Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-n9wrp Total loading time: 0 Render date: 2024-07-18T11:47:09.623Z Has data issue: false hasContentIssue false

1 - Mechanisms of homologous recombination in bacteria

Published online by Cambridge University Press:  06 August 2009

Marie-Agnès Petit
Affiliation:
Faculté de médecine Necker-Enfants Malades, France
Peter Mullany
Affiliation:
University College London
Get access

Summary

Homologous recombination promotes the pairing between identical — or nearly identical – DNA sequences and the subsequent exchange of genetic material between them. It is an important and widely conserved function in living organisms, from bacteria to humans, that serves to repair double-stranded breaks or single-stranded gaps in the DNA, arising as a consequence of ionizing radiations, ultraviolet (UV) light, or chemical treatments creating replication-blocking adducts (Kuzminov, 1999). More recently, homologous recombination functions were also found in bacteria to rescue replication forks that have stalled for various reasons, such as a missing factor (e.g., the helicase), or a particular difficulty upstream of the fork, such as supercoiling or intense traffic of proteins (Michel et al., 2001).

Besides its molecular role, homologous recombination has played a major role in genome dynamics, by changing gene copy numbers through deletions, duplications, and amplifications: Intrachromosomal recombination between ribosomal operons or between mobile elements scattered into the genome leads to deletion or tandem duplications of large regions within the genome, up to several hundred kilobases (Roth et al., 1996). The duplications are unstable. Mostly they recombine back to the parental organization, and, therefore, remain undetected, except when appropriate selection, by gene dosage mostly, is exerted (Petes and Hill, 1988). In contrast, such duplications are ideal substrate for the diversification of genes: One gene is kept intact whereas the other is mutagenized, which leads to the birth of gene families.

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

Abdulkarim, F., and Hughes, D. (1996). Homologous recombination between the tuf genes of Salmonella typhimurium. J Mol Biol, 260, 506–522CrossRefGoogle ScholarPubMed
Alonso, J. C., Luder, G., and Tailor, R. H. (1991). Characterization of Bacillus subtilis recombinational pathways. J Bacteriol, 173, 3977–3980CrossRefGoogle ScholarPubMed
Amundsen, S. K., and Smith, G. R. (2003). Interchangeable parts of the Escherichia coli recombination machinery. Cell, 112, 741–744CrossRefGoogle ScholarPubMed
Anderson, D. G., and Kowalczykowski, S. C. (1997). The translocating RecBCD enzyme stimulates recombination by directing RecA protein onto ssDNA in a chi-regulated manner. Cell, 90, 77–86CrossRefGoogle Scholar
Anderson, R. P., and Roth, J. R. (1977). Tandem genetic duplications in phage and bacteria. Annu Rev Microbiol, 31, 473–505CrossRefGoogle ScholarPubMed
Ayora, S., Missich, R., Mesa, P., Lurz, R., Yang, S., Egelman, E. H., and Alonso, J. C. (2002). Homologous-pairing activity of the Bacillus subtilis bacteriophage SPP1 replication protein G35P. J Biol Chem, 277, 35969–35979CrossRefGoogle ScholarPubMed
Ayora, S., Carrasco, B., Doncel, E., Lurz, R., and Alonso, J. C. (2004). Bacillus subtilis RecU protein cleaves Holliday junctions and anneals single-stranded DNA. Proc Natl Acad Sci U S A, 101, 452–457CrossRefGoogle ScholarPubMed
Baitin, D. M., Zaitsev, E. N., and Lanzov, V. A. (2003). Hyper-recombinogenic RecA protein from Pseudomonas aeruginosa with enhanced activity of its primary DNA binding site. J Mol Biol, 328, 1–7CrossRefGoogle ScholarPubMed
