Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-27T01:16:38.053Z Has data issue: false hasContentIssue false

8 - Free-Living to Freewheeling: The Evolution of Vibrio cholerae from Innocence to Infamy

Published online by Cambridge University Press:  10 August 2009

Rita R. Colwell
Affiliation:
Center of Marine Biotechnology, University of Maryland Biotechnology Institute, 701 East Pratt Street, Baltimore, MD 21202
Shah M. Faruque
Affiliation:
International Centre for Diarrhoeal Disease, Bangladesh, Dhaka, Bangladesh
G. Balakrish Nair
Affiliation:
International Centre for Diarrhoeal Disease, Bangladesh, Dhaka, Bangladesh
Krishna R. Dronamraju
Affiliation:
Foundation for Genetic Research, Houston, Texas
Get access

Summary

INTRODUCTION

Toxigenic strains of the Gram-negative bacterium Vibrio cholerae cause the most severe form of dehydrating diarrhea known as cholera and have been responsible for at least seven distinct pandemics of cholera (Faruque et al., 1998a). Since the first recorded pandemic, which began in 1817, V. cholerae strains associated with different pandemics are assumed to have undergone phenotypic and genetic changes with time. For example, although the seventh pandemic was caused by the El Tor biotype of V. cholerae, the sixth and possibly earlier pandemics were caused by the classical biotype. The genetic changes in V. cholerae associated with epidemics and pandemics, however, were not fully appreciated until the development and application of molecular techniques to analyze strains.

Recent application of molecular approaches has enabled extraordinary progress in our understanding of the evolution of V. cholerae. Like other bacteria, V. cholerae can be assumed to have existed well before human evolution and therefore to have originated primarily as a free-living aquatic microorganism. The fact that over a period of a million years some clones (serogroups) of this organism have acquired the ability to colonize transiently the human intestine and become an efficient human pathogen reflects the progressive acquisition of the genetic capability to become pathogenic for humans. They could have acquired this capability by embracing another “life style,” albeit accidentally, by carrying out a function in the human host that it performs either symbiotically or commensally for its nonhuman host but is, in effect, pathogenic to the human host, or alternatively, by finding a niche whereby the organism could amplify and perpetuate itself as effectively or more effectively, than it could in a free-living state.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2004

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

Aldova, E., Laznickova, K., Stepankova, , , E., and Lietava, J. (1968). Isolation of nonagglutinable vibrios from an enteritis outbreak in Czechoslovakia. Journal of Infectious Diseases, 118, 25–31CrossRefGoogle ScholarPubMed
Beltran, P., Delgado, G., Davarro, A., Trujillo, F., Selander, R. K., and Cravioto, A. (1999). Genetic diversity and population structure of Vibrio cholerae. Journal of Clinical Microbiology, 37, 581–90Google ScholarPubMed
Berche, P., Poyart, C., Abachin, E., Lelievre, H., Vandepitte, J., Dodin, A., and Fournier, J. M. (1994). The novel epidemic strain O139 is closely related to the pandemic strain O1 of Vibrio cholerae. Journal of Infectious Diseases, 170, 701–4CrossRefGoogle ScholarPubMed
Bik, E. M., Bunschoten, A. E., Gouw, R. D., and Mooi, F. R. (1995). Genesis of the novel epidemic Vibrio cholerae O139 strain: evidence for horizontal transfer of genes involved in polysaccharide synthesis. European Molecular Biology Organization Journal, 14, 209–16Google ScholarPubMed
