L., Lin, E. C. C., Low, K. B., Magasanik, B., Reznikoff, W. S., Riley, M., Schaechter, M., and Umbarger, H. E. (Eds.), Escherichia coli and Salmonella. Cellular and molecular biology, 2nd ed., pp. 146–157. Washington, DC: ASM PressGoogle Scholar
Lysnyansky, I., Ron, Y., and Yogev, D. (2001). Juxtaposition of an active promoter to vsp genes via site-specific DNA inversions generates antigenic variation in Mycoplasma bovis. J. Bacteriol. 183, 5698–5708CrossRefGoogle ScholarPubMed
Lysnyansky, I., Sachse, K., Rosenbusch, R., Levisohn, S., and Yogev, D. (1999). The vsp locus of Mycoplasma bovis: Gene organization and structural features. J. Bacteriol. 181, 5734–5741Google ScholarPubMed
MacNab, R. M. (1996). Flagella and motility. In Curtiss, R. III, Ingraham, J. L., Lin, E. C. C., Low, K. B., Magasanik, B., Reznikoff, W. S., Riley, et al M.. (Eds.), Escherichia coli and Salmonella. Cellular and molecular biology, 2nd ed., pp. 123–145. Washington, DC: ASM PressGoogle Scholar
McClain, M. S., Blomfield, I. C., and Eisenstein, B. I. (1991). Roles of fimB and fimE in site-specific DNA inversion associated with phase variation of type 1 fimbriae in Escherichia coli. J. Bacteriol. 173, 5308–5314CrossRefGoogle ScholarPubMed
McKenney, D., Hubner, J., Muller, E., Wang, Y., Goldmann, D. A., and Pier, G. B. (1998). The ica locus of Staphylococcus epidermidis encodes production of the capsular polysaccharide/adhesin. Infect. Immun. 66, 4711–4720Google ScholarPubMed
McQuiston, J. H., McQuiston, J. R., Cox, A. D., Wu, Y., Boyle, S. M., and Inzana, T. J. (2000). Characterization of a DNA region containing 5′-(CAAT)(n)-3′ DNA sequences involved in lipooligosaccharide biosynthesis in Haemophilus somnus. Microb. Pathog. 28, 301–312CrossRefGoogle ScholarPubMed
Mehr, I. J., Long, C. D., Serkin, C. D., and Seifert, H. S. (2000). A homologue of the recombination-dependent growth gene, rdgC, is involved in gonococcal pilin antigenic variation. Genetics 154, 523–532Google ScholarPubMed
Mehr, I. J., and Seifert, H. S. (1998). Differential roles of homologous recombination pathways in Neisseria gonorrhoeae pilin antigenic variation, DNA transformation and DNA repair. Mol. Microbiol. 30, 697–710CrossRefGoogle ScholarPubMed
Meier, J. T., Simon, M. I., and Barbour, A. G. (1985). Antigenic variation is associated with DNA rearrangements in a relapsing fever Borrelia. Cell 41, 403–409CrossRefGoogle Scholar
Merickel, S. K., Haykinson, M. J., and Johnson, R. C. (1998). Communication between Hin recombinase and Fis regulatory subunits during coordinate activation of Hin-catalyzed site-specific DNA inversion. Genes Dev 12, 2803–2816CrossRefGoogle ScholarPubMed
Meyer, T. F., Gibbs, C. P., and Haas, R. (1990). Variation and control of protein expression in Neisseria. Annu. Rev. Microbiol. 44, 451–477CrossRefGoogle ScholarPubMed
Meyer, T. F., Pohlner, J., and Putten, J. P. (1994). Biology of the pathogenic Neisseriae. Curr. Top. Microbiol. Immunol. 192, 283–317Google ScholarPubMed
Monk, I. R., Cook, G. M., Monk, B. C., and Bremer, P. J. (2004). Morphotypic conversion in Listeria monocytogenes biofilm formation: biological significance of rough colony isolates. Appl. Environ. Microbiol. 70, 6686–6694CrossRefGoogle ScholarPubMed
Moxon, E. R., Rainey, P. B., Nowak, M. A., and Lenski, R. E. (1994). Adaptive evolution of highly mutable loci in pathogenic bacteria. Curr. Biol. 4, 24–33CrossRefGoogle ScholarPubMed
Murphy, C. A., and Belas, R. (1999). Genomic rearrangements in the flagellin genes of Proteus mirabilis. Mol. Microbiol. 31, 679–690CrossRefGoogle ScholarPubMed
