全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

The Ecology of Bacterial Genes and the Survival of the New

DOI: 10.1155/2012/394026

Full-Text   Cite this paper   Add to My Lib

Abstract:

Much of the observed variation among closely related bacterial genomes is attributable to gains and losses of genes that are acquired horizontally as well as to gene duplications and larger amplifications. The genomic flexibility that results from these mechanisms certainly contributes to the ability of bacteria to survive and adapt in varying environmental challenges. However, the duplicability and transferability of individual genes imply that natural selection should operate, not only at the organismal level, but also at the level of the gene. Genes can be considered semiautonomous entities that possess specific functional niches and evolutionary dynamics. The evolution of bacterial genes should respond both to selective pressures that favor competition, mostly among orthologs or paralogs that may occupy the same functional niches, and cooperation, with the majority of other genes coexisting in a given genome. The relative importance of either type of selection is likely to vary among different types of genes, based on the functional niches they cover and on the tightness of their association with specific organismal lineages. The frequent availability of new functional niches caused by environmental changes and biotic evolution should enable the constant diversification of gene families and the survival of new lineages of genes. 1. Introduction Genomic science has brought about the possibility of disclosing the genetic underpinnings of entire organisms, and, for microbes, even of complex populations and communities. In doing so, it has unearthed a good number of surprising facts regarding the structure, organization, and variability of genomes, with strong evolutionary implications. In bacteria, the relatively small size of genomes has warranted the obtention of entire genomic sequences for a vast number of organisms, including numerous sets of sequences of closely related taxa. As genome sequences of closely related bacteria have accumulated, our view of bacterial genomes has radically changed. Genome comparisons have demonstrated that little 16S rRNA sequence divergence can be accompanied by large differences in total gene repertoire [1–7], and that even populations of a single 16S rRNA species can be made up of vast numbers of genomic varieties [8]. Much of the observed variation among closely related bacterial genomes has been attributed to gains and losses of genes that are acquired horizontally, often by way of mobile genetic elements (MGEs) as well as to a variety of genomic rearrangements that include gene duplications and large gene

References

[1]  N. T. Perna, G. Plunkett, V. Burland et al., “Genome sequence of enterohaemorrhagic Escherichia coli O157:H7,” Nature, vol. 409, no. 6819, pp. 529–533, 2001.
[2]  V. Daubin and H. Ochman, “Bacterial genomes as new gene homes: the genealogy of ORFans in E. coli,” Genome Research, vol. 14, no. 6, pp. 1036–1042, 2004.
[3]  H. Tettelin, V. Masignani, M. J. Cieslewicz et al., “Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: implications for the microbial ‘pan-genome’,” Proceedings of the National Academy of Sciences of the United States of America, vol. 102, no. 39, pp. 13950–13955, 2005.
[4]  M. L. Coleman, M. B. Sullivan, A. C. Martiny et al., “Genomic islands and the ecology and evolution of Prochlorococcus,” Science, vol. 311, no. 5768, pp. 1768–1770, 2006.
[5]  P. Normand, P. Lapierre, L. S. Tisa et al., “Genome characteristics of facultatively symbiotic Frankia sp. strains reflect host range and host plant biogeography,” Genome Research, vol. 17, no. 1, pp. 7–15, 2007.
[6]  M. W. J. Van Passel, P. R. Marri, and H. Ochman, “The emergence and fate of horizontally acquired genes in Escherichia coli,” PLoS Computational Biology, vol. 4, no. 4, Article ID e1000059, 2008.
[7]  A. Mira, A. B. Martín-Cuadrado, G. D'Auria, and F. Rodríguez-Valera, “The bacterial pan-genome: a new paradigm in microbiology,” International Microbiology, vol. 13, no. 2, pp. 45–57, 2010.
[8]  J. R. Thompson, S. Pacocha, C. Pharino et al., “Genotypic diversity within a natural coastal bacterioplankton population,” Science, vol. 307, no. 5713, pp. 1311–1313, 2005.
[9]  T. J. Treangen and E. P. C. Rocha, “Horizontal transfer, not duplication, drives the expansion of protein families in prokaryotes,” PLoS Genetics, vol. 7, no. 1, Article ID e1001284, 2011.
[10]  W. F. Doolittle and C. Sapienza, “Selfish genes, the phenotype paradigm and genome evolution,” Nature, vol. 284, no. 5757, pp. 601–603, 1980.
[11]  L. E. Orgel and F. H. C. Crick, “Selfish DNA: the ultimate parasite,” Nature, vol. 284, no. 5757, pp. 604–607, 1980.
[12]  M. P. Francino, “Gene survival in emergent genomes,” in Horizontal Gene Transfer in Microorganisms, M. P. Francino, Ed., pp. 1–22, Caister Academic Press, Norfolk, UK, 2012.
[13]  A. Mira, H. Ochman, and N. A. Moran, “Deletional bias and the evolution of bacterial genomes,” Trends in Genetics, vol. 17, no. 10, pp. 589–596, 2001.