Bik, E. M., Gouw, R. D., and Mooi, F. R. (1996). DNA fingerprinting of Vibrio cholerae strains with a novel insertion sequence element: a tool to identify epidemic strains. Journal of Clinical Microbiology, 34, 1453–61Google ScholarPubMed
Blank, T. E., Zhong, H., Bell, A. L., Whittam, T. S., and Donnenberg, M. S. (2000). Molecular variation among type IV pilin (bfpA) genes from diverse enteropathogenic Escherichia coli strains. Infection and Immunity, 68, 7028–38CrossRefGoogle ScholarPubMed
Bowler, L. D., Zhang, Q. Y., Riou, J. Y., and Spratt, B. G. (1994). Interspecies recombination between the penA genes of Neisseria meningitidis and commensal Neisseria species during the emergence of penicillin resistance in N. meningitidis: natural events and laboratory simulation. Journal of Bacteriology, 176, 333–7CrossRefGoogle ScholarPubMed
Boyd, E. F., Moyer, K. E., Shi, L., and Waldor, M. K. (2000). Infectious CTXΦ and the Vibrio pathogenicity island prophage in Vibrio mimicus: evidence for recent horizontal transfer between V. mimicus and V. cholerae. Infection and Immunity, 68, 1507–13CrossRefGoogle ScholarPubMed
Boyd, E. F., and Waldor, M. K. (2002). Evolutionary and functional analyses of variants of the toxin-coregulated pilus protein TcpA from toxigenic Vibrio cholerae non-O1/non-O139 serogroup isolates. Microbiology, 148, 1655–66CrossRefGoogle ScholarPubMed
Chakrabarti, A. K., Ghosh, A. N., Nair, G. B., Niyogi, S. K., Bhattacharya, S. K., and Sarkar, B. L. (2000). Development and evaluation of a phage typing scheme for Vibrio cholerae O139. Journal of Clinical Microbiology, 38, 44–9Google ScholarPubMed
Chakraborty, S., Mukhopadhyay, A. K., Bhadra, R. K., Ghosh, A. N., Mitra, R., Shimada, T., Yamasaki, S., Faruque, S. M., Takeda, Y., Colwell, R. R., and Nair, G. B. (2000). Virulence genes in environmental strains of Vibrio cholerae. Applied and Environmental Microbiology, 66, 4022–8CrossRefGoogle ScholarPubMed
Chakraborty, S., Garg, P., Ramamurthy, T., Thungapathra, M., Gautam, J. K., Kumar, C., Maiti, S., Yamasaki, S., Shimada, T., Takeda, Y., Ghosh, A., and Nair, G. B. (2001). Comparison of antibiogram, virulence genes, ribotypes and DNA fingerprints of Vibrio cholerae of matching serogroups isolated from hospitalised diarrhoea cases and from the environment during 1997–1998 in Calcutta. India Journal of Medicine Microbiology, 50, 879–88CrossRefGoogle ScholarPubMed
Chattopadhyay, D. J., Sarkar, B. L., Ansari, M. Q., Chakrabarti, B. K., Roy, M. K., Ghosh, A. N., and Pal, S. C. (1993). New phage typing scheme for Vibrio cholerae O1 biotype El Tor strains. Journal of Clinical Microbiology, 31, 1579–85Google ScholarPubMed
Chen, F., Evins, G. M., Cook, W. L., Almeida, R., Bean, H. N., and Wachsmuth, I. K. (1991). Genetic diversity among toxigenic and non-toxigenic Vibrio cholerae-O1 isolated from the western hemisphere. Epidemiology and Infection, 107, 225–33CrossRefGoogle ScholarPubMed
Cholera Working Group. (1993). Large epidemic of cholera-like disease in Bangladesh caused by Vibrio cholerae O139 synonym Bengal. Lancet, 342, 387–90CrossRef
Chowdhury, N. R., Chakraborty, S., Ramamurthy, T., Nishibuchi, M., Yamasaki, S., Takeda, Y., and Nair, G. B. (2000). Molecular evidence of clonal Vibrio parahaemolyticus pandemic strains. Emerging Infectious Diseases, 6, 631–6CrossRefGoogle ScholarPubMed
Citarella, R. V., and Colwell, R. R. (1970). Polyphasic taxonomy of the genus Vibrio: polynucleotide sequence relationships among selected Vibrio species. Journal of Bacteriology, 104, 434–42Google ScholarPubMed