Murphy, G. L., Connell, T. D., Barritt, D. S., Koomey, M., and Cannon, J. G. (1989). Phase variation of gonococcal protein II: Regulation of gene expression by slipped-strand mispairing of a repetitive DNA sequence. Cell 56, 539–547CrossRefGoogle ScholarPubMed
Nassif, X. (1999). Interaction mechanisms of encapsulated meningococci with eucaryotic cells: What does this tell us about the crossing of the blood–brain barrier by Neisseria meningitidis?Curr. Opin. Microbiol. 2, 71–77CrossRefGoogle ScholarPubMed
Nicholson, B., and Low, D. (2000). DNA methylation-dependent regulation of Pef expression in Salmonella typhimurium. Mol. Microbiol. 35, 728–742CrossRefGoogle ScholarPubMed
Nou, X., Braaten, B., Kaltenbach, L., and Low, D. A. (1995). Differential binding of Lrp to two sets of pap DNA binding sites mediated by Pap I regulates Pap phase variation in Escherichia coli. EMBO J. 14, 5785–5797Google Scholar
Nunes-Duby, S. E., Kwon, H. J., Tirumalai, R. S., Ellenberger, T., and Landy, A. (1998). Similarities and differences among 105 members of the Int family of site-specific recombinases. Nucleic Acids Res. 26, 391–406CrossRefGoogle ScholarPubMed
O'Gara, J. P., and Dorman, C. J. (2000). Effects of local transcription and H-NS on inversion of the fim switch of Escherichia coli. Mol. Microbiol. 36, 457–466CrossRefGoogle ScholarPubMed
O'Neill, K. H., Roche, D. M., Clarke, D. J., and Dowds, B. C. (2002). The ner gene of Photorhabdus: effects on primary-form-specific phenotypes and outer membrane protein composition. J. Bacteriol. 184, 3096–3105CrossRefGoogle ScholarPubMed
Owen, P., Meehan, M., Loughry-Doherty, H., and Henderson, I. (1996). Phase-variable outer membrane proteins in Escherichia coli. FEMS Immunol. Med. Microbiol. 16, 63–76CrossRefGoogle ScholarPubMed
Pallesen, L., Madsen, O., and Klemm, P. (1989). Regulation of the phase switch controlling expression of type 1 fimbriae in Escherichia coli. Mol. Microbiol. 3, 925–931CrossRefGoogle ScholarPubMed
Parkhill, J., Achtman, M., James, K. D., Bentley, S. D., Churcher, C., Klee, S. R., Morelli, G.. (2000a). Complete DNA sequence of a serogroup A strain of Neisseria meningitidis Z2491. Nature 404, 502–506CrossRefGoogle Scholar
Parkhill, J., Wren, B. W., Mungall, K., Ketley, J. M., Churcher, C., Basham, D., Chillingworth, T.. (2000b). The genome sequence of the food-borne pathogen Campylobacter jejuni reveals hypervariable sequences. Nature 403, 665–668CrossRefGoogle Scholar
Penningon, P. M., Cadavid, D., Bunikis, J., Norris, S. J., and Barbour, A. G. (1999). Extensive interplasmidic duplications change the virulence phenotype of the relapsing fever agent Borrelia turicatae. Mol. Microbiol. 34, 1120–1132CrossRefGoogle ScholarPubMed
Persson, A., Jacobsson, K., Frykberg, L., Johansson, K. E., and Poumarat, F. (2002). Variable surface protein Vmm of Mycoplasma mycoides subsp. mycoides small colony type. J. Bacteriol. 184, 3712–3722CrossRefGoogle ScholarPubMed
Plasterk, R. H., Simon, M. I., and Barbour, A. G. (1985). Transposition of structural genes to an expression sequence on a linear plasmid causes antigenic variation in the bacterium Borrelia hermsii. Nature 318, 257–263CrossRefGoogle Scholar
Prigent-Combaret, C., Vidal, O., Dorel, C., and Lejeune, P. (1999). Abiotic surface sensing and biofilm-dependent regulation of gene expression in Escherichia coli. J. Bacteriol. 181, 5993–6002Google ScholarPubMed
Restrepo, B. I., and Barbour, A. G. (1994). Antigen diversity in the bacterium B. hermsii through “somatic” mutations in rearranged vmp genes. Cell 78, 867–876CrossRefGoogle ScholarPubMed
