首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Imprinted gene expression corresponds to parental allele-specific DNA CpG methylation and chromatin composition. Histone tail covalent modifications have been extensively studied, but it is not known whether modifications in the histone globular domains can also discriminate between the parental alleles. Using multiplex chromatin immunoprecipitation-single nucleotide primer extension (ChIP-SNuPE) assays, we measured the allele-specific enrichment of H3K79 methylation and H4K91 acetylation along the H19/Igf2 imprinted domain. Whereas H3K79me1, H3K79me2, and H4K91ac displayed a paternal-specific enrichment at the paternally expressed Igf2 locus, H3K79me3 was paternally biased at the maternally expressed H19 locus, including the paternally methylated imprinting control region (ICR). We found that these allele-specific differences depended on CTCF binding in the maternal ICR allele. We analyzed an additional 11 differentially methylated regions (DMRs) and found that, in general, H3K79me3 was associated with the CpG-methylated alleles, whereas H3K79me1, H3K79me2, and H4K91ac enrichment was specific to the unmethylated alleles. Our data suggest that allele-specific differences in the globular histone domains may constitute a layer of the “histone code” at imprinted genes.Imprinted genes are defined by the characteristic monoallelic silencing of either the paternally or maternally inherited allele. Most imprinted genes exist in imprinted gene clusters (10), and these clusters are usually associated with one or more differentially methylated regions (DMRs) (27, 65). DNA methylation at DMRs is essential for the allele-specific expression of most imprinted genes (31). Maternal or paternal allele-specific DNA methylation of a subset of DMRs (germ line DMRs) is gamete specific (27, 39). These maternal or paternal methylation differences are established during oogenesis or spermatogenesis, respectively, by the de novo DNA methyltransferases Dnmt3a and Dnmt3b together with Dnmt3L (5, 26, 48). The gamete-specific methylation differences set the stage for the parental allele-specific action of germ line DMRs, some of which have been shown to control the monoallelic expression of the associated genes in the respective domains (11, 34, 36, 53, 66, 71-73, 77). These DMRs are called imprinting control regions (ICRs).Two recurring themes have been reported for ICR action. ICRs can function as DNA methylation-regulated promoters of a noncoding RNA or as methylation-regulated insulators. Recent evidence suggests that both of these mechanisms involve chromatin organization by either the noncoding RNA (45, 50) or the CTCF insulator protein (17, 32) along the respective imprinted domains. The CTCF insulator binds in the unmethylated maternal allele of the H19/Igf2 ICR and blocks the access of the Igf2 promoters to the shared downstream enhancers. CTCF cannot bind in the methylated paternal ICR allele; hence, here the Igf2 promoters have access to the enhancers (4, 18, 24, 25, 62). When CTCF binding is abolished in the ICR of the maternal allele, Igf2 expression becomes biallelic, and H19 expression is missing from both alleles (17, 52, 58, 63). Importantly, CTCF is the single major organizer of the allele-specific chromatin along the H19/Igf2 imprinted domain (17). Significantly, CTCF recruits, at a distance, Polycomb-mediated H3K27me3 repressive marks at the Igf2 promoter and at the Igf2 DMRs (17, 32).A role for chromatin composition is suggested in the parental allele-specific expression of imprinted genes. Repressive histone tail covalent modifications, such as H3K9me2 H3K9me3, H4K20me3, H3K27me3, and the symmetrically methylated H4R3me2 marks, are generally associated with the methylated DMR alleles, while activating histone tail covalent modifications, such as acetylated histone tails and also H3K4me2 and H3K4me3, are characteristic of the unmethylated alleles (7-9, 12-15, 17, 21, 33, 35, 43, 44, 51, 55, 56, 67, 69, 74, 75). Importantly, the maintenance of imprinted gene expression depends on the allele-specific chromatin differences. ICR-dependent H3K9me2 and H3K27me3 enrichment in the paternal allele (67) is required for paternal repression of a set of imprinted genes along the Kcnq1 imprinted domain in the placenta (30). Imprinted Cdkn1c and Cd81 expression depends on H3K27 methyltransferase Ezh2 activity in the extraembryonic ectoderm (64). Similarly, H3K9 methyltransferase Ehmt2 is required for parental allele-specific expression of a number of imprinted genes, including Osbpl5, Cd81, Ascl2, Tfpi2, and Slc22a3 in the placenta (44, 45, 70).There is increasing evidence that covalent modifications, not only in the histone tails but also in the histone globular