Colwell, R. R. (1973). Genetic and phenetic classification of bacteria. Advances in Applied Microbiology, 16, 137–76CrossRefGoogle Scholar
Colwell, R. R. (1996). Global climate and infectious disease: the cholera paradigm. Science, 274, 2025–31CrossRefGoogle ScholarPubMed
Colwell, R. R., and Grimes, D. J. (2000). Nonculturable Microorganisms in the Environment. American Society of Microbiology Press, Washington, DC, 354 pp
Colwell, R. R., Kaper, J., and Joseph, S. W. (1977). Vibrio cholerae, Vibrio parahaemolyticus, and other vibrios: occurrence and distribution in Chesapeake Bay. Science, 198, 394–6CrossRefGoogle ScholarPubMed
Colwell, R. R., Brayton, P. R., Grimes, D. J., Roszak, D. R., Huq, S. A., and Palmer, L. M. (1985). Viable but non-culturable Vibrio cholerae and related environmental pathogens in the environment: implication for release of genetically engineered microorganisms. Bio/Technology, 3, 817–20Google Scholar
Comstock, L. E., Johnson, J. A., Michalski, J. M., Morris, J. G. Jr., and Kaper, J. B. (1996). Cloning and sequence of a region encoding a surface polysaccharide of Vibrio cholerae O139 and characterization of the insertion site in the chromosome of Vibrio cholerae O1. Molecular Microbiology, 19, 815–26CrossRefGoogle ScholarPubMed
Dakin, W. P., Howell, D. J., Sutton, R. G., O'Keefe, M. F., and Thomas, P. (1974). Gastroenteritis due to non-agglutinable (non-cholera) vibrios. Medical Journal of Australia, 2, 487–90Google ScholarPubMed
Davis, B. M., Kimsey, H. H., Chang, W., and Waldor, M. K. (1999). The Vibrio cholerae O139 Calcutta bacteriophage CTXΦ is infectious and encodes a novel repressor. Journal of Bacteriology, 181, 6779–87Google ScholarPubMed
DiRita, V. J. (1992). Co-ordinate expression of virulence genes by ToxR in Vibrio cholerae. Molecular Microbiology, 6, 451–8CrossRefGoogle ScholarPubMed
DiRita, V. J., Parsot, C., Jander, G., and Mekalanos, J. J. (1991). Regulatory cascade controls virulence in Vibrio cholerae. Proceedings of the National Academy of Sciences USA, 88, 5403–7CrossRefGoogle ScholarPubMed
DuBose, R. F., Dykhuizen, D. E., and Hartl, D. L. (1988). Genetic exchange among natural isolates of bacteria: recombination within the phoA gene of Escherichia coli. Proceedings of the National Academy of Sciences USA, 85, 7036–40CrossRefGoogle ScholarPubMed
Dumontier, S., and Berche, P. (1998). Vibrio cholerae O22 might be a putative source of exogenous DNA resulting in the emergence of the new strain of Vibrio cholerae O139. FEMS Microbiology Letters, 164, 91–8CrossRefGoogle ScholarPubMed
Faruque, S. M., Roy, S. K., Alim, A. R. M. A., Siddique, A. K., and Albert, M. J. (1995). Molecular epidemiology of toxigenic Vibrio cholerae in Bangladesh studied by numerical analysis of rRNA gene restriction patterns. Journal of Clinical Microbiology, 33, 2833–8Google Scholar
Faruque, S. M., Ahmed, K. M., Alim, A. R. M. A., Qadri, F., Siddique, A. K., and Albert, M. J. (1997 a). Emergence of a new clone of toxigenic Vibrio cholerae biotype El Tor displacing V. cholerae O139 Bengal in Bangladesh. Journal of Clinical Microbiology, 35, 624–30Google ScholarPubMed
Faruque, S. M., Ahmed, K. M., Siddique, A. K., Zaman, K., Alim, A. R. M. A., and Albert, M. J. (1997 b). Molecular analysis of toxigenic Vibrio cholerae O139 Bengal isolated in Bangladesh between 1993 and 1996: evidence for the emergence of a new clone of the Bengal vibrios. Journal of Clinical Microbiology, 35, 2299–306Google ScholarPubMed
Faruque, S. M., Albert, M. J., and Mekalanos, J. J. (1998 a). Epidemiology, genetics, and ecology of toxigenic Vibrio cholerae. Microbiology and Molecular Biology Reviews, 62, 1301–14Google ScholarPubMed