Restrepo, B. I., Carter, C. J., and Barbour, A. G. (1994). Activation of a vmp pseudogene in Borrelia hermsii: An alternate mechanism of antigenic variation during relapsing fever. Mol. Microbiol. 13, 287–299CrossRefGoogle ScholarPubMed
Restrepo, B. I., Kitten, T., Carter, C. J., Infante, D., and Barbour, A. G. (1992). Subtelomeric expression regions of Borrelia hermsii linear plasmids are highly polymorphic. Mol. Microbiol. 6, 3299–3311CrossRefGoogle ScholarPubMed
Rocha, E. P., and Blanchard, A. (2002). Genomic repeats, genome plasticity and the dynamics of Mycoplasma evolution. Nucleic Acids Res. 30, 2031–2042CrossRefGoogle ScholarPubMed
Rocha, E. P., Matic, I., and Taddei, F. (2002). Over-representation of repeats in stress response genes: A strategy to increase versatility under stressful conditions?Nucleic Acids Res. 30, 1886–1894CrossRefGoogle ScholarPubMed
Roesch, P. L., and Blomfield, I. C. (1998). Leucine alters the interaction of the leucine-responsive regulatory protein (Lrp) with the fim switch to stimulate site-specific recombination in Escherichia coli. Mol. Microbiol. 27, 751–761CrossRefGoogle ScholarPubMed
Rohde, H., Knobloch, J. K., Horstkotte, M. A., and Mack, D. (2001). Correlation of biofilm expression types of Staphylococcus epidermidis with polysaccharide intercellular adhesin synthesis: Evidence for involvement of icaADBC genotype-independent factors. Med. Microbiol. Immunol. (Berl) 190, 105–112Google ScholarPubMed
Romling, U., Rohde, M., Olsen, A., Normark, S., and Reinkoster, J. (2000). AgfD, the checkpoint of multicellular and aggregative behaviour in Salmonella typhimurium regulates at least two independent pathways. Mol. Microbiol. 36, 10–23CrossRefGoogle ScholarPubMed
Romling, U., Sierralta, W. D., Eriksson, K., and Normark, S. (1998). Multicellular and aggregative behaviour of Salmonella typhimurium strains is controlled by mutations in the agfD promoter. Mol. Microbiol. 28, 249–264CrossRefGoogle ScholarPubMed
Ron, Y., Flitman-Tene, R., Dybvig, K., and Yogev, D. (2002). Identification and characterization of a site-specific tyrosine recombinase within the variable loci of Mycoplasma bovis, Mycoplasma pulmonis and Mycoplasma agalactiae. Gene 292, 205–211CrossRefGoogle ScholarPubMed
Rozsa, F. W., and Marrs, C. F. (1991). Interesting sequence differences between the pilin gene inversion regions of Moraxella lacunata ATCC 17956 and Moraxella bovis Epp63. J. Bacteriol. 173, 4000–4006CrossRefGoogle ScholarPubMed
Rudel, T., Facius, D., Barten, R., Scheuerpflug, I., Nonnenmacher, E., and Meyer, T. F. (1995a). Role of pili and the phase-variable PilC protein in natural competence for transformation of Neisseria gonorrhoeae. Proc. Natl. Acad. Sci. U S A 92, 7986–7990CrossRefGoogle Scholar
Rudel, T., Scheurerpflug, I., and Meyer, T. F. (1995b). Neisseria PilC protein identified as type-4 pilus tip-located adhesin. Nature 373, 357–359CrossRefGoogle Scholar
Rudel, T., Putten, J. P., Gibbs, C. P., Haas, R., and Meyer, T. F. (1992). Interaction of two variable proteins (PilE and PilC) required for pilus-mediated adherence of Neisseria gonorrhoeae to human epithelial cells. Mol. Microbiol. 6, 3439–3450CrossRefGoogle ScholarPubMed
Rytkonen, A., Johansson, L., Asp, V., Albiger, B., and Jonsson, A. B. (2001). Soluble pilin of Neisseria gonorrhoeae interacts with human target cells and tissue. Infect. Immun. 69, 6419–6426CrossRefGoogle Scholar
Sarkari, J., Pandit, N., Moxon, E. R., and Achtman, M. (1994). Variable expression of the Opc outer membrane protein in Neisseria meningitidis is caused by size variation of a promoter containing poly-cytidine. Mol. Microbiol. 13, 207–217CrossRefGoogle ScholarPubMed