Bierne, H., Seigneur, M., Ehrlich, S. D., and Michel, B. (1997). uvrD mutations enhance tandem repeat deletion in the Escherichia coli chromosome via SOS induction of the RecF recombination pathway. Mol Microbiol, 26, 557–567CrossRefGoogle ScholarPubMed
Bleuit, J. S., Xu, H., Ma, Y., Wang, T., Liu, J., and Morrical, S. W. (2001). Mediator proteins orchestrate enzyme-ssDNA assembly during T4 recombination-dependent DNA replication and repair. Proc Natl Acad Sci U S A, 98, 8298–8305CrossRefGoogle ScholarPubMed
Bolt, E. L., and Lloyd, R. G. (2002). Substrate specificity of RusA resolvase reveals the DNA structures targeted by RuvAB and RecG in vivo. Mol Cell, 10, 187–198CrossRefGoogle ScholarPubMed
Bruand, C., Farache, M., McGovern, S., Ehrlich, S. D., and Polard, P. (2001). DnaB, DnaD and DnaI proteins are components of the Bacillus subtilis replication restart primosome. Mol Microbiol, 42, 245–255CrossRefGoogle ScholarPubMed
Buchmeier, N. A., Lipps, C. J., So, M. Y., and Heffron, F. (1993). Recombination-deficient mutants of Salmonella typhimurium are avirulent and sensitive to the oxidative burst of macrophages. Mol Microbiol, 7, 933–936CrossRefGoogle ScholarPubMed
Canceill, D., and Ehrlich, S. D. (1996). Copy-choice recombination mediated by DNA polymerase III holoenzyme from Escherichia coli. Proc Natl Acad Sci U S A, 93, 6647–6652CrossRefGoogle ScholarPubMed
Capaldo, F. N., Ramsey, G., and Barbour, S. D. (1974). Analysis of the growth of recombination-deficient strains of Escherichia coli K-12. J Bacteriol, 118, 242–249Google ScholarPubMed
Chedin, F., Ehrlich, S. D., and Kowalczykowski, S. C. (2000). The Bacillus subtilis AddAB helicase/nuclease is regulated by its cognate Chi sequence in vitro. J Mol Biol, 298, 7–20CrossRefGoogle ScholarPubMed
Chedin, F., and Kowalczykowski, S. C. (2002). A novel family of regulated helicases/nucleases from gram-positive bacteria: Insights into the initiation of DNA recombination. Mol Microbiol, 43, 823–834CrossRefGoogle ScholarPubMed
Courcelle, J., Khodursky, A., Peter, B., Brown, P. O., and Hanawalt, P. C. (2001). Comparative gene expression profiles following UV exposure in wild-type and SOS-deficient Escherichia coli. Genetics, 158, 41–64Google ScholarPubMed
Cox, M. M. (2001). Recombinational DNA repair of damaged replication forks in Escherichia coli: Questions. Annu Rev Genet, 35, 53–82CrossRefGoogle ScholarPubMed
Cromie, G. A., Connelly, J. C., and Leach, D. R. (2001). Recombination at double-strand breaks and DNA ends: Conserved mechanisms from phage to humans. Mol Cell, 8, 1163–1174CrossRefGoogle Scholar
Cromie, G. A., and Leach, D. R. (2000). Control of crossing over. Mol Cell, 6, 815–826CrossRefGoogle ScholarPubMed
d'Alencon, E., Petranovic, M., Michel, B., Noirot, P., Aucouturier, A., Uzest, M., and Ehrlich, S. D. (1994). Copy-choice illegitimate DNA recombination revisited. Embo J, 13, 2725–2734Google ScholarPubMed
Dale, C., Wang, B., Moran, N., and Ochman, H. (2003). Loss of DNA recombinational repair enzymes in the initial stages of genome degeneration. Mol Biol Evol, 20, 1188–1194CrossRefGoogle ScholarPubMed
Datsenko, K. A., and Wanner, B. L. (2000). One-step inactivation of chromosomal genes in Escherichia coli K-12 using PCR products. Proc Natl Acad Sci U S A, 97, 6640–6645CrossRefGoogle ScholarPubMed
Vos, W. M., Vries, S. C., and Venema, G. (1983). Cloning and expression of the Escherichia coli recA gene in Bacillus subtilis. Gene, 25, 301–308CrossRefGoogle ScholarPubMed
Denamur, E., Bonacorsi, S., Giraud, A., Duriez, P., Hilali, F., Amorin, C., Bingen, E.. (2002). High frequency of mutator strains among human uropathogenic Escherichia coli isolates. J Bacteriol, 184, 605–609CrossRefGoogle ScholarPubMed
Dillingham, M. S., Spies, M., and Kowalczykowski, S. C. (2003). RecBCD enzyme is a bipolar DNA helicase. Nature, 423, 893–897CrossRefGoogle ScholarPubMed