Faruque, S. M., Asadulghani, S.Saha, M. N., Alim, A. R. M. A., Albert, M. J., Islam, K. M. N., and Mekalanos, J. J. (1998 b). Analysis of clinical and environmental strains of nontoxigenic Vibrio cholerae for susceptibility to CTXΦ: molecular basis for origination of new strains with epidemic potential. Infection and Immunity, 66, 5819–25Google ScholarPubMed
Faruque, S. M., Asadulghani, S.Alim, A. R. M. A., Albert, M. J., Islam, K. M. N., and Mekalanos, J. J. (1998 c). Induction of the lysogenic phage encoding cholera toxin in naturally occurring strains of toxigenic Vibrio cholerae O1 and O139. Infection and Immununity, 66, 3752–7Google ScholarPubMed
Faruque, S. M., Saha, M. N., Bag, Asadulghani P. K., Bhadra, R. K., Bhattacharya, S. K., Sack, R. B., Takeda, Y., and Nair, G. B. (2000). Genomic diversity among Vibrio cholerae O139 strains isolated in Bangladesh and India between 1992 and 1998. FEMS Microbiology Letters, 84, 279–84CrossRefGoogle Scholar
Field, M., Fromm, D., Al-Awqati, Q., and Greenough, W. B. III. (1972). Effect of cholera enterotoxin on ion transport across isolated ileal mucosa. Journal of Clinical Investigation, 51, 796–804CrossRefGoogle ScholarPubMed
Finkelstein, R. A. (1983). Antigenic and structural variations in the cholera/coli family of enterotoxins. Developments in Biological Standardization, 53, 93–5Google ScholarPubMed
Franzon, V. L., Barker, A., and Manning, P. A. (1993). Nucleotide sequence encoding the mannose-fucose-resistant hemagglutinin of Vibrio cholerae O1 and construction of a mutant. Infection and Immununity, 61, 3032–7Google ScholarPubMed
Freitas, F. S., Momen, H., and Salles, C. A. (2002). The zymovars of Vibrio cholerae: multilocus enzyme electrophoresis of Vibrio cholerae. Memórias do Instituto Oswaldo Cruz, 97, 511–6CrossRefGoogle ScholarPubMed
Goldberg, I., and Mekalanos, J. J. (1986). Effect of a recA mutation on cholera toxin gene amplification and deletion events. Journal of Bacteriology, 165, 723–31CrossRefGoogle ScholarPubMed
Hase, C. C., and Mekalanos, J. J. (1998). TcpP protein is a positive regulator of virulence gene expression in Vibrio cholerae. Proceedings of the National Academy of Sciences USA, 95, 730–4CrossRefGoogle ScholarPubMed
Heidelberg, J. F., Eisen, J. A., Nelson, W. C., Clayton, R. A., Gwinn, M. L., Dodson, R. J., Haft, D. H., Hickey, E. K., Peterson, J. D., Umayam, L., Gill, S. R., Nelson, K. E., Read, T. D., Tettelin, H., Richardson, D., Ermolaeva, M. D., Vamathevan, J., Bass, S., Qin, H., Dragoi, I., Sellers, P., McDonald, L., Utterback, T., Fleishman, R. D., Nierman, W. C., White, O., Salzberg, S. L., Smith, H. O., Colwell, R. R., Mekalanos, J. J., Ventor, J. C., and Frasier, C. M. (2000). DNA sequence of both chromosomes of the cholera pathogen Vibrio cholerae. Nature, 406, 477–83Google ScholarPubMed
Heilpern, A. J, and Waldor, M. K. (2000). CTXΦ infection of Vibrio cholerae requires the tolQRA gene products. Journal of Bacteriology, 182, 1739–47CrossRefGoogle ScholarPubMed
Herrington, D. A., Hall, R. H., Losonsky, G., Mekalanos, J. J., Taylor, R. K., and Levine, M. M. (1988). Toxin, toxin-coregulated pili and ToxR regulon are essential for Vibrio cholerae pathogenesis in humans. Journal of Experimental Medicine, 168, 1487–92CrossRefGoogle ScholarPubMed
Hughes, K. J., Everiss, K. D., Harkey, C. W., and Peterson, K. M. (1994). Identification of a Vibrio cholerae ToxR-activated gene (tagD) that is physically linked to the toxin coregulated pilus (tcp) gene cluster. Gene, 148, 97–100CrossRefGoogle ScholarPubMed