[14]  L. Gómez-Valero, E. P. C. Rocha, A. Latorre, and F. J. Silva, “Reconstructing the ancestor of Mycobacterium leprae: the dynamics of gene loss and genome reduction,” Genome Research, vol. 17, no. 8, pp. 1178–1185, 2007.
[15]  C. H. Kuo and H. Ochman, “Deletional bias across the three domains of life,” Genome Biology and Evolution, vol. 1, pp. 145–152, 2009.
[16]  I. Comas, A. Moya, and F. González-Candelas, “Phylogenetic signal and functional categories in Proteobacteria genomes,” BMC Evolutionary Biology, vol. 7, no. 1, supplement, article S7, 2007.
[17]  A. Wellner, M. N. Lurie, and U. Gophna, “Complexity, connectivity, and duplicability as barriers to lateral gene transfer,” Genome Biology, vol. 8, no. 8, article R156, 2007.
[18]  J. G. Lawrence, “Selfish operons and speciation by gene transfer,” Trends in Microbiology, vol. 5, no. 9, pp. 355–359, 1997.
[19]  J. G. Lawrence and H. Hendrickson, “Lateral gene transfer: when will adolescence end?” Molecular Microbiology, vol. 50, no. 3, pp. 739–749, 2003.
[20]  I. Comas and F. González-Candelas, “The evolution of horizontally transferred genes: a model for prokaryotes,” in Horizontal Gene Transfer in Microorganisms, M. P. Francino, Ed., pp. 75–91, Caister Academic Press, Norfolk, UK, 2012.
[21]  K. S. Makarova, Y. I. Wolf, and E. V. Koonin, “Comprehensive comparative-genomic analysis of Type 2 toxin-antitoxin systems and related mobile stress response systems in prokaryotes,” Biology Direct, vol. 4, article 19, 2009.
[22]  T. Naito, K. Kusano, and I. Kobayashi, “Selfish behavior of restriction-modification systems,” Science, vol. 267, no. 5199, pp. 897–899, 1995.
[23]  I. Kobayashi, “Behavior of restriction-modification systems as selfish mobile elements and their impact on genome evolution,” Nucleic Acids Research, vol. 29, no. 18, pp. 3742–3756, 2001.
[24]  H. Mihara, T. Fujii, S. I. Kato, T. Kurihara, Y. Hata, and N. Esaki, “Structure of external aldimine of Escherichia coli CsdB, an IscS/NifS homolog: implications for its specificity toward selenocysteine,” Journal of Biochemistry, vol. 131, no. 5, pp. 679–685, 2002.
[25]  G. Michel, A. W. Roszak, V. Sauvé et al., “Structures of shikimate dehydrogenase AroE and its paralog YdiB: a common structural framework for different activities,” The Journal of Biological Chemistry, vol. 278, no. 21, pp. 19463–19472, 2003.
[26]  M. Blaise, H. D. Becker, J. Lapointe, C. Cambillau, R. Giegé, and D. Kern, “Glu-Q-tRNAAsp synthetase coded by the yadB gene, a new paralog of aminoacyl-tRNA synthetase that glutamylates tRNAAsp anticodon,” Biochimie, vol. 87, no. 9-10, pp. 847–861, 2005.
[27]  Y. Yin and J. F. Kirsch, “Identification of functional paralog shift mutations: conversion of Escherichia coli malate dehydrogenase to a lactate dehydrogenase,” Proceedings of the National Academy of Sciences of the United States of America, vol. 104, no. 44, pp. 17353–17357, 2007.
[28]  L. A. Nahum, S. Goswami, and M. H. Serres, “Protein families reflect the metabolic diversity of organisms and provide support for functional prediction,” Physiological Genomics, vol. 38, no. 3, pp. 250–260, 2009.
[29]  L. Mandrich and G. Manco, “Evolution in the amidohydrolase superfamily: substrate-assisted gain of function in the E183K mutant of a phosphotriesterase-like metal-carboxylesterase,” Biochemistry, vol. 48, no. 24, pp. 5602–5612, 2009.
[30]  A. Law and M. J. Boulanger, “Defining a structural and kinetic rationale for paralogous copies of phenylacetate-CoA ligases from the cystic fibrosis pathogen Burkholderia cenocepacia J2315,” The Journal of Biological Chemistry, vol. 286, no. 17, pp. 15577–15585, 2011.
[31]  D. McNally and M. A. Fares, “In silico identification of functional divergence between the multiple groEL gene paralogs in Chlamydiae,” BMC Evolutionary Biology, vol. 7, article 81, 2007.
[32]  G. Sanchez-Perez, A. Mira, G. Nyiro, L. Pa?i?, and F. Rodriguez-Valera, “Adapting to environmental changes using specialized paralogs,” Trends in Genetics, vol. 24, no. 4, pp. 154–158, 2008.
[33]  J. Li, Y. Wang, C. Y. Zhang et al., “Myxococcus xanthus viability depends on GroEL supplied by either of two genes, but the paralogs have different functions during heat shock, predation, and development,” Journal of Bacteriology, vol. 192, no. 7, pp. 1875–1881, 2010.
