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 46 毫秒
1.
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   

2.
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   

3.
4.
5.
In this report, we describe the selective cloning of large DNA fragments from magnetotactic metagenomes from various aquatic habitats. This was achieved by a two-step magnetic enrichment which allowed the mass collection of environmental magnetotactic bacteria (MTB) virtually free of nonmagnetic contaminants. Four fosmid libraries were constructed and screened by end sequencing and hybridization analysis using heterologous magnetosome gene probes. A total of 14 fosmids were fully sequenced. We identified and characterized two fosmids, most likely originating from two different alphaproteobacterial strains of MTB that contain several putative operons with homology to the magnetosome island (MAI) of cultivated MTB. This is the first evidence that uncultivated MTB exhibit similar yet differing organizations of the MAI, which may account for the diversity in biomineralization and magnetotaxis observed in MTB from various environments.Magnetotactic bacteria (MTB) synthesize magnetosomes, which are membrane-enclosed organelles comprising crystals of magnetite (Fe3O4) or, less commonly, greigite (Fe3S4) (3) that are aligned in intracellular chains along dedicated cytoskeletal structures (26, 36, 38). Magnetic alignment along the magnetic field lines of the earth facilitates navigation in the stratified environment within freshwater and marine sediments (3, 13). MTB do not form a coherent phylogenetic group, but the trait of magnetotaxis is found in species within different phylogenetic clades, including Alphaproteobacteria, Deltaproteobacteria, Gammaproteobacteria, and the Nitrospira phylum (1, 3, 10, 41). Different species produce magnetosome crystals with a multitude of different morphologies displaying a broad variety of intracellular arrangements, including one, two, or multiple chains (3, 14). The perfectly shaped magnetosome crystals and highly ordered chain structures cannot be synthesized by chemical methods as yet. Therefore, an understanding of the genetic mechanisms controlling magnetosome formation is also of great interest for the inorganic production of advanced magnetic nanomaterials (3, 13, 28).Most genes controlling magnetosome formation and magnetotaxis in Magnetospirillum gryphiswaldense and other freshwater magnetospirilla are clustered within four major operons (mamAB, mamGFDC, mms6, and mamXY) (18, 34, 37, 49) that are part of a large genomic magnetosome island (MAI) (49). It was recently shown that the MAI is also conserved in marine MTB, including the MV-1 magnetotactic vibrio strain and the MC-1 magnetic coccus strain. The homologous genomic regions display similar gene contents and, to a lesser extent, a conserved gene synteny (23). It has been suggested that the MAI was transferred horizontally between different MTB (37). However, the divergence between the MAI regions of strain MV-1, strain MC-1, and the magnetospirilla suggests that the events of horizontal gene transfer (HGT) did not occur very recently.Despite continued efforts by many laboratories, the majority of MTB are still not available in pure culture. In particular, the huge diversity of uncultivated species with respect to different morpho- and phylotypes and, in particular, magnetosome crystal shapes is not nearly fully represented by cultivated species. Thus, understanding of the genetic diversity of the magnetotaxis and magnetosome biosynthetic machinery has to rely on culture-independent techniques such as the metagenomic analysis of environmental MTB (24).It has been demonstrated that single genes and even entire operons can be cloned and functionally expressed from uncultivated soil or marine bacteria by using large insert libraries that provide contiguous sections from single organisms (4, 21, 22). The potential to identify and clone genes for metabolic pathways with relevance for biotechnological applications has already been demonstrated in metagenomic projects, such as the identification of polyketide synthase genes from microbial consortia of marine sponges (25) or other environmental samples (8, 31). The cost of sequencing and the challenges that are associated with the management of vast datasets, however, preclude comprehensive genomic studies of highly complex communities. Consequently, approaches that are based on the analysis of a group of bacteria with reduced species diversity are favored. This requires that the sample material is enriched for the target organisms before DNA preparation, for example, by flow sorting, centrifugation, or other physical enrichment techniques (32, 42) or by focusing on natural communities with reduced species diversity (48).Unlike other uncultivated bacteria, MTB exhibit magnetically directed swimming behavior, which enables their selective enrichment from environmental samples without the need of cultivation (16). This approach was utilized in a number of earlier studies uncovering the morphological and phylogenetic diversity of MTB found in environmental populations (10, 43, 45-47). However, these investigations were confined to PCR-based analysis of 16S rRNA genes, ultrastructural studies, and fluorescence in situ hybridization.In this study we used an improved magnetic collection technique to selectively harvest large numbers of uncultivated MTBs, which allowed the extraction of genomic DNA for the construction of large insert metagenomic libraries from different aquatic habitats. Large parts of the MAI from two uncultivated MTB were identified by hybridization using heterologous magnetosome gene probes and end sequencing. We demonstrate for the first time that uncultivated MTB exhibit a clustered organization of magnetosome genes which resembles that of cultivated species and yet displays variations that may account for the observed diversity in biomineralization and magnetotaxis in MTB from various environments. The levels of similarity between and synteny of magnetosome genes of uncultivated and cultivated MTB provide further evidence for HGT.  相似文献   

6.