Islam, M. S., Drasar, B. S., and Sack, R. B. (1994 a). The aquatic flora and fauna as reservoirs of Vibrio cholerae: a review. Journal of Diarrhoeal Disease Research, 12, 87–96Google ScholarPubMed
Islam, M. S., Drasar, B. S., and Sack, R. B. (1994 b). The aquatic environment as a reservoir of Vibrio cholerae a review. Journal of Diarrhoeal Disease Research, 11, 197–206Google Scholar
Johnson, J. A., Salles, C. A., Panigrahi, P., Albert, M. J., Wright, A. C., Johnson, R. J., and Morris, J. G. Jr. (1994). Vibrio cholerae O139 synonym Bengal is closely related to Vibrio cholerae El Tor but has important differences. Infection and Immunity, 62, 2108–10Google ScholarPubMed
Jonson, G., Lebens, M., and Holmgren, J. (1994). Cloning and sequencing of Vibrio cholerae mannose-sensitive haemagglutinin pilin gene: localization of mshA within a cluster of type IV pilin genes. Molecular Microbiology, 13, 109–18CrossRefGoogle Scholar
Karaolis, D. K., Johnson, J. A., Bailey, C. C., Boedeker, E. C., Kaper, J. B., and Reeves, P. R. (1998). A Vibrio cholerae pathogenicity island associated with epidemic and pandemic strains. Proceedings of the National Academy of Sciences USA, 95, 3134–9CrossRefGoogle ScholarPubMed
Karaolis, D. K., Somara, S., Maneval, D. R. Jr., Johnson, J. A., and Kaper, J. B. (1999). A bacteriophage encoding a pathogenicity island, a type-IV pilus and a phage receptor in cholera bacteria. Nature, 399, 375–9CrossRefGoogle Scholar
Karaolis, D. K. R., Lan, R., Kaper, J. B., and Reeves, P. R. (2001). Comparison of Vibrio cholerae pathogenicity islands in sixth and seventh pandemic strains. Infection and Immunity, 69, 1947–52CrossRefGoogle ScholarPubMed
Kaufman, M. R., Shaw, C. E., Jones, I. D., and Taylor, R. K. (1993). Biogenesis and regulation of the Vibrio cholerae toxin-coregulated pilus: analogies to other virulence factor secretory systems. Gene, 126, 43–9CrossRefGoogle ScholarPubMed
Kelly, M. T. (1991). Pathogenic Vibrionaceae in patients and the environment. Undersea Biomedical Research, 18, 193–6Google ScholarPubMed
Kimsey, H. H., and Waldor, M. K. (1998). CTXφ immunity: application in the development of cholera vaccines. Proceedings of the National Academy of Sciences USA, 95, 7035–9CrossRefGoogle ScholarPubMed
Kimsey, H. H., Nair, G. B., Ghosh, A., and Waldor, M. K. (1998). Diverse CTXphis and evolution of new pathogenic Vibrio cholerae. Lancet, 352, 457–8CrossRefGoogle ScholarPubMed
Kirn, T. J., Lafferty, M. J., Sandoe, C. M., and Taylor, R. K. (2000). Delineation of pilin domains required for bacterial association into microcolonies and intestinal colonization by Vibrio cholerae. Molecular Microbiology, 35, 896–910CrossRefGoogle ScholarPubMed
Kovach, M. E., Shaffer, M. D., and Peterson, K. M. (1996). A putative integrase gene defines the distal end of a large cluster of toxR-regulated colonization genes in Vibrio cholerae. Microbiology, 142, 2165–74CrossRefGoogle ScholarPubMed
Lawrence, J. G., and Ochman, H. (1998). Molecular archaeology of the Escherichia coli genome. Proceedings of the National Academy of Sciences USA, 95, 9413–17CrossRefGoogle ScholarPubMed
Lazar, S., and Waldor, M. K. (1998). ToxR-independent expression of cholera toxin from the replicative form of CTXΦ. Infection and Immunity, 66, 394–7Google ScholarPubMed
Lipp, E. K., Huq, A., and Colwell, R. R. (2002). Effects of global climate on infectious disease: the cholera model. Clinical Microbiology Reviews, 15, 757–70CrossRefGoogle ScholarPubMed