domains, carry essential information for development and gene regulation. The H3K79 methyltransferase gene is essential for development in Drosophila (60) and in mice (22). H3K79 methylation is required for telomeric heterochromatin silencing in Drosophila (60), Saccharomyces cerevisiae (47, 68), and mice (22). The H4K91 residue regulates nucleosome assembly (76). Whereas mutations at single acetylation sites in the histone tails have only minor consequences, mutation of the H4K91 site in the histone H4 globular domain causes severe defects in silent chromatin formation and DNA repair in yeast (37, 42, 76).Contrary to the abundant information that exists regarding the allele-specific chromatin composition at DMRs of imprinted genes, no information is available about the parental allele-specific marking in the histone globular domains at the DMRs. We hypothesized that chromatin marks in the globular domains of histones also distinguish the parental alleles of germ line DMRs. In order to demonstrate this, we measured the allele-specific enrichment of H3K79me1, H3K79me2, H3K79me3, and H4K91ac at 11 mouse DMRs using quantitative multiplex chromatin immunoprecipitation-single nucleotide primer extension (ChIP-SNuPE) assays. In general, H3K79me3 was associated with the methylated allele at most DMRs, whereas the unmethylated allele showed enrichment for H3K79me1, H3K79me2, and H4K91ac. These results are consistent with the possibility that allele-specific differences in the globular domains of histones contribute to the “histone code” at DMRs.  相似文献   

3.
4.
5.
6.
Analysis of Lyme borreliosis (LB) spirochetes, using a novel multilocus sequence analysis scheme, revealed that OspA serotype 4 strains (a rodent-associated ecotype) of Borrelia garinii were sufficiently genetically distinct from bird-associated B. garinii strains to deserve species status. We suggest that OspA serotype 4 strains be raised to species status and named Borrelia bavariensis sp. nov. The rooted phylogenetic trees provide novel insights into the evolutionary history of LB spirochetes.Multilocus sequence typing (MLST) and multilocus sequence analysis (MLSA) have been shown to be powerful and pragmatic molecular methods for typing large numbers of microbial strains for population genetics studies, delineation of species, and assignment of strains to defined bacterial species (4, 13, 27, 40, 44). To date, MLST/MLSA schemes have been applied only to a few vector-borne microbial populations (1, 6, 30, 37, 40, 41, 47).Lyme borreliosis (LB) spirochetes comprise a diverse group of zoonotic bacteria which are transmitted among vertebrate hosts by ixodid (hard) ticks. The most common agents of human LB are Borrelia burgdorferi (sensu stricto), Borrelia afzelii, Borrelia garinii, Borrelia lusitaniae, and Borrelia spielmanii (7, 8, 12, 35). To date, 15 species have been named within the group of LB spirochetes (6, 31, 32, 37, 38, 41). While several of these LB species have been delineated using whole DNA-DNA hybridization (3, 20, 33), most ecological or epidemiological studies have been using single loci (5, 9-11, 29, 34, 36, 38, 42, 51, 53). Although some of these loci have been convenient for species assignment of strains or to address particular epidemiological questions, they may be unsuitable to resolve evolutionary relationships among LB species, because it is not possible to define any outgroup. For example, both the 5S-23S intergenic spacer (5S-23S IGS) and the gene encoding the outer surface protein A (ospA) are present only in LB spirochete genomes (36, 43). The advantage of using appropriate housekeeping genes of LB group spirochetes is that phylogenetic trees can be rooted with sequences of relapsing fever spirochetes. This renders the data amenable to detailed evolutionary studies of LB spirochetes.LB group spirochetes differ remarkably in their patterns and levels of host association, which are likely to affect their population structures (22, 24, 46, 48). Of the three main Eurasian Borrelia species, B. afzelii is adapted to rodents, whereas B. valaisiana and most strains of B. garinii are maintained by birds (12, 15, 16, 23, 26, 45). However, B. garinii OspA serotype 4 strains in Europe have been shown to be transmitted by rodents (17, 18) and, therefore, constitute a distinct ecotype within B. garinii. These strains have also been associated with high pathogenicity in humans, and their finer-scale geographical distribution seems highly focal (10, 34, 52, 53).In this study, we analyzed the intra- and interspecific phylogenetic relationships of B. burgdorferi, B. afzelii, B. garinii, B. valaisiana, B. lusitaniae, B. bissettii, and B. spielmanii by means of a novel MLSA scheme based on chromosomal housekeeping genes (30, 48).  相似文献   

7.