7.
8.
Spores of Bacillus subtilis contain a number of small, acid-soluble spore proteins (SASP) which comprise up to 20% of total spore core protein. The multiple α/β-type SASP have been shown to confer resistance to UV radiation, heat, peroxides, and other sporicidal treatments. In this study, SASP-defective mutants of B. subtilis and spores deficient in dacB, a mutation leading to an increased core water content, were used to study the relative contributions of SASP and increased core water content to spore resistance to germicidal 254-nm and simulated environmental UV exposure (280 to 400 nm, 290 to 400 nm, and 320 to 400 nm). Spores of strains carrying mutations in sspA, sspB, and both sspA and sspB (lacking the major SASP-α and/or SASP-β) were significantly more sensitive to 254-nm and all polychromatic UV exposures, whereas the UV resistance of spores of the sspE strain (lacking SASP-γ) was essentially identical to that of the wild type. Spores of the dacB-defective strain were as resistant to 254-nm UV-C radiation as wild-type spores. However, spores of the dacB strain were significantly more sensitive than wild-type spores to environmental UV treatments of >280 nm. Air-dried spores of the dacB mutant strain had a significantly higher water content than air-dried wild-type spores. Our results indicate that α/β-type SASP and decreased spore core water content play an essential role in spore resistance to environmentally relevant UV wavelengths whereas SASP-γ does not.Spores of Bacillus spp. are highly resistant to inactivation by different physical stresses, such as toxic chemicals and biocidal agents, desiccation, pressure and temperature extremes, and high fluences of UV or ionizing radiation (reviewed in references 33, 34, and 48). Under stressful environmental conditions, cells of Bacillus spp. produce endospores that can stay dormant for extended periods. The reason for the high resistance of bacterial spores to environmental extremes lies in the structure of the spore. Spores possess thick layers of highly cross-linked coat proteins, a modified peptidoglycan spore cortex, a low core water content, and abundant intracellular constituents, such as the calcium chelate of dipicolinic acid and α/β-type small, acid-soluble spore proteins (α/β-type SASP), the last two of which protect spore DNA (6, 42, 46, 48, 52). DNA damage accumulated during spore dormancy is also efficiently repaired during spore germination (33, 47, 48). UV-induced DNA photoproducts are repaired by spore photoproduct lyase and nucleotide excision repair, DNA double-strand breaks (DSB) by nonhomologous end joining, and oxidative stress-induced apurinic/apyrimidinic (AP) sites by AP endonucleases and base excision repair (15, 26-29, 34, 43, 53, 57).Monochromatic 254-nm UV radiation has been used as an efficient and cost-effective means of disinfecting surfaces, building air, and drinking water supplies (31). Commonly used test organisms for inactivation studies are bacterial spores, usually spores of Bacillus subtilis, due to their high degree of resistance to various sporicidal treatments, reproducible inactivation response, and safety (1, 8, 19, 31, 48). Depending on the Bacillus species analyzed, spores are 10 to 50 times more resistant than growing cells to 254-nm UV radiation. In addition, most of the laboratory studies of spore inactivation and radiation biology have been performed using monochromatic 254-nm UV radiation (33, 34). Although 254-nm UV-C radiation is a convenient germicidal treatment and relevant to disinfection procedures, results obtained by using 254-nm UV-C are not truly representative of results obtained using UV wavelengths that endospores encounter in their natural environments (34, 42, 50, 51, 59). However, sunlight reaching the Earth''s surface is not monochromatic 254-nm radiation but a mixture of UV, visible, and infrared radiation, with the UV portion spanning approximately 290 to 400 nm (33, 34, 36). Thus, our knowledge of spore UV resistance has been constructed largely using a wavelength of UV radiation not normally reaching the Earth''s surface, even though ample evidence exists that both DNA photochemistry and microbial responses to UV are strongly wavelength dependent (2, 30, 33, 36).Of recent interest in our laboratories has been the exploration of factors that confer on B. subtilis spores resistance to environmentally relevant extreme conditions, particularly solar UV radiation and extreme desiccation (23, 28, 30, 34 36, 48, 52). It has been reported that α/β-type SASP but not SASP-γ play a major role in spore resistance