Maiden, M. C., Bygraves, J. A., Feil, E., Morelli, G., Russell, J. E., Urwin, R., Zhang, Q., Zhou, J., Zurth, K., Caugant, D. A., Feavers, I. M., Achtman, M., and Spratt, B. G. (1998). Multilocus sequence typing: a portable approach to the identification of clones within populations of pathogenic microorganisms. Proceedings of the National Academy of Sciences USA, 95, 3140–5CrossRefGoogle ScholarPubMed
Manning, P. A. (1997). The tcp gene cluster of Vibrio cholerae. Gene, 192, 63–70CrossRefGoogle ScholarPubMed
Manning, P. A., Stroeher, U. H., and Morona, R. (1994). Molecular basis for O-antigen biosynthesis in Vibrio cholerae: Ogawa-Inaba switching. In Vibrio cholerae and Cholera: Molecular to Global Perspectives, eds. I. K. Wachsmuth, P. A. Blake, and O. Olsvik, pp. 77–94. ASM Press, Washington, DCCrossRef
Matsumoto, C., Okuda, J., Ishibashi, M., Iwanaga, M., Garg, P., Ramamurthy, T., Wong, H. C., Depaola, A., Kim, Y. B., Albert, M. J., and Nishibuchi, M. (2000). Pandemic spread of an O3:K6 clone of Vibrio parahaemolyticus and emergence of related strains evidenced by arbitrarily primed PCR and toxRS sequence analyses. Journal of Clinical Microbiology, 38, 578–85Google ScholarPubMed
Mazel, D., Dychinco, B., Webb, V. A., and Davies, J. (1998). A distinctive class of integron in the Vibrio cholerae genome. Science, 280, 605–8CrossRefGoogle ScholarPubMed
Mekalanos, J. J. (1983). Duplication and amplification of toxin genes in Vibrio cholerae. Cell, 35, 253–63CrossRefGoogle ScholarPubMed
Mekalanos, J. J., Rubin, E. J., and Waldor, M. K. (1997). Cholera: molecular basis for emergence and pathogenesis. FEMS Immunology and Medical Microbiology, 18, 241–8CrossRefGoogle ScholarPubMed
Merrell, D. S., Butler, S. M., Qadri, F., Dolganov, N. A., Alam, A., Cohen, M. B., Calderwood, S. B., Schoolnik, G. K., and Camilli, A. (2002). Host-induced epidemic spread of the cholera bacterium. Nature, 417, 642–5CrossRefGoogle ScholarPubMed
Miao, E. A., and Miller, S. I. (1999). Bacteriophages in the evolution of pathogen-host interactions. Proceedings of the National Academy of Sciences USA, 96, 9452–4CrossRefGoogle ScholarPubMed
Morris, J. G. (1990). Non-O group 1 Vibrio cholerae: a look at the epidemiology of an occasional pathogen. Epidemiological Reviews 12, 179–91CrossRefGoogle Scholar
Morris, J. G. Jr., and Black, R. E. (1985). Cholera and other vibrios in the United States. New England Journal of Medicine, 312, 343–50CrossRefGoogle ScholarPubMed
Mukhopadhyay, A. K., Chakraborty, S., Takeda, Y., Nair, G. B., and Berg, D. E. (2001). Characterization of VPI pathogenicity island and CTXΦ prophage in environmental strains of Vibrio cholerae. Journal of Bacteriology, 183, 4737–46CrossRefGoogle ScholarPubMed
Nair, G. B., Garg, S., Mukhopadhyay, A. K., Shimada, T., and Takeda, Y. (1994 a). Laboratory diagnosis of Vibrio cholerae O139 Bengal, the new pandemic strain of cholera. LabMedica International, , 8–11Google Scholar
Nair, G. B., Ramamurthy, T., Bhattacharya, S. K., Mukhopadhyay, A. K., Garg, S., Bhattacharya, M. K., Takeda, T., Shimada, T., Takeda, Y., and Deb, B. C. (1994 (b). Spread of Vibrio cholerae O139 Bengal in India. Journal of Infectious Diseases, 169, 1029–34CrossRefGoogle ScholarPubMed
Nair, G. B., Faruque, S. M., Bhuiyan, N. A., Kamruzzaman, M., Siddique, A. K., and Sack, D. A. (2002). New variants of Vibrio cholerae O1 biotype El Tor with attributes of the classical biotype from hospitalized patients with acute diarrhea in Bangladesh. Journal of Clinical Microbiology, 40, 3296–9CrossRefGoogle ScholarPubMed
Ochman, H., Lawrence, J. G., and Groisman, E. A. (2000). Lateral gene transfer and the nature of bacterial innovation. Nature, 405, 299–304CrossRefGoogle ScholarPubMed