[34]  L. Van Valen, “Predation and species diversity,” Journal of Theoretical Biology, vol. 44, no. 1, pp. 19–21, 1974.
[35]  C. Gubry-Rangin, B. Hai, C. Quince et al., “Niche specialization of terrestrial archaeal ammonia oxidizers,” Proceedings of the National Academy of Sciences of the United States of America, vol. 108, no. 52, pp. 21206–21211, 2011.
[36]  N. Shapir, E. F. Mongodin, M. J. Sadowsky, S. C. Daugherty, K. E. Nelson, and L. P. Wackett, “Evolution of catabolic pathways: genomic insights into microbial s-triazine metabolism,” Journal of Bacteriology, vol. 189, no. 3, pp. 674–682, 2007.
[37]  M. Cai and L. Xun, “Organization and regulation of pentachlorophenol-degrading genes in Sphingobium chlorophenolicum ATCC 39723,” Journal of Bacteriology, vol. 184, no. 17, pp. 4672–4680, 2002.
[38]  M. H. Dai, J. B. Rogers, J. R. Warner, and S. D. Copley, “A previously unrecognized step in pentachlorophenol degradation in Sphingobium chlorophenolicum is catalyzed by tetrachlorobenzoquinone reductase (PcpD),” Journal of Bacteriology, vol. 185, no. 1, pp. 302–310, 2003.
[39]  G. J. Poelarends, M. Wilkens, M. J. Larkin, J. D. Van Elsas, and D. B. Janssen, “Degradation of 1,3-dichloropropene by Pseudomonas cichorii 170,” Applied and Environmental Microbiology, vol. 64, no. 8, pp. 2931–2936, 1998.
[40]  K. S. Ju and R. E. Parales, “Application of nitroarene dioxygenases in the design of novel strains that degrade chloronitrobenzenes,” Microbial Biotechnology, vol. 2, no. 2, pp. 241–252, 2009.
[41]  M. Kivisaar, “Evolution of catabolic pathways and their regulatory systems in synthetic nitroaromatic compounds degrading bacteria,” Molecular Microbiology, vol. 82, no. 2, pp. 265–268, 2011.
[42]  D. H. Pieper and M. Seeger, “Bacterial metabolism of polychlorinated biphenyls,” Journal of Molecular Microbiology and Biotechnology, vol. 15, no. 2-3, pp. 121–138, 2008.
[43]  S. D. Copley, “Evolution of efficient pathways for degradation of anthropogenic chemicals,” Nature Chemical Biology, vol. 5, no. 8, pp. 559–566, 2009.
[44]  A. A. Medeiros, “Evolution and dissemination of β-lactamases accelerated by generations of β-lactam antibiotics,” Clinical Infectious Diseases, vol. 24, no. 1, supplement, pp. S19–S45, 1997.
[45]  M. Barlow, J. Caywood, S. Lai, J. Finley, and C. Swanlund, “Horizontal gene transfer and recombination in the evolution of antibiotic resistance genes,” in Horizontal Gene Transfer in Microorganisms, M. P. Francino, Ed., pp. 165–176, Caister Academic Press, Norfolk, UK, 2012.
[46]  B. G. Hall, “The EBG system of E. coli: origin and evolution of a novel β-galactosidase for the metabolism of lactose,” Genetica, vol. 118, no. 2-3, pp. 143–156, 2003.
[47]  A. Aharoni, L. Gaidukov, O. Khersonsky, S. M. Gould, C. Roodveldt, and D. S. Tawfik, “The 'evolvability' of promiscuous protein functions,” Nature Genetics, vol. 37, no. 1, pp. 73–76, 2005.
[48]  V. López-Canut, M. Roca, J. Bertrán, V. Moliner, and I. Tu?ón, “Promiscuity in alkaline phosphatase superfamily. Unraveling evolution through molecular simulations,” Journal of the American Chemical Society, vol. 133, no. 31, pp. 12050–12062, 2011.
[49]  B. Frantz, T. Aldrich, and A. M. Chakrabarty, “Microbial degradation of synthetic recalcitrant compounds,” Biotechnology Advances, vol. 5, no. 1, pp. 85–99, 1987.
[50]  J. G. Lawrence, “Gene Transfer in Bacteria: speciation without species?” Theoretical Population Biology, vol. 61, no. 4, pp. 449–460, 2002.
[51]  W. P. Hanage, C. Fraser, and B. G. Spratt, “Fuzzy species among recombinogenic bacteria,” BMC Biology, vol. 3, article 6, 2005.
[52]  A. C. Retchless and J. G. Lawrence, “Temporal fragmentation of speciation in bacteria,” Science, vol. 317, no. 5841, pp. 1093–1096, 2007.
[53]  P. Shen and H. V. Huang, “Homologous recombination in Escherichia coli: dependence on substrate length and homology,” Genetics, vol. 112, no. 3, pp. 441–457, 1986.