Sauer, K., Camper, A. K., Ehrlich, G. D., Costerton, J. W., and Davies, D. G. (2002). Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J. Bacteriol. 184, 1140–1154CrossRefGoogle ScholarPubMed
Saunders, N. J., Jeffries, A. C., Peden, J. F., Hood, D. W., Tettelin, H., Rappuoli, R., and Moxon, E. R. (2000). Repeat-associated phase variable genes in the complete genome sequence of Neisseria meningitidis strain MC58. Mol. Microbiol. 37, 207–215CrossRefGoogle ScholarPubMed
Schembri, M. A., and Klemm, P. (2001). Coordinate gene regulation by fimbriae-induced signal transduction. EMBO J. 20, 3074–3081CrossRefGoogle ScholarPubMed
Schembri, M. A., Hjerrild, L., Gjermansen, M., and Per Klemm, P. (2003). Differential expression of the Escherichia coli autoaggregation factor antigen 43. J. Bacteriol. 185, 2236–2242CrossRefGoogle ScholarPubMed
Schwan, T. G., and Hinnebusch, B. J. (1998). Bloodstream- versus tick-associated variants of a relapsing fever bacterium. Science 280, 1938–1940CrossRefGoogle ScholarPubMed
Schwan, W. R., Lee, J. L., Lenard, F. A., Matthews, B. T., Beck, M. T. (2002). Osmolarity and pH growth conditions regulate fim gene transcription and type 1 pilus expression in uropathogenic Escherichia coli. Infect Immun. 70, 1391–1402CrossRefGoogle ScholarPubMed
Seifert, H. S. (1996). Questions about gonococcal pilus phase and antigenic variation. Mol. Microbiol. 21, 433–440CrossRefGoogle ScholarPubMed
Serkin, C. D., and Seifert, H. S. (1998). Frequency of pilin antigenic variation in Neisseria gonorrhoeae. J. Bacteriol. 180, 1955–1958Google ScholarPubMed
Shen, X., Gumulak, J., Yu, H., French, C. T., Zou, N., and Dybvig, K. (2000). Gene rearrangements in the vsa locus of Mycoplasma pulmonis. J. Bacteriol. 182, 2900–2908CrossRefGoogle ScholarPubMed
Sleytr, U. B., Messner, P., Pum, D., and Sara, M. (1993). Crystalline bacterial cell surface layers. Mol. Microbiol. 10, 911–916CrossRefGoogle ScholarPubMed
Smigielski, A. J., Akhurst, R. J., and Boemare, N. E. (1994). Phase variation in Xenorhabdus nematophilus and Photorhabdus luminescens: Differences in respiratory activity and membrane energization. Appl. Environ. Microbiol. 60, 120–125Google ScholarPubMed
Smith, S. G., and Dorman, C. J. (1999). Functional analysis of the FimE integrase of Escherichia coli K-12: Isolation of mutant derivatives with altered DNA inversion preferences. Mol. Microbiol. 34, 965–979CrossRefGoogle ScholarPubMed
Snyder, L. A., Butcher, S. A., and Saunders, N. J. (2001). Comparative whole-genome analyses reveal over 100 putative phase-variable genes in the pathogenic Neisseria spp. Microbiology 147, 2321–2332CrossRefGoogle ScholarPubMed
Stern, A., Brown, M., Nickel, P., and Meyer, T. F. (1986). Opacity genes in Neisseria gonorrhoeae: Control of phase and antigenic variation. Cell 47, 61–71CrossRefGoogle ScholarPubMed
Stoenner, H. G., Dodd, T., and Larsen, C. (1982). Antigenic variation of Borrelia hermsii. J. Exp. Med. 156, 1297–1311CrossRefGoogle ScholarPubMed
Stoodley, P., Sauer, K., Davies, D. G., and Costerton, J. W. (2002). Biofilms as complex differentiated communities. Annu. Rev. Microbiol. 56, 187–209CrossRefGoogle ScholarPubMed
Storz, G., and Imlay, J. A. (1999). Oxidative stress. Curr. Opin. Microbiol. 2, 188–194CrossRefGoogle ScholarPubMed
Taddei, F., Radman, M., Maynard-Smith, J., Toupance, B., Gouyon, P. H., and Godelle, B. (1997). Role of mutator alleles in adaptive evolution. Nature 387, 700–702CrossRefGoogle ScholarPubMed
Tobiason, D. M., Lenich, A. G., and Glasgow, A. C. (1999). Multiple DNA binding activities of the novel site-specific recombinase, Piv, from Moraxella lacunata. J. Biol. Chem. 274, 9698–9706CrossRefGoogle ScholarPubMed