Okuda, J., Ishibashi, M., Hayakawa, E., Nishino, T., Takeda, Y., Mukhopadhyay, A. K., Garg, S., Bhattacharya, S. K., Nair, G. B., and Nishibuchi, M. (1997). Emergence of a unique O3:K6 clone of Vibrio parahaemolyticus in Calcutta, India, and isolation of strains from the same clonal group from southeast Asian travellers arriving in Japan. Journal of Clinical Microbiology, 35, 3150–5Google Scholar
Parsot, C., Taxman, E., and Mekalanos, J. J. (1991). ToxR regulates the production of lipoproteins and the expression of serum resistance in Vibrio cholerae. Proceedings of the National Academy of Sciences USA, 88, 1641–5CrossRefGoogle ScholarPubMed
Pearson, G. D. N., Woods, A., Chiang, S. L., and Mekalanos, J. J. (1993). CTX genetic element encodes a site-specific recombination system and an intestinal colonization factor. Proceedings of the National Academy of Sciences USA, 90, 3750–4CrossRefGoogle ScholarPubMed
Peek, J. A., and Taylor, R. K. (1992). Characterization of a periplasmic thiol: disulfide interchange protein required for the functional maturation of secreted virulence factors of Vibrio cholerae. Proceedings of the National Academy of Sciences USA, 89, 6210–4CrossRefGoogle ScholarPubMed
Peterson, K. M., and Mekalanos, J. J. (1988). Characterization of the Vibrio cholerae ToxR regulon: identification of novel genes involved in intestinal colonization. Infection and Immunity, 56, 2822–9Google ScholarPubMed
Popovic, T., Bopp, C., Olsvik, O., and Wachsmuth, K. (1993). Epidemiologic application of a standardized ribotype scheme for Vibrio cholerae O1. Journal of Clinical Microbiology, 31, 2474–82Google ScholarPubMed
Reidl, J., and Klose, K. E. (2002). Vibrio cholerae and cholera: out of the water and into the host. FEMS Microbiology Reviews, 26, 125–39CrossRefGoogle Scholar
Rivera, I. N. G., Chun, J., Huq, A., Sack, R. B., and Colwell, R. R. (2001). Genotypes associated with virulence in environmental isolates of Vibrio cholerae. Applied and Environmental Microbiology, 67, 2421–9CrossRefGoogle ScholarPubMed
Rowe-Magnus, D. A., and Mazel, D. (1999). Resistance gene capture. Current Opinion in Microbiology, 2, 483–8CrossRefGoogle ScholarPubMed
Sakazaki, R., and Donovan, T. J. (1984). Serology and epidemiology of Vibrio cholerae and Vibrio mimicus. In Methods in Microbiology, Vol. 16, ed. Y. Bergan, pp. 271–89. Academic Press, LondonCrossRef
Salles, C. A., and Momen, H. (1991). Identification of Vibrio cholerae by enzyme electrophoresis. Transactions of the Royal Society of Tropical Medicine and Hygiene, 85, 544–7CrossRefGoogle ScholarPubMed
Sengupta, D. K., Sengupta, T. K., and Ghose, A. C. (1992). Major outer membrane proteins of Vibrio cholerae and their role in induction of protective immunity through inhibition of intestinal colonization. Infection and Immunity, 60, 4848–55Google ScholarPubMed
Skorupski, K., and Taylor, R. K. (1997). Control of the ToxR virulence regulon in Vibrio cholerae by environmental stimuli. Molecular Microbiology, 25, 1003–9CrossRefGoogle ScholarPubMed
Sozhamannan, S., Deng, Y. K., Li, M., Sulakvelidze, A., Kaper, J. B., Johnson, J. A., Nair, G. B., and Morris, J. G. Jr. (1999). Cloning and sequencing of the genes downstream of the wbf gene cluster of Vibrio cholerae serogroup O139 and analysis of the junction genes in other serogroups. Infection and Immunity, 67, 5033–40Google ScholarPubMed
Stroeher, U. H., Jedani, K. E., Dredge, B. K., Morona, R., Brown, M. H., Karageorgos, L. E., Albert, J. M., and Manning, P. A. (1995). Genetic rearrangements in the rfb regions of Vibrio cholerae O1 and O139. Proceedings of the National Academy of Sciences USA, 92, 10,374–8CrossRefGoogle ScholarPubMed