Dubnau, D., and Lovett, C. M., Jr. (2002). “Transformation and recombination.” In Sonenshein, A. L., Hoch, J. A., and Losick, R. (eds.), Bacillus subtilis and its closest relatives. From genes to cells. Washington, DC: ASM Press, pp. 453–471Google Scholar
Earl, A. M., Mohundro, M. M., Mian, I. S., and Battista, J. R. (2002). The IrrE protein of Deinococcus radiodurans R1 is a novel regulator of recA expression. J Bacteriol, 184, 6216–6224CrossRefGoogle ScholarPubMed
el Karoui, M., Biaudet, V., Schbath, S., and Gruss, A. (1999). Characteristics of Chi distribution on different bacterial genomes. Res Microbiol, 150, 579–587CrossRefGoogle ScholarPubMed
el Karoui, M., Ehrlich, D., and Gruss, A. (1998). Identification of the lactococcal exonuclease/recombinase and its modulation by the putative Chi sequence. Proc Natl Acad Sci U S A, 95, 626–631CrossRefGoogle ScholarPubMed
Fernandez, S., Ayora, S., and Alonso, J. C. (2000). Bacillus subtilis homologous recombination: genes and products. Res Microbiol, 151, 481–486CrossRefGoogle ScholarPubMed
Fernandez De Henestrosa, A. R., Ogi, T., Aoyagi, S., Chafin, D., Hayes, J. J., Ohmori, H., and Woodgate, R. (2000). Identification of additional genes belonging to the LexA regulon in Escherichia coli. Mol Microbiol, 35, 1560–1572CrossRefGoogle ScholarPubMed
Forterre, (1999). Displacement of cellular proteins by functional analogues from plasmids or viruses could explain puzzling phylogenies of many DNA informational proteins. Mol Microbiol 33, 457–465CrossRefGoogle ScholarPubMed
Hall, S. D., and Kolodner, R. D. (1994). Homologous pairing and strand exchange promoted by the Escherichia coli RecT protein. Proc Natl Acad Sci U S A, 91, 3205–3209CrossRefGoogle ScholarPubMed
Halpern, D., Gruss, A., Claverys, J. P., and El-Karoui, M. (2004). rexAB mutants in Streptococcus pneumoniae. Microbiology, 150, 2409–2414CrossRefGoogle ScholarPubMed
Hedayati, M. A., Steffen, S. E., and Bryant, F. R. (2002). Effect of the Streptococcus pneumoniae MmsA protein on the RecA protein-promoted three-strand exchange reaction. Implications for the mechanism of transformational recombination. J Biol Chem, 277, 24863–24869CrossRefGoogle ScholarPubMed
Hsieh, P., Camerini-Otero, C. S., and Camerini-Otero, R. D. (1992). The synapsis event in the homologous pairing of DNAs: RecA recognizes and pairs less than one helical repeat of DNA. Proc Natl Acad Sci U S A, 89, 6492–6496CrossRefGoogle ScholarPubMed
Humbert, O., Prudhomme, M., Hakenbeck, R., Dowson, C. G., and Claverys, J. P. (1995). Homeologous recombination and mismatch repair during transformation in Streptococcus pneumoniae: Saturation of the Hex mismatch repair system. Proc Natl Acad Sci U S A, 92, 9052–9056CrossRefGoogle ScholarPubMed
Ivancic-Bace, I., Peharec, P., Moslavac, S., Skrobot, N., Salaj-Smic, E., and Brcic-Kostic, K. (2003). RecFOR function is required for DNA repair and recombination in a RecA loading-deficient recB mutant of Escherichia coli. Genetics, 163, 485–494Google Scholar
Kim, J. I., and Cox, M. M. (2002). The RecA proteins of Deinococcus radiodurans and Escherichia coli promote DNA strand exchange via inverse pathways. Proc Natl Acad Sci U S A, 99, 7917–7921CrossRefGoogle ScholarPubMed
Kodama, K., Kobayashi, T., Niki, H., Hiraga, S., Oshima, T., Mori, H., and Horiuchi, T. (2002). Amplification of hot DNA segments in Escherichia coli. Mol Microbiol, 45, 1575–1588CrossRefGoogle ScholarPubMed
Kogoma, T., Cadwell, G. W., Barnard, K. G., and Asai, T. (1996). The DNA replication priming protein, PriA, is required for homologous recombination and double-strand break repair. J Bacteriol, 178, 1258–1264CrossRefGoogle ScholarPubMed