[54]  J. Majewski, P. Zawadzki, P. Pickerill, F. M. Cohan, and C. G. Dowson, “Barriers to genetic exchange between bacterial species: Streptococcus pneumoniae transformation,” Journal of Bacteriology, vol. 182, no. 4, pp. 1016–1023, 2000.
[55]  P. Meier and W. Wackernagel, “Impact of mutS inactivation on foreign DNA acquisition by natural transformation in Pseudomonas stutzeri,” Journal of Bacteriology, vol. 187, no. 1, pp. 143–154, 2005.
[56]  A. B. Reams and E. L. Neidle, “Selection for gene clustering by tandem duplication,” Annual Review of Microbiology, vol. 58, pp. 119–142, 2004.
[57]  A. B. Reams, E. Kofoid, M. Savageau, and J. R. Roth, “Duplication frequency in a population of Salmonella enterica rapidly approaches steady state with or without recombination,” Genetics, vol. 184, no. 4, pp. 1077–1094, 2010.
[58]  H. Hendrickson, E. S. Slechta, U. Bergthorsson, D. I. Andersson, and J. R. Roth, “Amplification-mutagenesis: evidence that “directed” adaptive mutation and general hypermutability result from growth with a selected gene amplification,” Proceedings of the National Academy of Sciences of the United States of America, vol. 99, no. 4, pp. 2164–2169, 2002.
[59]  F. A. Kondrashov, I. B. Rogozin, Y. I. Wolf, and E. V. Koonin, “Selection in the evolution of gene duplications,” Genome Biology, vol. 3, no. 2, article RESEARCH0008, 2002.
[60]  S. D. Hooper and O. G. Berg, “On the nature of gene innovation: duplication patterns in microbial genomes,” Molecular Biology and Evolution, vol. 20, no. 6, pp. 945–954, 2003.
[61]  M. P. Francino, “An adaptive radiation model for the origin of new gene functions,” Nature Genetics, vol. 37, no. 6, pp. 573–577, 2005.
[62]  D. I. Andersson and D. Hughes, “Gene amplification and adaptive evolution in bacteria,” Annual Review of Genetics, vol. 43, pp. 167–195, 2009.
[63]  D. Romero and R. Palacios, “Gene amplification and genomic plasticity in prokaryotes,” Annual Review of Genetics, vol. 31, pp. 91–111, 1997.
[64]  M. M. Riehle, A. F. Bennett, and A. D. Long, “Genetic architecture of thermal adaptation in Escherichia coli,” Proceedings of the National Academy of Sciences of the United States of America, vol. 98, no. 2, pp. 525–530, 2001.
[65]  A. B. Reams and E. L. Neidle, “Genome plasticity in Acinetobacter: new degradative capabilities acquired by the spontaneous amplification of large chromosomal segments,” Molecular Microbiology, vol. 47, no. 5, pp. 1291–1304, 2003.
[66]  S. Zhong, A. Khodursky, D. E. Dykhuizen, and A. M. Dean, “Evolutionary genomics of ecological specialization,” Proceedings of the National Academy of Sciences of the United States of America, vol. 101, no. 32, pp. 11719–11724, 2004.
[67]  M. Gerstein, “A structural census of genomes: comparing bacterial, eukaryotic, and archaeal genomes in terms of protein structure,” Journal of Molecular Biology, vol. 274, no. 4, pp. 562–576, 1997.
[68]  M. A. Huynen and E. Van Nimwegen, “The frequency distribution of gene family sizes in complete genomes,” Molecular Biology and Evolution, vol. 15, no. 5, pp. 583–589, 1998.
[69]  J. Qian, N. M. Luscombe, and M. Gerstein, “Protein family and fold occurrence in genomes: power-law behaviour and evolutionary model,” Journal of Molecular Biology, vol. 313, no. 4, pp. 673–681, 2001.
[70]  P. M. Harrison and M. Gerstein, “Studying genomes through the aeons: protein families, pseudogenes and proteome evolution,” Journal of Molecular Biology, vol. 318, no. 5, pp. 1155–1174, 2002.
[71]  G. P. Karev, Y. I. Wolf, and E. V. Koonin, “Simple stochastic birth and death models of genome evolution: was there enough time for us to evolve?” Bioinformatics, vol. 19, no. 15, pp. 1889–1900, 2003.
[72]  O. Lespinet, Y. I. Wolf, E. V. Koonin, and L. Aravind, “The role of lineage-specific gene family expansion in the evolution of eukaryotes,” Genome Research, vol. 12, no. 7, pp. 1048–1059, 2002.
[73]  M. H. Serres, A. R. W. Kerr, T. J. McCormack, and M. Riley, “Evolution by leaps: gene duplication in bacteria,” Biology Direct, vol. 4, article 46, 2009.
[74]  E. Alm, K. Huang, and A. Arkin, “The evolution of two-component systems in bacteria reveals different strategies for niche adaptation,” PLoS Computational Biology, vol. 2, no. 11, article e143, pp. 1329–1342, 2006.