to 254-nm UV-C radiation (20, 21) and to wet heat, dry heat, and oxidizing agents (48). In contrast, increased spore water content was reported to affect B. subtilis spore resistance to moist heat and hydrogen peroxide but not to 254-nm UV-C (12, 40, 48). However, the possible roles of SASP-α, -β, and -γ and core water content in spore resistance to environmentally relevant solar UV wavelengths have not been explored. Therefore, in this study, we have used B. subtilis strains carrying mutations in the sspA, sspB, sspE, sspA and sspB, or dacB gene to investigate the contributions of SASP and increased core water content to the resistance of B. subtilis spores to 254-nm UV-C and environmentally relevant polychromatic UV radiation encountered on Earth''s surface.  相似文献   

9.
10.
11.
We examined the diversity of a marker gene for homoacetogens in two cockroach gut microbial communities. Formyltetrahydrofolate synthetase (FTHFS or fhs) libraries prepared from a wood-feeding cockroach, Cryptocercus punctulatus, were dominated by sequences that affiliated with termite gut treponemes. No spirochete-like sequences were recovered from the omnivorous roach Periplaneta americana, which was dominated by Firmicutes-like sequences.The guts of wood-feeding termites and Cryptocercus punctulatus cockroaches share an unusual pattern of electron flow, as high rates of CO2-reductive acetogenesis typically supplant methanogenesis as the terminal electron sink (2, 3). Past studies have shown that from 10 to 30% of gut acetate produced in environments of termites and wood-feeding cockroaches is microbially generated from CO2 (3, 28), ultimately powering 18 to 26% of the host insect''s own respiratory energy metabolism (25). Nevertheless, most termites emit methane (2), and termite emissions constitute approximately 4% of the global methane budget (27). Cockroaches have been proposed to represent an additional source of note (9). Interestingly, methanogenic termites and cockroaches exhibit increased acetogenesis following addition of exogenous H2 (3, 29). This suggests that these insects are host to a robust population of bacteria that are capable of homoacetogenesis but may be primarily using alternative electron donors (and other substrates and pathways) in vivo.Acetogenic bacteria belonging to two bacterial phyla, Firmicutes and Spirochaetes, have been isolated from the guts of termites (1, 4, 11, 12, 14). Several surveys have targeted and used the gene for formyltetrahydrofolate synthetase (FTHFS), a key gene in the Wood-Ljungdahl pathway of acetogenesis (16), as a potential marker for the pathway (15, 18). For the wood-feeding termites that have been examined, the studies have revealed an abundance of FTHFS sequences that form a coherent phylogenetic cluster, together with genes from homoacetogenic termite gut spirochetes belonging to the genus Treponema (24, 26, 30). This suggests that treponemes may be among the more abundant of the homoacetogens active in these environments.Little is known about the population structure and biology of CO2-reducing, acetogenic bacteria in the guts of either omnivorous or wood-feeding cockroaches. The wood-feeding cockroach Cryptocercus hosts an abundance of flagellate protozoa closely related to those believed to dominate polysaccharide fermentation in the guts of termites (5, 6, 22), suggesting that at least one key environmental niche is filled by similar microbes in both termites and Cryptocercidae. Additionally, Cryptocercidae cockroaches, like termites, house diverse spirochetes and are the site of intense CO2 reduction into acetate (3, 7). Possibly, spirochetes capable of CO2 reduction into acetate are present in the hindguts of cockroaches. However, no evidence has yet been presented for the existence of homoacetogenic treponemes in environments other than the guts of termites, and FTHFS surveys of human (21) or horse (15) fecal matter and bovine rumen samples (20) revealed only Firmicutes-like and other FTHFS alleles that are distinct from those comprising the termite treponeme cluster.Here, by examining FTHFS gene diversity in Cryptocercus punctulatus and Periplaneta americana guts, we endeavored to learn more about the distribution and origins of homoacetogenic treponemes (and their genes) that are found in wood-feeding termites. In particular, we wished to ascertain whether FTHFS genes present in either of the two cockroaches are termite treponeme-like and, if so, whether analysis reveals any obvious signal indicating recent or ancient lateral community transfer events between insect lineages.  相似文献   

12.