Tomb, J. F., White, O., Kerlavage, A. R., Clayton, R. A., Sutton, G. G., Fleischmann, R. D., Ketchum, K. A.. (1997). The complete genome sequence of the gastric pathogen Helicobacter pylori. Nature 388, 539–547CrossRefGoogle ScholarPubMed
Belkum, A., Scherer, S., Alphen, L., and Verbrugh, H. (1998). Short-sequence DNA repeats in prokaryotic genomes. Microbiol. Mol. Biol. Rev. 62, 275–293Google ScholarPubMed
Putte, P., and Goosen, N. (1992). DNA inversions in phages and bacteria. Trends Genet. 8, 457–462CrossRefGoogle ScholarPubMed
Ende, A., Hopman, C. T., Zaat, S., Essink, B. B., Berkhout, B., and Dankert, J. (1995). Variable expression of class 1 outer membrane protein in Neisseria meningitidis is caused by variation in the spacing between the –10 and –35 regions of the promoter. J. Bacteriol. 177, 2475–2480CrossRefGoogle ScholarPubMed
Ham, S. M., Alphen, L., Mooi, F. R., and Putten, J. P. (1993). Phase variation of H. influenzae fimbriae: Transcriptional control of two divergent genes through a variable combined promoter region. Cell 73, 1187–1196Google ScholarPubMed
Wainwright, L. A., Frangipane, J. V., and Seifert, H. S. (1997). Analysis of protein binding to the Sma/Cla DNA repeat in pathogenic Neisseriae. Nucleic Acids Res. 25, 1362–1368CrossRefGoogle ScholarPubMed
Wainwright, L. A., Pritchard, K. H., and Seifert, H. S. (1994). A conserved DNA sequence is required for efficient gonococcal pilin antigenic variation. Mol. Microbiol. 13, 75–87CrossRefGoogle ScholarPubMed
Waldron, D. E., Owen, P., and Dorman, C. J. (2002). Competitive interaction of the OxyR DNA-binding protein and the Dam methylase at the antigen 43 gene regulatory region in Escherichia coli. Mol. Microbiol. 44, 509–520CrossRefGoogle ScholarPubMed
Wallecha, A., Correnti, J., Munster, V., and Woude, M. (2003). Phase variation of Ag43 is independent of the oxidation state of OxyR. J. Bacteriol. 185, 2203–2209CrossRefGoogle ScholarPubMed
Wallecha, A., Munster, V., Correnti, J., Chan, T., and Woude, M. (2002). Dam- and OxyR-dependent phase variation of agn43: Essential elements and evidence for a new role of DNA methylation. J. Bacteriol. 184, 3338–3347CrossRefGoogle ScholarPubMed
Wang, G., Ge, Z., Rasko, D. A., and Taylor, D. E. (2000). Lewis antigens in Helicobacter pylori: Biosynthesis and phase variation. Mol. Microbiol. 36, 1187–1196CrossRefGoogle ScholarPubMed
White-Ziegler, C. A., Angus Hill, M. L., Braaten, B. A., Woude, M. W., and Low, D. A. (1998). Thermoregulation of Escherichia coli pap transcription: H-NS is a temperature-dependent DNA methylation blocking factor. Mol. Microbiol. 28, 1121–1137CrossRefGoogle ScholarPubMed
White-Ziegler, C. A., Black, A. M., Eliades, S. H., Young, S., and Porter, K. (2002). The N-acetyltransferase RimJ responds to environmental stimuli to repress pap fimbrial transcription in Escherichia coli. J. Bacteriol. 184, 4334–4342CrossRefGoogle ScholarPubMed
White-Ziegler, C. A., Blyn, L. B., Braaten, B. A., and Low, D. A. (1990). Identification of an Escherichia coli genetic locus involved in thermoregulation of the pap operon. J. Bacteriol. 172, 1775–1782CrossRefGoogle ScholarPubMed
White-Ziegler, C. A., and Low, D. A. (1992). Thermoregulation of the pap operon: Evidence for the involvement of RimJ, the N-terminal acetylase of ribosomal protein S5. J. Bacteriol. 174, 7003–7012CrossRefGoogle ScholarPubMed
Whiteley, M., Bangera, M. G., Bumgarner, R. E., Parsek, M. R., Teitzel, G. M., Lory, S., and Greenberg, E. P. (2001). Gene expression in Pseudomonas aeruginosa biofilms. Nature 413, 860–864CrossRefGoogle ScholarPubMed