[75]  B. S. Goldman, W. C. Nierman, D. Kaiser et al., “Evolution of sensory complexity recorded in a myxobacterial genome,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 41, pp. 15200–15205, 2006.
[76]  W. Viratyosin, L. A. Campbell, C. C. Kuo, and D. D. Rockey, “Intrastrain and interstrain genetic variation within a paralogous gene family in Chlamydia pneumoniae,” BMC Microbiology, vol. 2, article 1, pp. 1–11, 2002.
[77]  J. G. Lawrence, “Gene transfer, speciation, and the evolution of bacterial genomes,” Current Opinion in Microbiology, vol. 2, no. 5, pp. 519–523, 1999.
[78]  J. P. Gogarten, W. F. Doolittle, and J. G. Lawrence, “Prokaryotic evolution in light of gene transfer,” Molecular Biology and Evolution, vol. 19, no. 12, pp. 2226–2238, 2002.
[79]  L. D. Alcaraz, G. Olmedo, G. Bonilla et al., “The genome of Bacillus coahuilensis reveals adaptations essential for survival in the relic of an ancient marine environment,” Proceedings of the National Academy of Sciences of the United States of America, vol. 105, no. 15, pp. 5803–5808, 2008.
[80]  K. Penn, C. Jenkins, M. Nett et al., “Genomic islands link secondary metabolism to functional adaptation in marine Actinobacteria,” ISME Journal, vol. 3, no. 10, pp. 1193–1203, 2009.
[81]  J. Wiedenbeck and F. M. Cohan, “Origins of bacterial diversity through horizontal genetic transfer and adaptation to new ecological niches,” FEMS Microbiology Reviews, vol. 35, no. 5, pp. 957–976, 2011.
[82]  X. Wu, S. Monchy, S. Taghavi, W. Zhu, J. Ramos, and D. van der Lelie, “Comparative genomics and functional analysis of niche-specific adaptation in Pseudomonas putida,” FEMS Microbiology Reviews, vol. 35, no. 2, pp. 299–323, 2011.
[83]  E. A. Groisman and H. Ochman, “Pathogenicity islands: bacterial evolution in quantum leaps,” Cell, vol. 87, no. 5, pp. 791–794, 1996.
[84]  H. Ochman and E. A. Groisman, “Distribution of pathogenicity islands in Salmonella spp.,” Infection and Immunity, vol. 64, no. 12, pp. 5410–5412, 1996.
[85]  E. F. Boyd and D. L. Hartl, “Salmonella virulence plasmid: modular acquisition of the spv virulence region by an F-plasmid in Salmonella enterica subspecies I and insertion into the chromosome of subspecies II, IIIa, IV and VII isolates,” Genetics, vol. 149, no. 3, pp. 1183–1190, 1998.
[86]  J. Hacker and J. B. Kaper, “Pathogenicity islands and the evolution of microbes,” Annual Review of Microbiology, vol. 54, pp. 641–679, 2000.
[87]  H. Ochman, J. G. Lawrence, and E. A. Grolsman, “Lateral gene transfer and the nature of bacterial innovation,” Nature, vol. 405, no. 6784, pp. 299–304, 2000.
[88]  H. Ochman and N. A. Moran, “Genes lost and genes found: evolution of bacterial pathogenesis and symbiosis,” Science, vol. 292, no. 5519, pp. 1096–1098, 2001.
[89]  I. K. Toth, L. Pritchard, and P. R. J. Birch, “Comparative genomics reveals what makes an enterobacterial plant pathogen,” Annual Review of Phytopathology, vol. 44, pp. 305–336, 2006.
[90]  T. P. Stinear, T. Seemann, P. F. Harrison et al., “Insights from the complete genome sequence of Mycobacterium marinum on the evolution of Mycobacterium tuberculosis,” Genome Research, vol. 18, no. 5, pp. 729–741, 2008.
[91]  E. C. Berglund, A. C. Frank, A. Calteau et al., “Run-off replication of host-adaptability genes is associated with gene transfer agents in the genome of mouse-infecting Bartonella grahamii,” PLoS Genetics, vol. 5, no. 7, Article ID e1000546, 2009.
[92]  M. Marchetti, D. Capela, M. Glew et al., “Experimental evolution of a plant pathogen into a legume symbiont,” PLoS Biology, vol. 8, no. 1, Article ID e1000280, 2010.
[93]  M. W. Silby, C. Winstanley, S. A. Godfrey, S. B. Levy, and R. W. Jackson, “Pseudomonas genomes: diverse and adaptable,” FEMS Microbiology Reviews, vol. 35, no. 4, pp. 652–680, 2011.
[94]  M. A. Schmidt, “LEEways: tales of EPEC, ATEC and EHEC,” Cellular Microbiology, vol. 12, no. 11, pp. 1544–1552, 2010.
[95]  F. De la Cruz and J. Davies, “Horizontal gene transfer and the origin of species: lessons from bacteria,” Trends in Microbiology, vol. 8, no. 3, pp. 128–133, 2000.