An intI-targeted PCR assay was optimized to evaluate the frequency of partial class 2-like integrases relative to putative, environmental IntI elements in clone libraries generated from 17 samples that included various terrestrial, marine, and deep-sea habitats with different exposures to human influence. We identified 169 unique IntI phylotypes (≤98% amino acid identity) relative to themselves and with respect to those previously described. Among these, six variants showed an undescribed, extended, IntI-specific additional domain. A connection between human influence and the dominance of IntI-2-like variants was also observed. IntI phylotypes 80 to 99% identical to class 2 integrases comprised ∼70 to 100% (n = 65 to 87) of the IntI elements detected in samples with a high input of fecal waste, whereas IntI2-like sequences were undetected in undisturbed settings and poorly represented (1 to 10%; n = 40 to 79) in environments with moderate or no recent fecal or anthropogenic impact. Eleven partial IntI2-like sequences lacking the signature ochre 179 codon were found among samples of biosolids and agricultural soil supplemented with swine manure, indicating a wider distribution of potentially functional IntI2 variants than previously reported. To evaluate IntI2 distribution patterns beyond the usual hosts, namely, the Enterobacteriaceae, we coupled PCR assays targeted at intI and 16S rRNA loci to G+C fractionation of total DNA extracted from manured cropland. IntI2-like sequences and 16S rRNA phylotypes related to Firmicutes (Clostridium and Bacillus) and Bacteroidetes (Chitinophaga and Sphingobacterium) dominated a low-G+C fraction (∼40 to 45%), suggesting that these groups could be important IntI2 hosts in manured soil. Moreover, G+G fractionation uncovered an additional set of 36 novel IntI phylotypes (≤98% amino acid identity) undetected in bulk DNA and revealed the prevalence of potentially functional IntI2 variants in the low-G+C fraction.Integrons are genetic modules described in pathogenic and commensal bacteria that confer the ability to capture and express promoterless DNA units, called gene cassettes, which encode a variety of adaptive functions including antibiotic resistance (9, 42, 64). The acquisition of gene cassettes occurs through a site-specific recombination mechanism catalyzed by an integron-encoded integrase (IntI). The integrative recombination reaction occurs primarily between an integron receptor site (attI) and a cassette-associated sequence known as the attC site or 59-base element (11). However, integron integrases are able to recognize and process nonspecific secondary targets as well as attI and attC sites with a high degree of sequence variation (20, 25). This versatility facilitates the exchange of exogenous genes between different integrons through various recombination reactions (attI × attC, attI × attI, and attC × attC) that propel the adaptability and evolution of bacterial genomes (8, 11, 31, 38, 55, 58). Although integrons can be chromosomally encoded, they also may be horizontally transferred via transduction or by transposons associated with conjugative plasmids (42, 61). Three major groups (classes 1 to 3) are known to be associated with laterally transferred elements and highly prevalent in the clinical scene. In most of the cases, these have also been reported to harbor almost exclusively gene cassettes encoding antibiotic resistance functions (42). All together, these traits have led to their designation as “mobile” (9) or “clinical” (22) integrons. Although integrons have been traditionally classified according to the percent identity of the nucleotide or predicted amino acid sequence of their respective intI genes (9, 43, 71), several structural features and differences in abundance patterns have been identified which distinguish classes 1 to 3 (9, 42).Class 1 integrons are the most widely studied variant and are typically linked to replicative Tn21 transposons, which appears to contribute to their extensive distribution (48). A key feature commonly reported within the class 1 module is the presence of a highly conserved 3′ region comprised of a qacEΔ gene and a sul1 gene, which provide protection against quaternary ammonium compounds and sulfa drugs, respectively. In contrast, class 2 integrons are routinely associated with nonreplicative Tn7 transposons, are less frequently detected and, hence, remain an understudied group relative to their class 1 counterparts (42, 48, 65). Even less is known about the class 3 variants, which so far have been described in only three instances (71).Except for the identical IntI2 elements recently reported in Providencia stuartii and Escherichia coli strains isolated from beef cattle sources and the human urinary tract, respectively, all known integrases encoded by class 2 integrons are considered nonfunctional due to the presence of the ochre 179 codon (6, 40, 42). Nevertheless, it has been argued that integrons with truncated class 2 integrases might be implicated in the transfer and high prevalence of antibiotic resistance genes among clinical isolates, possibly via the in trans activities of other functional integrases or the suppression of the stop codon (27). So far, class 2 integrons have been described in association with isolates affiliated to the gamma, beta, and epsilon subdivisions of the Proteobacteria but have been more frequently reported among members of the Gammaproteobacteria group, particularly the Enterobacteriaceae (1, 14, 19, 56, 57). However, most of these studies have focused on easily culturable, aerobic bacteria or those of clinical importance, leading to the exclusion of unculturable or difficult-to-grow commensals that could be inconspicuous but important reservoirs of class 2 elements in the environment. Although the occurrence and quantification of integrons and integron-associated genes by means of molecular, culture-independent methods are being increasingly documented outside the clinical scene (18, 22, 28, 48, 49, 51, 65, 70), the estimates of the extant diversity of the integron platform in nature are still rudimentary. Likewise, further work is needed for the identification of environmental hosts of integrons commonly found in clinical strains without the bias associated with culture techniques (48).In order to provide a comprehensive view of integron integrase variation and prevalence patterns of IntI2 elements in the environment, we PCR amplified partial intI sequences from metagenomic DNA isolated from various terrestrial, marine, and deep-sea habitats exposed to various degrees of anthropogenic or fecal impact. Amplification conditions were optimized to facilitate the assessment of the frequency of IntI2-like sequences relative to that of environmental integron integrases. Additionally, since the guanine-plus-cytosine content of DNA corresponds to taxonomy (68), we coupled G+C fractionation of total DNA (4, 5, 29, 30) with PCR assays targeted at intI and 16S rRNA genes to identify potential, unconventional hosts of class 2 integrons in soil that had received swine manure.  相似文献   

13.
14.
15.
Soil substrate membrane systems allow for microcultivation of fastidious soil bacteria as mixed microbial communities. We isolated established microcolonies from these membranes by using fluorescence viability staining and micromanipulation. This approach facilitated the recovery of diverse, novel isolates, including the recalcitrant bacterium Leifsonia xyli, a plant pathogen that has never been isolated outside the host.The majority of bacterial species have never been recovered in the laboratory (1, 14, 19, 24). In the last decade, novel cultivation approaches have successfully been used to recover “unculturables” from a diverse range of divisions (23, 25, 29). Most strategies have targeted marine environments (4, 23, 25, 32), but soil offers the potential for the investigation of vast numbers of undescribed species (20, 29). Rapid advances have been made toward culturing soil bacteria by reformulating and diluting traditional media, extending incubation times, and using alternative gelling agents (8, 21, 29).The soil substrate membrane system (SSMS) is a diffusion chamber approach that uses extracts from the soil of interest as the growth substrate, thereby mimicking the environment under investigation (12). The SSMS enriches for slow-growing oligophiles, a proportion of which are subsequently capable of growing on complex media (23, 25, 27, 30, 32). However, the SSMS results in mixed microbial communities, with the consequent difficulty in isolation of individual microcolonies for further characterization (10).Micromanipulation has been widely used for the isolation of specific cell morphotypes for downstream applications in molecular diagnostics or proteomics (5, 15). This simple technology offers the opportunity to select established microcolonies of a specific morphotype from the SSMS when combined with fluorescence visualization (3, 11). Here, we have combined the SSMS, fluorescence viability staining, and advanced micromanipulation for targeted isolation of viable, microcolony-forming soil bacteria.  相似文献   

16.
17.