Phenoxyalkanoic acid (PAA) herbicides are widely used in agriculture. Biotic degradation of such herbicides occurs in soils and is initiated by α-ketoglutarate- and Fe2+-dependent dioxygenases encoded by tfdA-like genes (i.e., tfdA and tfdAα). Novel primers and quantitative kinetic PCR (qPCR) assays were developed to analyze the diversity and abundance of tfdA-like genes in soil. Five primer sets targeting tfdA-like genes were designed and evaluated. Primer sets 3 to 5 specifically amplified tfdA-like genes from soil, and a total of 437 sequences were retrieved. Coverages of gene libraries were 62 to 100%, up to 122 genotypes were detected, and up to 389 genotypes were predicted to occur in the gene libraries as indicated by the richness estimator Chao1. Phylogenetic analysis of in silico-translated tfdA-like genes indicated that soil tfdA-like genes were related to those of group 2 and 3 Bradyrhizobium spp., Sphingomonas spp., and uncultured soil bacteria. Soil-derived tfdA-like genes were assigned to 11 clusters, 4 of which were composed of novel sequences from this study, indicating that soil harbors novel and diverse tfdA-like genes. Correlation analysis of 16S rRNA and tfdA-like gene similarity indicated that any two bacteria with D > 20% of group 2 tfdA-like gene-derived protein sequences belong to different species. Thus, data indicate that the soil analyzed harbors at least 48 novel bacterial species containing group 2 tfdA-like genes. Novel qPCR assays were established to quantify such new tfdA-like genes. Copy numbers of tfdA-like genes were 1.0 × 106 to 65 × 106 per gram (dry weight) soil in four different soils, indicating that hitherto-unknown, diverse tfdA-like genes are abundant in soils.Phenoxyalkanoic acid (PAA) herbicides such as MCPA (4-chloro-2-methyl-phenoxyacetic acid) and 2,4-D (2,4-dichlorophenoxyacetic acid) are widely used to control broad-leaf weeds in agricultural as well as nonagricultural areas (19, 77). Degradation occurs primarily under oxic conditions in soil, and microorganisms play a key role in the degradation of such herbicides in soil (62, 64). Although relatively rapidly degraded in soil (32, 45), both MCPA and 2,4-D are potential groundwater contaminants (10, 56, 70), accentuating the importance of bacterial PAA herbicide-degrading bacteria in soils (e.g., references 3, 5, 6, 20, 41, 59, and 78).Degradation can occur cometabolically or be associated with energy conservation (15, 54). The first step in the degradation of 2,4-D and MCPA is initiated by the product of cadAB or tfdA-like genes (29, 30, 35, 67), which constitutes an α-ketoglutarate (α-KG)- and Fe2+-dependent dioxygenase. TfdA removes the acetate side chain of 2,4-D and MCPA to produce 2,4-dichlorophenol and 4-chloro-2-methylphenol, respectively, and glyoxylate while oxidizing α-ketoglutarate to CO2 and succinate (16, 17).Organisms capable of PAA herbicide degradation are phylogenetically diverse and belong to the Alpha-, Beta-, and Gammproteobacteria and the Bacteroidetes/Chlorobi group (e.g., references 2, 14, 29-34, 39, 60, 68, and 71). These bacteria harbor tfdA-like genes (i.e., tfdA or tfdAα) and are categorized into three groups on an evolutionary and physiological basis (34). The first group consists of beta- and gammaproteobacteria and can be further divided into three distinct classes based on their tfdA genes (30, 46). Class I tfdA genes are closely related to those of Cupriavidus necator JMP134 (formerly Ralstonia eutropha). Class II tfdA genes consist of those of Burkholderia sp. strain RASC and a few strains that are 76% identical to class I tfdA genes. Class III tfdA genes are 77% identical to class I and 80% identical to class II tfdA genes and linked to MCPA degradation in soil (3). The second group consists of alphaproteobacteria, which are closely related to Bradyrhizobium spp. with tfdAα genes having 60% identity to tfdA of group 1 (18, 29, 34). The third group also harbors the tfdAα genes and consists of Sphingomonas spp. within the alphaproteobacteria (30).Diverse PAA herbicide degraders of all three groups were identified in soil by cultivation-dependent studies (32, 34, 41, 78). Besides CadAB, TfdA and certain TfdAα proteins catalyze the conversion of PAA herbicides (29, 30, 35). All groups of tfdA-like genes are potentially linked to the degradation of PAA herbicides, although alternative primary functions of group 2 and 3 TfdAs have been proposed (30, 35). However, recent cultivation-independent