[96]  F. M. Cohan, “Bacterial species and speciation,” Systematic Biology, vol. 50, no. 4, pp. 513–524, 2001.
[97]  L. Chistoserdova, J. A. Vorholt, R. K. Thauer, and M. E. Lidstrom, “C1 transfer enzymes and coenzymes linking methylotrophic bacteria and methanogenic archaea,” Science, vol. 281, no. 5373, pp. 99–102, 1998.
[98]  J. Xiong, K. Inoue, and C. E. Bauer, “Tracking molecular evolution of photosynthesis by characterization of a major photosynthesis gene cluster from Heliobacillus mobilis,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 25, pp. 14851–14856, 1998.
[99]  P. S. Novichkov, M. V. Omelchenko, M. S. Gelfand, A. A. Mironov, Y. I. Wolf, and E. V. Koonin, “Genome-wide molecular clock and horizontal gene transfer in bacterial evolution,” Journal of Bacteriology, vol. 186, no. 19, pp. 6575–6585, 2004.
[100]  W. Hao and G. Brian Golding, “The fate of laterally transferred genes: life in the fast lane to adaptation or death,” Genome Research, vol. 16, no. 5, pp. 636–643, 2006.
[101]  J. Haiko, L. Laakkonen, B. Westerlund-Wikstr?m, and T. K. Korhonen, “Molecular adaptation of a plant-bacterium outer membrane protease towards plague virulence factor Pla,” BMC Evolutionary Biology, vol. 11, no. 1, article 43, 2011.
[102]  S. D. Hooper and O. G. Berg, “Duplication is more common among laterally transferred genes than among indigenous genes,” Genome Biology, vol. 4, no. 8, article R48, 2003.
[103]  I. Matic, C. Rayssiguier, and M. Radman, “Interspecies gene exchange in bacteria: the role of SOS and mismatch repair systems in evolution of species,” Cell, vol. 80, no. 3, pp. 507–515, 1995.
[104]  M. K. Ashby and J. Houmard, “Cyanobacterial two-component proteins: structure, diversity, distribution, and evolution,” Microbiology and Molecular Biology Reviews, vol. 70, no. 2, pp. 472–509, 2006.
[105]  M. Brigulla and W. Wackernagel, “Molecular aspects of gene transfer and foreign DNA acquisition in prokaryotes with regard to safety issues,” Applied Microbiology and Biotechnology, vol. 86, no. 4, pp. 1027–1041, 2010.
[106]  S. Yang, J. R. Arguello, X. Li et al., “Repetitive element-mediated recombination as a mechanism for new gene origination in Drosophila,” PLoS Genetics, vol. 4, no. 1, article e3, pp. 0078–0087, 2008.
[107]  V. Daubin, E. Lerat, and G. Perrière, “The source of laterally transferred genes in bacterial genomes,” Genome Biology, vol. 4, no. 9, article R57, 2003.
[108]  R. Jain, M. C. Rivera, and J. A. Lake, “Horizontal gene transfer among genomes: the complexity hypothesis,” Proceedings of the National Academy of Sciences of the United States of America, vol. 96, no. 7, pp. 3801–3806, 1999.
[109]  C. Brochier, E. Bapteste, D. Moreira, and H. Philippe, “Eubacterial phylogeny based on translational apparatus proteins,” Trends in Genetics, vol. 18, no. 1, pp. 1–5, 2002.
[110]  E. W. Brown, J. E. LeClerc, B. Li, W. L. Payne, and T. A. Cebula, “Phylogenetic evidence for horizontal transfer of mutS alleles among naturally occurring Escherichia coli strains,” Journal of Bacteriology, vol. 183, no. 5, pp. 1631–1644, 2001.
[111]  Y. Nakamura, T. Itoh, H. Matsuda, and T. Gojobori, “Biased biological functions of horizontally-transferred genes in prokaryotic genomes,” Nature Genetics, vol. 36, no. 7, pp. 760–766, 2004.
[112]  F. D. Ciccarelli, T. Doerks, C. Von Mering, C. J. Creevey, B. Snel, and P. Bork, “Toward automatic reconstruction of a highly resolved tree of life,” Science, vol. 311, no. 5765, pp. 1283–1287, 2006.
[113]  R. Sorek, Y. Zhu, C. J. Creevey, M. P. Francino, P. Bork, and E. M. Rubin, “Genome-wide experimental determination of barriers to horizontal gene transfer,” Science, vol. 318, no. 5855, pp. 1449–1452, 2007.
[114]  S. S. Abby, E. Tannier, M. Gouy, and V. Daubin, “Lateral gene transfer as a support for the tree of life,” Proceedings of the National Academy of Sciences of the United States of America, vol. 109, no. 13, pp. 4962–4967, 2012.
[115]  R. L. Charlebois and W. F. Doolittle, “Computing prokaryotic gene ubiquity: rescuing the core from extinction,” Genome Research, vol. 14, no. 12, pp. 2469–2477, 2004.