Kooistra, J., Haijema, B. J., and Venema, G. (1993). The Bacillus subtilis addAB genes are fully functional in Escherichia coli. Mol Microbiol, 7, 915–923CrossRefGoogle ScholarPubMed
Kowalczykowski, S. C., Dixon, D. A., Eggleston, A. K., Lauder, S. D., and Rehrauer, W. M. (1994). Biochemistry of homologous recombination in Escherichia coli. Microbiol Rev, 58, 401–465Google ScholarPubMed
Kuzminov, A. (1999). Recombinational repair of DNA damage in Escherichia coli and bacteriophage lambda. Microbiol Mol Biol Rev, 63, 751–813Google ScholarPubMed
Lloyd, R. G. (1991). Conjugational recombination in resolvase-deficient ruvC mutants of Escherichia coli K-12 depends on recG. J Bacteriol, 173, 5414–5418CrossRefGoogle ScholarPubMed
Lloyd, R. G., and Low, K. B. (1996). “Homologous recombination.” In Neidhart, F. C. (ed.), Escherichia coli and Salmonella, Vol. 2. Washington, DC: ASM Press, pp. 2236–2255Google Scholar
Lloyd, R. G., and Sharples, G. J. (1993). Dissociation of synthetic Holliday junctions by E. coli RecG protein. EMBO J, 12, 17–22Google ScholarPubMed
Loughlin, M. F., Barnard, F. M., Jenkins, D., Sharples, G. J., and Jenks, P. J. (2003). Helicobacter pylori mutants defective in RuvC Holliday junction resolvase display reduced macrophage survival and spontaneous clearance from the murine gastric mucosa. Infect Immun, 71, 2022–2031CrossRefGoogle ScholarPubMed
Lovett, C. M. Jr., Love, P. E., and Yasbin, R. E. (1989). Competence-specific induction of the Bacillus subtilis RecA protein analog: Evidence for dual regulation of a recombination protein. J Bacteriol, 171, 2318–2322CrossRefGoogle ScholarPubMed
Majewski, J., and Cohan, F. M. (1998). The effect of mismatch repair and heteroduplex formation on sexual isolation in Bacillus. Genetics, 148, 13–18Google ScholarPubMed
Marians, K. J. (1999). PriA: At the crossroads of DNA replication and recombination. Prog Nucleic Acid Res Mol Biol, 63, 39–67CrossRefGoogle Scholar
Marians, K. J. (2000). PriA-directed replication fork restart in Escherichia coli. Trends Biochem Sci, 25, 185–189CrossRefGoogle ScholarPubMed
Marsin, S., McGovern, S., Ehrlich, S. D., Bruand, C., and Polard, P. (2001). Early steps of Bacillus subtilis primosome assembly. J Biol Chem, 276, 45818–45825CrossRefGoogle ScholarPubMed
Martin, B., Sharples, G. J., Humbert, O., Lloyd, R. G., and Claverys, J. P. (1996). The mmsA locus of Streptococcus pneumoniae encodes a RecG-like protein involved in DNA repair and in three-strand recombination. Mol Microbiol, 19, 1035–1045CrossRefGoogle ScholarPubMed
McGlynn, P., and Lloyd, R. G. (2000). Modulation of RNA polymerase by (p)ppGpp reveals a RecG-dependent mechanism for replication fork progression. Cell, 101, 35–45CrossRefGoogle ScholarPubMed
Michel, B., Flores, M. J., Viguera, E., Grompone, G., Seigneur, M., and Bidnenko, V. (2001). Rescue of arrested replication forks by homologous recombination. Proc Natl Acad Sci U S A, 98, 8181–8188CrossRefGoogle ScholarPubMed
Michel, B., Recchia, G. D., Penel-Colin, M., Ehrlich, S. D., and Sherratt, D. J. (2000). Resolution of Holliday junctions by RuvABC prevents dimer formation in rep mutants and UV-irradiated cells. Mol Microbiol, 37, 180–191CrossRefGoogle ScholarPubMed
Miller, E. S., Kutter, E., Mosig, G., Arisaka, F., Kunisawa, T., and Ruger, W. (2003). Bacteriophage T4 genome. Microbiol Mol Biol Rev, 67, 86–156CrossRefGoogle ScholarPubMed
Morimatsu, K., and Kowalczykowski, S. C. (2003). RecFOR proteins load RecA protein onto gapped DNA to accelerate DNA strand exchange: A universal step of recombinational repair. Mol Cell, 11, 1337–1347CrossRefGoogle ScholarPubMed
Muyrers, J. P., Zhang, Y., Buchholz, F., and Stewart, A. F. (2000). RecE/RecT and Redalpha/Redbeta initiate double-stranded break repair by specifically interacting with their respective partners. Genes Dev, 14, 1971–1982Google ScholarPubMed