studies focused on 16S rRNA genes or solely on group 1 tfdA sequences in soil (e.g., references 3-5, 13, and 41). Whether group 2 and 3 tfdA-like genes are also quantitatively linked to the degradation of PAA herbicides in soils is unknown. Thus, tools to target a broad range of tfdA-like genes are needed to resolve such an issue. Primers used to assess the diversity of tfdA-like sequences used in previous studies were based on the alignment of approximately 50% or less of available sequences to date (3, 20, 29, 32, 39, 47, 58, 73). Primers specifically targeting all major groups of tfdA-like genes to assess and quantify a broad diversity of potential PAA degraders in soil are unavailable. Thus, the objectives of this study were (i) to develop primers specific for all three groups of tfdA-like genes, (ii) to establish quantitative kinetic PCR (qPCR) assays based on such primers for different soil samples, and (iii) to assess the diversity and abundance of tfdA-like genes in soil.  相似文献   

8.
9.
10.
Immunogold localization revealed that OmcS, a cytochrome that is required for Fe(III) oxide reduction by Geobacter sulfurreducens, was localized along the pili. The apparent spacing between OmcS molecules suggests that OmcS facilitates electron transfer from pili to Fe(III) oxides rather than promoting electron conduction along the length of the pili.There are multiple competing/complementary models for extracellular electron transfer in Fe(III)- and electrode-reducing microorganisms (8, 18, 20, 44). Which mechanisms prevail in different microorganisms or environmental conditions may greatly influence which microorganisms compete most successfully in sedimentary environments or on the surfaces of electrodes and can impact practical decisions on the best strategies to promote Fe(III) reduction for bioremediation applications (18, 19) or to enhance the power output of microbial fuel cells (18, 21).The three most commonly considered mechanisms for electron transfer to extracellular electron acceptors are (i) direct contact between redox-active proteins on the outer surfaces of the cells and the electron acceptor, (ii) electron transfer via soluble electron shuttling molecules, and (iii) the conduction of electrons along pili or other filamentous structures. Evidence for the first mechanism includes the necessity for direct cell-Fe(III) oxide contact in Geobacter species (34) and the finding that intensively studied Fe(III)- and electrode-reducing microorganisms, such as Geobacter sulfurreducens and Shewanella oneidensis MR-1, display redox-active proteins on their outer cell surfaces that could have access to extracellular electron acceptors (1, 2, 12, 15, 27, 28, 31-33). Deletion of the genes for these proteins often inhibits Fe(III) reduction (1, 4, 7, 15, 17, 28, 40) and electron transfer to electrodes (5, 7, 11, 33). In some instances, these proteins have been purified and shown to have the capacity to reduce Fe(III) and other potential electron acceptors in vitro (10, 13, 29, 38, 42, 43, 48, 49).Evidence for the second mechanism includes the ability of some microorganisms to reduce Fe(III) that they cannot directly contact, which can be associated with the accumulation of soluble substances that can promote electron shuttling (17, 22, 26, 35, 36, 47). In microbial fuel cell studies, an abundance of planktonic cells and/or the loss of current-producing capacity when the medium is replaced is consistent with the presence of an electron shuttle (3, 14, 26). Furthermore, a soluble electron shuttle is the most likely explanation for the electrochemical signatures of some microorganisms growing on an electrode surface (26, 46).Evidence for the third mechanism is more circumstantial (19). Filaments that have conductive properties have been identified in Shewanella (7) and Geobacter (41) species. To date, conductance has been measured only across the diameter of the filaments, not along the length. The evidence that the conductive filaments were involved in extracellular electron transfer in Shewanella was the finding that deletion of the genes for the c-type cytochromes OmcA and MtrC, which are necessary for extracellular electron transfer, resulted in nonconductive filaments, suggesting that the cytochromes were associated with the filaments (7). However, subsequent studies specifically designed to localize these cytochromes revealed that, although the cytochromes were extracellular, they were attached to the cells