[116]  S. R. Santos and H. Ochman, “Identification and phylogenetic sorting of bacterial lineages with universally conserved genes and proteins,” Environmental Microbiology, vol. 6, no. 7, pp. 754–759, 2004.
[117]  E. Lerat, V. Daubin, and N. A. Moran, “From gene trees to organismal phylogeny in prokaryotes: the case of the γ-Proteobacteria,” PLoS Biology, vol. 1, no. 1, article E19, 2003.
[118]  O. Lukjancenko, T. M. Wassenaar, and D. W. Ussery, “Comparison of 61 sequenced Escherichia coli genomes,” Microbial Ecology, vol. 60, no. 4, pp. 708–720, 2010.
[119]  A. Jacobsen, R. S. Hendriksen, F. M. Aaresturp, D. W. Ussery, and C. Friis, “The Salmonella enterica Pan-genome,” Microbial Ecology, vol. 62, no. 3, pp. 487–504, 2011.
[120]  M. Touchon, C. Hoede, O. Tenaillon et al., “Organised genome dynamics in the Escherichia coli species results in highly diverse adaptive paths,” PLoS Genetics, vol. 5, no. 1, Article ID e1000344, 2009.
[121]  J. G. Lawrence and H. Ochman, “Molecular archaeology of the Escherichia coli genome,” Proceedings of the National Academy of Sciences of the United States of America, vol. 95, no. 16, pp. 9413–9417, 1998.
[122]  E. Lerat, V. Daubin, H. Ochman, and N. A. Moran, “Evolutionary origins of genomic repertoires in bacteria,” PLoS Biology, vol. 3, no. 5, article e130, 2005.
[123]  C. H. Kuo and H. Ochman, “The fate of new bacterial genes,” FEMS Microbiology Reviews, vol. 33, no. 1, pp. 38–43, 2009.
[124]  F. M. Cohan, “Towards a conceptual and operational union of bacterial systematics, ecology, and evolution,” Philosophical Transactions of the Royal Society B, vol. 361, no. 1475, pp. 1985–1996, 2006.
[125]  E. Maisonneuve, L. J. Shakespeare, M. G. J?rgensen, and K. Gerdes, “Bacterial persistence by RNA endonucleases,” Proceedings of the National Academy of Sciences of the United States of America, vol. 108, no. 32, pp. 13206–13211, 2011.
[126]  O. Vesper, S. Amitai, M. Belitsky et al., “Selective translation of leaderless mRNAs by specialized ribosomes generated by MazF in Escherichia coli,” Cell, vol. 147, no. 1, pp. 147–157, 2011.
[127]  J. E. Rebollo, V. Francois, and J. M. Louarn, “Detection and possible role of two large nondivisible zones on the Escherichia coli chromosome,” Proceedings of the National Academy of Sciences of the United States of America, vol. 85, no. 24, pp. 9391–9395, 1988.
[128]  F. R. Blattner, G. Plunkett, C. A. Bloch et al., “The complete genome sequence of Escherichia coli K-12,” Science, vol. 277, no. 5331, pp. 1453–1462, 1997.
[129]  F. Boccard, E. Esnault, and M. Valens, “Spatial arrangement and macrodomain organization of bacterial chromosomes,” Molecular Microbiology, vol. 57, no. 1, pp. 9–16, 2005.
[130]  E. Esnault, M. Valens, O. Espéli, and F. Boccard, “Chromosome structuring limits genome plasticity in Escherichia coli,” PLoS genetics, vol. 3, no. 12, article e226, 2007.
[131]  P. Sobetzko, A. Travers, and G. Muskhelishvili, “Gene order and chromosome dynamics coordinate spatiotemporal gene expression during the bacterial growth cycle,” Proceedings of the National Academy of Sciences of the United States of America, vol. 109, no. 2, pp. E42–E50, 2012.
[132]  K. P. Williams, “Integration sites for genetic elements in prokaryotic tRNA and tmRNA genes: sublocation preference of integrase subfamilies,” Nucleic Acids Research, vol. 30, no. 4, pp. 866–875, 2002.
[133]  T. Ikemura, “Codon usage and tRNA content in unicellular and multicellular organisms,” Molecular Biology and Evolution, vol. 2, no. 1, pp. 13–34, 1985.
[134]  P. M. Sharp and W. H. Li, “An evolutionary perspective on synonymous codon usage in unicellular organisms,” Journal of Molecular Evolution, vol. 24, no. 1-2, pp. 28–38, 1986.
[135]  M. Bulmer, “The selection-mutation-drift theory of synonymous codon usage,” Genetics, vol. 129, no. 3, pp. 897–907, 1991.
[136]  S. Kanaya, Y. Yamada, Y. Kudo, and T. Ikemura, “Studies of codon usage and tRNA genes of 18 unicellular organisms and quantification of Bacillus subtilis tRNAs: gene expression level and species-specific diversity of codon usage based on multivariate analysis,” Gene, vol. 238, no. 1, pp. 143–155, 1999.