Myers, R. S., and Stahl, F. W. (1994). Chi and the RecBC D enzyme of Escherichia coli. Annu Rev Genet, 28, 49–70CrossRefGoogle ScholarPubMed
Narumi, I., Satoh, K., Kikuchi, M., Funayama, T., Yanagisawa, T., Kobayashi, Y., Watanabe, H., and Yamamoto, K. (2001). The LexA protein from Deinococcus radiodurans is not involved in RecA induction following gamma irradiation. J Bacteriol, 183, 6951–6956CrossRefGoogle Scholar
Noirot, P., Gupta, R. C., Radding, C. M., and Kolodner, R. D. (2003). Hallmarks of homology recognition by RecA-like recombinases are exhibited by the unrelated Escherichia coli RecT protein. Embo J, 22, 324–334CrossRefGoogle ScholarPubMed
Noirot, P., and Kolodner, R. D. (1998). DNA strand invasion promoted by Escherichia coli RecT protein. J Biol Chem, 273, 12274–12280CrossRefGoogle 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
Ogura, M., Yamaguchi, H., Kobayashi, K., Ogasawara, N., Fujita, Y., and Tanaka, T. (2002). Whole-genome analysis of genes regulated by the Bacillus subtilis competence transcription factor ComK. J Bacteriol, 184, 2344–2351CrossRefGoogle ScholarPubMed
Petes, T. D., and Hill, C. W. (1988). Recombination between repeated genes in microorganisms. Annu Rev Genet, 22, 147–168CrossRefGoogle ScholarPubMed
Petit, M. A., Dimpfl, J., Radman, M., and Echols, H. (1991). Control of large chromosomal duplications in Escherichia coli by the mismatch repair system. Genetics, 129, 327–332Google ScholarPubMed
Polard, P., Marsin, S., McGovern, S., Velten, M., Wigley, D. B., Ehrlich, S. D., and Bruand, C. (2002). Restart of DNA replication in gram-positive bacteria: Functional characterisation of the Bacillus subtilis PriA initiator. Nucleic Acids Res, 30, 1593–1605CrossRefGoogle ScholarPubMed
Quiberoni, A., Biswas, I., El Karoui, M., Rezaiki, L., Tailliez, P., and Gruss, A. (2001). In vivo evidence for two active nuclease motifs in the double-strand break repair enzyme RexAB of Lactococcus lactis. J Bacteriol, 183, 4071–4078CrossRefGoogle ScholarPubMed
Rayssiguier, C., Thaler, D. S., and Radman, M. (1989). The barrier to recombination between Escherichia coli and Salmonella typhimurium is disrupted in mismatch-repair mutants. Nature, 342, 396–401CrossRefGoogle ScholarPubMed
Rinken, R., and Wackernagel, W. (1992). Inhibition of the recBCD-dependent activation of Chi recombinational hot spots in SOS-induced cells of Escherichia coli. J Bacteriol, 174, 1172–1178CrossRefGoogle ScholarPubMed
Robu, M. E., Inman, R. B., and Cox, M. M. (2001). RecA protein promotes the regression of stalled replication forks in vitro. Proc Natl Acad Sci U S A, 98, 8211–8218CrossRefGoogle ScholarPubMed
Roca, A. I., and Cox, M. M. (1997). RecA protein: Structure, function, and role in recombinational DNA repair. Prog Nucleic Acid Res Mol Biol, 56, 129–223CrossRefGoogle ScholarPubMed
Roth, J. R., Benson, N., Galitski, T., Haak, K., Lawrence, J., and Miesel, L. (1996). “Rearrangements of the bacterial chromosome: Formation and applications.” In Neidhart, F. C. (ed.), Escherichia coli and Salmonella, Vol. 2. Washington, DC: ASM Press, pp. 2256–2276Google Scholar
Saveson, C. J., and Lovett, S. T. (1997). Enhanced deletion formation by aberrant DNA replication in Escherichia coli. Genetics, 146, 457–470Google ScholarPubMed
Sawitzke, J. A., and Stahl, F. W. (1997). Roles for lambda Orf and Escherichia coli RecO, RecR and RecF in lambda recombination. Genetics, 147, 357–369Google ScholarPubMed
Schapiro, J. M., Libby, S. J., and Fang, F. C. (2003). Inhibition of bacterial DNA replication by zinc mobilization during nitrosative stress. Proc Natl Acad Sci U S A, 100, 8496–8501CrossRefGoogle ScholarPubMed