or in the exopolymeric matrix and not aligned along the pili (24, 25, 30, 40, 43). Subsequent reviews of electron transfer to Fe(III) in Shewanella oneidensis (44, 45) appear to have dropped the nanowire concept and focused on the first and second mechanisms.Geobacter sulfurreducens has a number of c-type cytochromes (15, 28) and multicopper proteins (12, 27) that have been demonstrated or proposed to be on the outer cell surface and are essential for extracellular electron transfer. Immunolocalization and proteolysis studies demonstrated that the cytochrome OmcB, which is essential for optimal Fe(III) reduction (15) and highly expressed during growth on electrodes (33), is embedded in the outer membrane (39), whereas the multicopper protein OmpB, which is also required for Fe(III) oxide reduction (27), is exposed on the outer cell surface (39).OmcS is one of the most abundant cytochromes that can readily be sheared from the outer surfaces of G. sulfurreducens cells (28). It is essential for the reduction of Fe(III) oxide (28) and for electron transfer to electrodes under some conditions (11). Therefore, the localization of this important protein was further investigated.  相似文献   

11.
12.
13.
14.
Halogenases have been shown to play a significant role in biosynthesis and introducing the bioactivity of many halogenated secondary metabolites. In this study, 54 reduced flavin adenine dinucleotide (FADH2)-dependent halogenase gene-positive strains were identified after the PCR screening of a large collection of 228 reference strains encompassing all major families and genera of filamentous actinomycetes. The wide distribution of this gene was observed to extend to some rare lineages with higher occurrences and large sequence diversity. Subsequent phylogenetic analyses revealed that strains containing highly homologous halogenases tended to produce halometabolites with similar structures, and halogenase genes are likely to propagate by horizontal gene transfer as well as vertical inheritance within actinomycetes. Higher percentages of halogenase gene-positive strains than those of halogenase gene-negative ones contained polyketide synthase genes and/or nonribosomal peptide synthetase genes or displayed antimicrobial activities in the tests applied, indicating their genetic and physiological potentials for producing secondary metabolites. The robustness of this halogenase gene screening strategy for the discovery of particular biosynthetic gene clusters in rare actinomycetes besides streptomycetes was further supported by genome-walking analysis. The described distribution and phylogenetic implications of the FADH2-dependent halogenase gene present a guide for strain selection in the search for novel organohalogen compounds from actinomycetes.It is well known that actinomycetes, notably filamentous actinomycetes, have a remarkable capacity to produce bioactive molecules for drug development (4, 6). However, novel technologies are demanded for the discovery of new bioactive secondary metabolites from these microbes to meet the urgent medical need for drug candidates (5, 9, 31).Genome mining recently has been used to search for new drug leads (7, 20, 42, 51). Based on the hypothesis that secondary metabolites with similar structures are biosynthesized by gene clusters that harbor certain homologous genes, such homologous genes could serve as suitable markers for distinct natural-product gene clusters (26, 51). A wide range of structurally diverse bioactive compounds are synthesized by polyketide synthase (PKS) and nonribosomal peptide synthetase (NRPS) systems in actinomycetes, therefore much attention has been given to revealing a previously unrecognized biosynthetic potential of actinomycetes through the genome mining of these genes (2, 3, 22). However, the broad distribution of PKS and NRPS genes and their high numbers even in a single actinomycete complicate their use (2, 3). To rationally exploit the genetic potential of actinomycetes, more and more special genes, such as tailoring enzyme genes, are being utilized for this sequence-guided genetic screening strategy (20, 38).Tailoring enzymes, which are responsible for the introduction and generation of diversity and bioactivity in several structural classes during or after NRPS, PKS, or NRPS/PKS assembly lines, usually include acyltransferases, aminotransferases, cyclases, glycosyltransferases, halogenases, ketoreductases, methyltransferases, and oxygenases (36, 