Recently, methicillin-resistant Staphylococcus aureus (MRSA) and methicillin-resistant Staphylococcus pseudintermedius (MRSP) have been increasingly isolated from veterinarians and companion animals. With a view to preventing the spread of MRSA and MRSP, we evaluated the occurrence and molecular characteristics of each in a veterinary college. MRSA and MRSP were isolated from nasal samples from veterinarians, staff members, and veterinary students affiliated with a veterinary hospital. Using stepwise logistic regression, we identified two factors associated with MRSA carriage: (i) contact with an identified animal MRSA case (odds ratio [OR], 6.9; 95% confidence interval [95% CI], 2.2 to 21.6) and (ii) being an employee (OR, 6.2; 95% CI, 2.0 to 19.4). The majority of MRSA isolates obtained from individuals affiliated with the veterinary hospital and dog patients harbored spa type t002 and a type II staphylococcal cassette chromosome mec (SCCmec), similar to the hospital-acquired MRSA isolates in Japan. MRSA isolates harboring spa type t008 and a type IV SCCmec were obtained from one veterinarian on three different sampling occasions and also from dog patients. MRSA carriers can also be a source of MRSA infection in animals. The majority of MRSP isolates (85.2%) carried hybrid SCCmec type II-III, and almost all the remaining MRSP isolates (11.1%) carried SCCmec type V. MRSA and MRSP were also isolated from environmental samples collected from the veterinary hospital (5.1% and 6.4%, respectively). The application of certain disinfection procedures is important for the prevention of nosocomial infection, and MRSA and MRSP infection control strategies should be adopted in veterinary medical practice.Methicillin-resistant Staphylococcus aureus (MRSA) is an important cause of nosocomial infections in human hospitals. The prevalence of hospital-acquired MRSA (HA-MRSA) infection among inpatients in intensive care units (ICUs) continues to increase steadily in Japan. Recently, cases of community-acquired MRSA (CA-MRSA) have been documented in persons without an established risk factor for HA-MRSA infection (14, 32, 36, 49).There has also been an increase in the number of reports of the isolation of MRSA from veterinarians and companion animals (5, 21, 23-26, 28, 31, 34, 38, 44, 50, 51, 53). Values reported for the prevalence of MRSA among veterinary staff include 17.9% in the United Kingdom (21), 10% in Japan (38), 3.9% in Scotland (13), and 3.0% in Denmark (28). Loeffler et al. reported that the prevalence of MRSA among dog patients and healthy dogs owned by veterinary staff members was 8.9% (21). In Japan, an MRSA isolate was detected in only one inpatient dog (3.8%) and could not be detected in any of 31 outpatient dogs (38). In the United States, MRSA isolates were detected in both dog (0.1%) and cat (0.1%) patients (31). The prevalence of MRSA among healthy dogs has been reported to be 0.7% (5). Hanselman et al. suggested that MRSA colonization may be an occupational risk for large-animal veterinarians (12). Recently, Burstiner et al. reported that the frequency of MRSA colonization among companion-animal veterinary personnel was equal to the frequency among large-animal veterinary personnel (6).In addition, other methicillin-resistant coagulase-positive staphylococci (MRCPS), such as methicillin-resistant Staphylococcus pseudintermedius (MRSP) and methicillin-resistant Staphylococcus schleiferi (MRSS), isolated from dogs, cats, and a veterinarian have been reported (11, 31, 38, 40, 52). MRSP isolates have also been detected among inpatient dogs (46.2%) and outpatient dogs (19.4%) in a Japanese veterinary teaching hospital (38). In Canada, however, MRSP and MRSS isolates were detected in only 2.1% and 0.5% of dog patients, respectively (11).Methicillin-resistant staphylococci produce penicillin-binding protein 2′, which reduces their affinity for β-lactam antibiotics. This protein is encoded by the mecA gene (48), which is carried on the staphylococcal cassette chromosome mec (SCCmec). SCCmec is a mobile genetic element characterized by the combination of the mec and ccr complexes (16), and it is classified into subtypes according to differences in the junkyard regions (43). SCCmec typing can be used as a molecular tool (22, 27, 30, 33, 36, 55) for examining the molecular epidemiology of methicillin-resistant staphylococci.In this study, we investigated the occurrence and characteristics of MRCPS isolates in a veterinary hospital in order to establish the transmission route of MRCPS in a veterinary hospital and with a view to preventing the spread of MRCPS infection. In addition, we evaluated the factors associated with MRCPS. Further, as Heller et al. have reported the distribution of MRSA within veterinary hospital environments and suggested the necessity to review cleaning protocols of hospital environments (13), we also attempted to isolate MRCPS from environmental samples collected in a veterinary hospital for an evaluation of MRSA transmission cycle though environmental surfaces in the veterinary hospital.  相似文献   

18.