Wu, Y., McQuiston, J. H., Cox, A., Pack, T. D., and Inzana, T. J. (2000). Molecular cloning and mutagenesis of a DNA locus involved in lipooligosaccharide biosynthesis in Haemophilus somnus. Infect. Immun. 68, 310–319CrossRefGoogle ScholarPubMed
Xia, Y., Gally, D., Forsman-Semb, K., and Uhlin, B. E. (2000). Regulatory cross-talk between adhesin operons in Escherichia coli: Inhibition of type 1 fimbriae expression by the PapB protein. EMBO J. 19, 1450–1457CrossRefGoogle ScholarPubMed
Yang, Q. L., and Gotschlich, E. C. (1996). Variation of gonococcal lipooligosaccharide structure is due to alterations in poly-G tracts in lgt genes encoding glycosyl transferases. J. Exp. Med. 183, 323–327CrossRefGoogle ScholarPubMed
Yildiz, F. H., Liu, X. S., Heydorn, A., and Schoolnik, G. K. (2004). Molecular analysis of rugosity in a Vibrio cholerae O1 El Tor phase variant. Mol. Microbiol. 53, 497–515CrossRefGoogle Scholar
Yildiz, F. H., Dolganov, N. A., and Schoolnik, G. K. (2001). VpsR, a member of the response regulators of the two-component regulatory systems, is required for expression of vps biosynthesis genes and EPS(ETr)-associated phenotypes in Vibrio cholerae O1 El Tor. J. Bacteriol. 183, 1716–1726CrossRefGoogle ScholarPubMed
Yildiz, F. H., and Schoolnik, G. K. (1999). Vibrio cholerae O1 El Tor: Identification of a gene cluster required for the rugose colony type, exopolysaccharide production, chlorine resistance, and biofilm formation. Proc. Natl. Acad. Sci. U S A 96, 4028–4033CrossRefGoogle ScholarPubMed
Yogev, D., Rosengarten, R., Watson-McKown, R., and Wise, K. S. (1991). Molecular basis of Mycoplasma surface antigenic variation: A novel set of divergent genes undergo spontaneous mutation of periodic coding regions and 5′ regulatory sequences. EMBO J. 10, 4069–4079Google ScholarPubMed
Yoshida, T., Furuya, N., Ishikura, M., Isobe, T., Haino-Fukushima, K., Ogawa, T., and Komano, T. (1998). Purification and characterization of thin pili of IncI1 plasmids ColIb-P9 and R64: Formation of PilV-specific cell aggregates by type IV pili. J. Bacteriol. 180, 2842–2848Google ScholarPubMed
Zachar, Z., and Savage, D. C. (1979). Microbial interference and colonization of the murine gastrointestinal tract by Listeria monocytogenes. Infect. Immun. 23, 168–174Google ScholarPubMed
Zhang, Q., and Wise, K. S. (1997). Localized reversible frameshift mutation in an adhesin gene confers a phase-variable adherence phenotype in mycoplasma. Mol. Microbiol. 25, 859–869CrossRefGoogle Scholar
Zhao, H., Li, X., Johnson, D. E., Blomfield, I., and Mobley, H. L. (1997). In vivo phase variation of MR/P fimbrial gene expression in Proteus mirabilis infecting the urinary tract. Mol. Microbiol. 23, 1009–1019CrossRefGoogle ScholarPubMed
Zheng, M., Aslund, F., and Storz, G. (1998). Activation of the OxyR transcription factor by reversible disulfide bond formation. Science 279, 1718–1721CrossRefGoogle ScholarPubMed
Ziebuhr, W., Heilmann, C., Gotz, F., Meyer, P., Wilms, K., Straube, E., and Hacker, J. (1997). Detection of the intercellular adhesion gene cluster (ica) and phase variation in Staphylococcus epidermidis blood culture strains and mucosal isolates. Infect. Immun. 65, 890–896Google ScholarPubMed
Ziebuhr, W., Krimmer, V., Rachid, S., Lossner, I., Gotz, F., and Hacker, J. (1999). A novel mechanism of phase variation of virulence in Staphylococcus epidermidis: Evidence for control of the polysaccharide intercellular adhesin synthesis by alternating insertion and excision of the insertion sequence element IS256. Mol. Microbiol. 32, 345–356CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×