[137]  E. P. C. Rocha, “Codon usage bias from tRNA's point of view: redundancy, specialization, and efficient decoding for translation optimization,” Genome Research, vol. 14, no. 11, pp. 2279–2286, 2004.
[138]  D. Amorós-Moya, S. Bedhomme, M. Hermann, and I. G. Bravo, “Evolution in regulatory regions rapidly compensates the cost of nonoptimal codon usage,” Molecular Biology and Evolution, vol. 27, no. 9, pp. 2141–2151, 2010.
[139]  A. Medrano-Soto, G. Moreno-Hagelsieb, P. Vinuesa, J. A. Christen, and J. Collado-Vides, “Successful lateral transfer requires codon usage compatibility between foreign genes and recipient genomes,” Molecular Biology and Evolution, vol. 21, no. 10, pp. 1884–1894, 2004.
[140]  J. G. Lawrence and H. Ochman, “Amelioration of bacterial genomes: rates of change and exchange,” Journal of Molecular Evolution, vol. 44, no. 4, pp. 383–397, 1997.
[141]  G. S. Vernikos, N. R. Thomson, and J. Parkhill, “Genetic flux over time in the Salmonella lineage,” Genome Biology, vol. 8, no. 6, article R100, 2007.
[142]  M. N. Price, P. S. Dehal, and A. P. Arkin, “Horizontal gene transfer and the evolution of transcriptional regulation in Escherichia coli,” Genome Biology, vol. 9, no. 1, article R4, 2008.
[143]  O. Gal-Mor, D. Elhadad, W. Deng, G. Rahav, and B. B. Finlay, “The Salmonella enterica PhoP directly activates the horizontally acquired SPI-2 gene ssel and is functionally different from a S. bongori ortholog,” PLoS ONE, vol. 6, no. 5, Article ID e20024, 2011.
[144]  L. C. Martínez, H. Yakhnin, M. I. Camacho et al., “Integration of a complex regulatory cascade involving the SirA/BarA and Csr global regulatory systems that controls expression of the Salmonella SPI-1 and SPI-2 virulence regulons through HilD,” Molecular Microbiology, vol. 80, no. 6, pp. 1637–1656, 2011.
[145]  M. J. Lercher and C. Pál, “Integration of horizontally transferred genes into regulatory interaction networks takes many million years,” Molecular Biology and Evolution, vol. 25, no. 3, pp. 559–567, 2008.
[146]  W. Davids and Z. Zhang, “The impact of horizontal gene transfer in shaping operons and protein interaction networks—direct evidence of preferential attachment,” BMC Evolutionary Biology, vol. 8, no. 1, article 23, 2008.
[147]  C. Pál, B. Papp, and M. J. Lercher, “Adaptive evolution of bacterial metabolic networks by horizontal gene transfer,” Nature Genetics, vol. 37, no. 12, pp. 1372–1375, 2005.
[148]  A. Kreimer, E. Borenstein, U. Gophna, and E. Ruppin, “The evolution of modularity in bacterial metabolic networks,” Proceedings of the National Academy of Sciences of the United States of America, vol. 105, no. 19, pp. 6976–6981, 2008.
[149]  M. D'Antonio and F. D. Ciccarelli, “Modification of gene duplicability during the evolution of protein interaction network,” PLoS Computational Biology, vol. 7, no. 4, Article ID e1002029, 2011.
[150]  A. Hintze and C. Adami, “Evolution of complex modular biological networks,” PLoS Computational Biology, vol. 4, no. 2, article e23, 2008.
[151]  C. Espinosa-Soto and A. Wagner, “Specialization can drive the evolution of modularity,” PLoS Computational Biology, vol. 6, no. 3, article e1000719, 2010.
[152]  Y. Bar-Yam, Dynamics of Complex Systems, Westview Press, Boulder, CO, USA, 1997.
[153]  J. G. Lawrence, “Microbial evolution: enforcing cooperation by partial kin selection,” Current Biology, vol. 19, no. 20, pp. R943–R945, 2009.
[154]  T. Nogueira, D. J. Rankin, M. Touchon, F. Taddei, S. P. Brown, and E. P. C. Rocha, “Horizontal gene transfer of the secretome drives the evolution of bacterial cooperation and virulence,” Current Biology, vol. 19, no. 20, pp. 1683–1691, 2009.
[155]  S. E. Mc Ginty, D. J. Rankin, and S. P. Brown, “Horizontal gene transfer and the evolution of bacterial cooperation,” Evolution, vol. 65, no. 1, pp. 21–32, 2011.
[156]  D. J. Rankin, E. P. C. Rocha, and S. P. Brown, “What traits are carried on mobile genetic elements, and why,” Heredity, vol. 106, no. 1, pp. 1–10, 2011.
[157]  L. E. de Vries, Y. Vallès, Y. Agers? et al., “The gut as reservoir of antibiotic resistance: microbial diversity of tetracycline resistance in mother and infant,” PLoS ONE, vol. 6, no. 6, Article ID e21644, 2011.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133

WeChat 1538708413