45). Halogenation, an important feature for the bioactivity of a large number of distinct natural products (16, 18, 30), frequently is introduced by one type of halogenase, called reduced flavin adenine dinucleotide (FADH2)-dependent (or flavin-dependent) halogenase (10, 12, 35). More than 4,000 halometabolites have been discovered (15), including commercially important antibiotics such as chloramphenicol, vancomycin, and teicoplanin (43).Previous investigations of FADH2-dependent halogenase genes were focused largely on related gene clusters in the genera Amycolatopsis (33, 44, 53) and Streptomyces (8, 10, 21, 27, 32, 34, 47-49) and also on those in the genera Actinoplanes (25), Actinosynnema (50), Micromonospora (1), and Nonomuraea (39); however, none of these studies has led to the rest of the major families and genera of actinomycetes. In addition, there is evidence that FADH2-dependent halogenase genes of streptomycetes usually exist in halometabolite biosynthetic gene clusters (20), but we lack knowledge of such genes and clusters in other actinomycetes.In the present study, we show that the distribution of the FADH2-dependent halogenase gene in filamentous actinomycetes does indeed correlate with the potential for halometabolite production based on other genetic or physiological factors. We also showed that genome walking near the halogenase gene locus could be employed to identify closely linked gene clusters that likely encode pathways for organohalogen compound production in actinomycetes other than streptomycetes.  相似文献   

15.
16.
17.
18.
Magnetosome biomineralization and magnetotaxis in magnetotactic bacteria are controlled by numerous, mostly unknown gene functions that are predominantly encoded by several operons located within the genomic magnetosome island (MAI). Genetic analysis of magnetotactic bacteria has remained difficult and requires the development of novel tools. We established a Cre-lox-based deletion method which allows the excision of large genomic fragments in Magnetospirillum gryphiswaldense. Two conjugative suicide plasmids harboring lox sites that flanked the target region were subsequently inserted into the chromosome by homologous recombination, requiring only one single-crossover event, respectively, and resulting in a double cointegrate. Excision of the targeted chromosomal segment that included the inserted plasmids and their resistance markers was induced by trans expression of Cre recombinase, which leaves behind a scar of only a single loxP site. The Cre helper plasmid was then cured from the deletant strain by relief of antibiotic selection. We have used this method for the deletion of 16.3-kb, 61-kb, and 67.3-kb fragments from the genomic MAI, either in a single round or in subsequent rounds of deletion, covering a region of approximately 87 kb that comprises the mamAB, mms6, and mamGFDC operons. As expected, all mutants were Mag and some were Mot; otherwise, they showed normal growth patterns, which indicates that the deleted region is not essential for viability in the laboratory. The method will facilitate future functional analysis of magnetosome genes and also can be utilized for large-scale genome engineering in magnetotactic bacteria.Magnetosomes are unique membrane-enveloped organelles that are formed by magnetotactic bacteria (MTB) for magnetic navigation (2, 37). The mechanism of magnetosome formation is within the focus of a multidisciplinary interest and has relevance for biotechnological applications (5). It has been recognized that the biomineralization of inorganic magnetite crystals and their assembly into highly ordered magnetosome chains are under strict genetic control. Recent studies combining proteomic and bioinformatic approaches suggested that the genetic determination of magnetosome formation is complex and may potentially involve 25 to 50 gene functions (15), with unknown numbers of accessory genes and those controlling signal transduction and motility to achieve effective magnetotaxis (8, 9, 12, 26, 27, 29). However, the functional characterization of these candidate genes has been lagging behind. This is due to technical difficulties and the lack of facile tools for genetic manipulation of MTB. Allelic replacement systems have been established for Magnetospirillum magneticum (18) and Magnetospirillum gryphiswaldense (39, 40), but so far, there are only few examples of these for magnetosome genes that were functionally characterized because of