Taylor, R. K., Miller, V. L., Furlong, D. B., and Mekalanos, J. J. (1987). Use of phoA gene fusions to identify a pilus colonization factor coordinately regulated with cholera toxin. Proceedings of the National Academy of Sciences USA, 84, 2833–7CrossRefGoogle ScholarPubMed
Taylor, R., Shaw, C., Peterson, K., Spears, P., and Mekalanos, J. (1988). Safe, live Vibrio cholerae vaccines? Vaccine, 6, 151–4CrossRefGoogle ScholarPubMed
Trucksis, M., Michalski, J., Deng, Y. K., and Kaper, J. B. (1998). The Vibrio cholerae genome contains two unique circular chromosomes. Proceedings of the National Academy of Sciences USA, 95, 14,464–9CrossRefGoogle ScholarPubMed
Voss, E., and Attridge, S. R. (1993). In vitro production of toxin-coregulated pili by Vibrio cholerae El Tor. Microbial Pathogenesis, 15, 255–68CrossRefGoogle ScholarPubMed
Wachsmuth, I. K., Evins, G. M., Fields, P. I., Olsvic, O., Popovic, T., Bopp, C. A., Wells, J. G., Cerrillo, C., and Blake, P. A. (1993). The molecular epidemiology of cholera in Latin America. Journal of Infectious Diseases, 167, 621–6CrossRefGoogle ScholarPubMed
Wachsmuth, I. K., Olsvik, O., Evins, G. M., and Popovic, T. (1994). Molecular epidemiology of cholera, In Vibrio cholerae and Cholera: Molecular to Global Perspectives, eds. K. Wachsmuth, P. A. Blake, and O. Olsvik, pp. 357–70. American Society of Microbiology, Washington, DC
Waldor, M. K., and Mekalanos, J. J. (1994). Vibrio cholerae O139 specific gene sequences. Lancet, 343, 1366CrossRefGoogle ScholarPubMed
Waldor, M. K., and Mekalanos, J. J. (1996). Lysogenic conversion by a filamentous bacteriophage encoding cholera toxin. Science, 272, 1910–14CrossRefGoogle ScholarPubMed
Waldor, M. K., and RayChaudhuri, D. (2000). Bacterial genomics: Treasure trove for cholera research. Nature, 406(6795), 469–70CrossRefGoogle Scholar
Waldor, M. K., Rubin, E. J., Pearson, G. D. N., Kimsey, H., and Mekalanos, J. J. (1997). Regulation, replication, and integration functions of the Vibrio cholerae CTXΦ are encoded by regions RS2. Molecular Microbiology, 24, 917–26CrossRefGoogle ScholarPubMed
Watnick, P., and Kolter, R. (2000). Biofilm, city of microbes. Journal of Bacteriology, 182, 2675–9CrossRefGoogle ScholarPubMed
Watnick, P. I., Lauriano, C. M., Klose, K. E., Croal, L., and Kolter, R. (2001). The absence of a flagellum leads to altered colony morphology, biofilm development and virulence in Vibrio cholerae O139. Molecular Microbiology, 39, 223–35CrossRefGoogle ScholarPubMed
Xu, H. S., Roberts, N. C., Singleton, F. L., Attwell, R. W., Grimes, D. J., and Colwell, R. R. (1982). Survival and viability of nonculturable Escherichia coli and Vibrio cholerae in the estuarine and marine environment. Microbial Ecology, 8, 313–23CrossRefGoogle Scholar
Yamai, S., Okitsu, T., Shimada, T., and Katsube, Y. (1997). Distribution of serogroups of Vibrio cholerae non-O1 non-O139 with specific reference to their ability to produce cholera toxin, and addition of novel serogroups. Kansenshogaku Zasshi, 71, 1037–45. (In Japanese.)CrossRef
Yamasaki, S., Garg, S., Nair, G. B., and Takeda, Y. (1999 a). Distribution of Vibrio cholerae O1 antigen biosynthesis genes among O139 and other non-O1 serogroups of Vibrio cholerae. FEMS Microbiology Letters, 179, 115–21CrossRefGoogle ScholarPubMed
Yamasaki, S., Shimizu, T., Hoshino, K., Ho, S. T., Shimada, T., Nair, G. B., and Takeda, Y. (1999 b). The genes responsible for O-antigen synthesis of Vibrio cholerae O139 are closely related to those of Vibrio cholerae O22. Gene, 237, 321–32CrossRefGoogle 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
×