Schmitt, W., Odenbreit, S., Heuermann, D., and Haas, R. (1995). Cloning of the Helicobacter pylori recA gene and functional characterization of its product. Mol Gen Genet, 248, 563–572CrossRefGoogle ScholarPubMed
Seigneur, M., Bidnenko, V., Ehrlich, S. D., and Michel, B. (1998). RuvAB acts at arrested replication forks. Cell, 95, 419–430CrossRefGoogle ScholarPubMed
Seigneur, M., Ehrlich, S. D., and Michel, B. (2000). RuvABC-dependent double-strand breaks in dnaBts mutants require recA. Mol Microbiol, 38, 565–574CrossRefGoogle ScholarPubMed
Sharples, G. J. (2001). The X philes: Structure-specific endonucleases that resolve Holliday junctions. Mol Microbiol, 39, 823–834CrossRefGoogle ScholarPubMed
Sharples, G. J., Ingleston, S. M., and Lloyd, R. G. (1999). Holliday junction processing in bacteria: Insights from the evolutionary conservation of RuvABC, RecG, and RusA. J Bacteriol, 181, 5543–5550Google ScholarPubMed
Singleton, M. R., Dillingham, M. S., Gaudier, M., Kowalczykowski, S. C., and Wigley, D. B. (2004). Crystal structure of RecBCD enzyme reveals a machine for processing DNA breaks. Nature, 432, 187–193CrossRefGoogle ScholarPubMed
Takahashi, N. K., Kusano, K., Yokochi, T., Kitamura, Y., Yoshikura, H., and Kobayashi, I. (1993). Genetic analysis of double-strand break repair in Escherichia coli. J Bacteriol, 175, 5176–5185CrossRefGoogle ScholarPubMed
Tatusov, R. L., Koonin, E. V., and Lipman, D. J. (1997). A genomic perspective on protein families. Science, 278, 631–637CrossRefGoogle ScholarPubMed
Tatusov, R. L., Natale, D. A., Garkavtsev, I. V., Tatusova, T. A., Shankavaram, U. T., Rao, B. S., Kiryutin, B., Galperin, M. Y., Fedorova, N. D., and Koonin, E. V. (2001). The COG database: New developments in phylogenetic classification of proteins from complete genomes. Nucleic Acids Res, 29, 22–28CrossRefGoogle ScholarPubMed
Taylor, A. F., and Smith, G. R. (2003). RecBCD enzyme is a DNA helicase with fast and slow motors of opposite polarity. Nature, 423, 889–893CrossRefGoogle ScholarPubMed
Velten, M., McGovern, S., Marsin, S., Ehrlich, S. D., Noirot, P., and Polard, P. (2003). A two-protein strategy for the functional loading of a cellular replicative DNA helicase. Mol Cell, 11, 1009–1020CrossRefGoogle ScholarPubMed
Viguera, E., Canceill, D., and Ehrlich, S. D. (2001). Replication slippage involves DNA polymerase pausing and dissociation. EMBO J, 20, 2587–2595CrossRefGoogle ScholarPubMed
Voloshin, O. N., Ramirez, B. E., Bax, A., and Camerini-Otero, R. D. (2001). A model for the abrogation of the SOS response by an SOS protein: A negatively charged helix in DinI mimics DNA in its interaction with RecA. Genes Dev, 15, 415–427CrossRefGoogle ScholarPubMed
Vulic, M., Dionisio, F., Taddei, F., and Radman, M. (1997). Molecular keys to speciation: DNA polymorphism and the control of genetic exchange in enterobacteria. Proc Natl Acad Sci U S A, 94, 9763–9767CrossRefGoogle ScholarPubMed
Xu, L., and Marians, K. J. (2003). PriA mediates DNA replication pathway choice at recombination intermediates. Mol Cell, 11, 817–826CrossRefGoogle ScholarPubMed
Yu, D., Ellis, H. M., Lee, E. C., Jenkins, N. A., Copeland, N. G., and Court, D. L. (2000). An efficient recombination system for chromosome engineering in Escherichia coli. Proc Natl Acad Sci U S A, 97, 5978–5983CrossRefGoogle ScholarPubMed
Zhang, Y., Buchholz, F., Muyrers, J. P., and Stewart, A. F. (1998). A new logic for DNA engineering using recombination in Escherichia coli. Nat Genet, 20, 123–128CrossRefGoogle ScholarPubMed
Zulty, J. J., and Barcak, G. J. (1993). Structural organization, nucleotide sequence, and regulation of the Haemophilus influenzae rec-1+ gene. J Bacteriol, 175, 7269–7281CrossRefGoogle 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
×