the tedious and cumbersome procedures required for mutant generation (11, 19, 28, 31-32). Most genes controlling magnetosome formation in these and other MTB are located within a genomic magnetosome island (MAI) (34), which is genetically instable during stationary growth (47) and more or less conserved in other MTB (12, 13, 35). Most known magnetosome genes are organized within several conserved operons, which are interspersed with large, poorly conserved genome sections of unknown functions that have been speculated to represent genetic junk irrelevant for magnetotaxis but to cause genetic instability by their high content of repeats and transposable elements (34, 47). Thus, for large-scale functional genome analysis and rearrangements of the MAI, there is a great need for additional and more efficient genetic methods.Artificial genome recombination systems have been described for a number of bacteria. Many of them are based on the Cre-loxP system of the P1 phage (42). The Cre-loxP recombination system is a simple two-component system that is recognized as a powerful genetic tool in a multitude of eukaryotic and prokaryotic organisms (4, 6, 48). The Cre protein belongs to the integrase family of site-specific recombinases and catalyzes reciprocal site-specific recombination of DNA at 34-bp loxP sites, resulting in either excision or inversion, depending on the parallel or antiparallel orientation of the loxP sites, respectively (21). It does not require any host cofactors or accessory proteins (7). Cre-lox deletion has several advantages over other methods, such as a high efficiency and the independency of the length of DNA located between the two lox sites. The utility of Cre-lox systems has been demonstrated in a wide variety of Gram-positive and Gram-negative bacteria (17, 22-23). In several studies, it was applied for the generation of large-scale deletions, as in for example, the Gram-positive Corynebacterium glutamicum (43-46) and Bacillus subtilis (49).In M. gryphiswaldense, the functionality of a Cre-loxP antibiotic marker recycling system (25) has been previously demonstrated by deletion of a single gene based on double-crossover insertion of two loxP sites, followed by subsequent Cre-mediated excision (31). In this study, we describe a novel strategy for Cre-loxP-mediated deletion of large genomic fragments which requires only two single crossovers. The system has been validated by the generation of three large deletions, two single and one combination within the MAI, which demonstrated that the total deleted region of approximately 87 kb is not essential for viability and growth in the laboratory.  相似文献   

19.
Streptococcus sanguinis is an important cause of infective endocarditis. Previous studies have identified lipoproteins as virulence determinants in other streptococcal species. Using a bioinformatic approach, we identified 52 putative lipoprotein genes in S. sanguinis strain SK36 as well as genes encoding the lipoprotein-processing enzymes prolipoprotein diacylglyceryl transferase (lgt) and signal peptidase II (lspA). We employed a directed signature-tagged mutagenesis approach to systematically disrupt these genes and screen each mutant for the loss of virulence in an animal model of endocarditis. All mutants were viable. In competitive index assays, mutation of a putative phosphate transporter reduced in vivo competitiveness by 14-fold but also reduced in vitro viability by more than 20-fold. Mutations in lgt, lspA, or an uncharacterized lipoprotein gene reduced competitiveness by two- to threefold in the animal model and in broth culture. Mutation of ssaB, encoding a putative metal transporter, produced a similar effect in culture but reduced in vivo competiveness by >1,000-fold. [3H]palmitate labeling and Western blot analysis confirmed that the lgt mutant failed to acylate lipoproteins, that the lspA mutant had a general defect in lipoprotein cleavage, and that SsaB was processed differently in both mutants. These results indicate that the loss of a single lipoprotein, SsaB, dramatically reduces endocarditis virulence, whereas the loss of most other lipoproteins or of normal lipoprotein processing has no more than a minor effect on virulence.Streptococcus sanguinis is a member of the viridans group of streptococci and is a primary colonizer of teeth (8). The viridans species and, in particular, S. sanguinis (15, 18) are a leading cause of infective endocarditis, a serious infection of the valves or lining of the heart (48). Damage to the heart resulting from rheumatic fever or certain congenital heart defects dramatically increases the risk of developing endocarditis (48, 71). The damage is thought to result in the formation of sterile cardiac “vegetations” composed of platelets and fibrin (48) that can be colonized by certain bacteria during periods of bacteremia. This view is supported by animal studies in which formation of sterile vegetation by cardiac catheterization is required for the efficient establishment of streptococcal endocarditis (17). Prevention of infective endocarditis currently relies upon prophylactic administration of antibiotics prior to dental or other surgical procedures that are likely to produce bacteremia. The growing realization that oral bacteria such as S. sanguinis can enter the bloodstream through routine daily activities such as eating has led the American Heart Association (71) and others (57) to question the value of using antibiotic prophylaxis for dental procedures. Clearly, a better understanding of the bacterial virulence factors that contribute to endocarditis could lead to better preventive measures, such as a vaccine that could potentially afford continuous protection to high-risk patients (71).In a previous study, we used the signature-tagged mutagenesis (STM) technique to search for endocarditis virulence factors of S. sanguinis in a rabbit model (53). This study identified a number of housekeeping enzymes that contribute to endocarditis. Because these proteins are not likely to be surface localized, they hold little promise as vaccine candidates. One class of streptococcal surface proteins that is rich in both virulence factors (4, 7, 25, 33, 38, 60) and promising vaccine candidates (6, 39, 42, 51, 70) is the lipoproteins. Lipoprotein activities that have been suggested to contribute to streptococcal virulence include adhesion (4, 7, 63), posttranslational modification (25, 29, 51), and ATP-binding cassette (ABC)-mediated transport (33, 52, 60). In the last instance, lipoproteins anchored to the cell membrane by their lipid tails appear to serve the same transport function as the periplasmic substrate-binding proteins of gram-negative bacteria (66). STM studies performed with Streptococcus pneumoniae (26, 41, 55) and Streptococcus agalactiae (34) have identified multiple lipoprotein mutants among collections of reduced virulence mutants. In an attempt to determine the cumulative contribution of streptococcal lipoproteins to virulence, some investigators have created mutations in the lgt or lspA genes, encoding lipoprotein-processing enzymes (12, 25, 27, 36). The lgt gene encodes prolipoprotein diacylglyceryl transferase, which catalyzes the transfer of a diacylglycerol lipid unit to a cysteine in the conserved N-terminal “lipobox” of lipoproteins, while lspA encodes the signal peptidase II enzyme that cleaves the signal peptide of the prolipoprotein just prior to the conserved cysteine (59, 65). While mutation of these genes has been shown to be lethal in gram-negative bacteria (21, 73), many gram-positive bacterial species have been shown to tolerate such mutations, often with only minor effects on growth (3, 12, 13, 25, 27, 36, 54). Some of these studies indicated a deleterious effect on the virulence of the lgt (25, 54) or lspA (36) mutation, but others found no effect (12) or an enhancement of virulence (27). It is clear from these and other studies (3, 13) that neither the loss of acylation due to lgt inactivation nor the loss of signal peptidase II-mediated cleavage completely eliminates lipoprotein function, necessitating alternative approaches for assessing the global contribution of lipoproteins to virulence.We have used bioinformatic approaches to identify every putative lipoprotein encoded by S. sanguinis strain SK36. To determine the contribution of these lipoproteins to the endocarditis virulence of S. sanguinis, we have systematically mutagenized each of these genes, as well as the lgt and lspA genes, and evaluated these mutants for virulence by using STM in an animal model. Selected mutants were further examined for virulence in competitive index (CI) assays. A strain with a disrupted ssaB gene, which encodes a putative metal transport protein, was found to exhibit a profound defect in virulence that was far greater than that of any other strain tested, including the lgt or lspA mutant.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号