Adhesive pili on the surface of the serotype M1 Streptococcus pyogenes strain SF370 are composed of a major backbone subunit (Spy0128) and two minor subunits (Spy0125 and Spy0130), joined covalently by a pilin polymerase (Spy0129). Previous studies using recombinant proteins showed that both minor subunits bind to human pharyngeal (Detroit) cells (A. G. Manetti et al., Mol. Microbiol. 64:968-983, 2007), suggesting both may act as pilus-presented adhesins. While confirming these binding properties, studies described here indicate that Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role as a wall linker. Pili were localized predominantly to cell wall fractions of the wild-type S. pyogenes parent strain and a spy0125 deletion mutant. In contrast, they were found almost exclusively in culture supernatants in both spy0130 and srtA deletion mutants, indicating that the housekeeping sortase (SrtA) attaches pili to the cell wall by using Spy0130 as a linker protein. Adhesion assays with antisera specific for individual subunits showed that only anti-rSpy0125 serum inhibited adhesion of wild-type S. pyogenes to human keratinocytes and tonsil epithelium to a significant extent. Spy0125 was localized to the tip of pili, based on a combination of mutant analysis and liquid chromatography-tandem mass spectrometry analysis of purified pili. Assays comparing parent and mutant strains confirmed its role as the adhesin. Unexpectedly, apparent spontaneous cleavage of a labile, proline-rich (8 of 14 residues) sequence separating the N-terminal ∼1/3 and C-terminal ∼2/3 of Spy0125 leads to loss of the N-terminal region, but analysis of internal spy0125 deletion mutants confirmed that this has no significant effect on adhesion.The group A Streptococcus (S. pyogenes) is an exclusively human pathogen that commonly colonizes either the pharynx or skin, where local spread can give rise to various inflammatory conditions such as pharyngitis, tonsillitis, sinusitis, or erysipelas. Although often mild and self-limiting, GAS infections are occasionally very severe and sometimes lead to life-threatening diseases, such as necrotizing fasciitis or streptococcal toxic shock syndrome. A wide variety of cell surface components and extracellular products have been shown or suggested to play important roles in S. pyogenes virulence, including cell surface pili (1, 6, 32). Pili expressed by the serotype M1 S. pyogenes strain SF370 mediate specific adhesion to intact human tonsil epithelia and to primary human keratinocytes, as well as cultured keratinocyte-derived HaCaT cells, but not to Hep-2 or A549 cells (1). They also contribute to adhesion to a human pharyngeal cell line (Detroit cells) and to biofilm formation (29).Over the past 5 years, pili have been discovered on an increasing number of important Gram-positive bacterial pathogens, including Bacillus cereus (4), Bacillus anthracis (4, 5), Corynebacterium diphtheriae (13, 14, 19, 26, 27, 44, 46, 47), Streptococcus agalactiae (7, 23, 38), and Streptococcus pneumoniae (2, 3, 24, 25, 34), as well as S. pyogenes (1, 29, 32). All these species produce pili that are composed of a single major subunit plus either one or two minor subunits. During assembly, the individual subunits are covalently linked to each other via intermolecular isopeptide bonds, catalyzed by specialized membrane-associated transpeptidases that may be described as pilin polymerases (4, 7, 25, 41, 44, 46). These are related to the classical housekeeping sortase (usually, but not always, designated SrtA) that is responsible for anchoring many proteins to Gram-positive bacterial cell walls (30, 31, 33). The C-terminal ends of sortase target proteins include a cell wall sorting (CWS) motif consisting, in most cases, of Leu-Pro-X-Thr-Gly (LPXTG, where X can be any amino acid) (11, 40). Sortases cleave this substrate between the Thr and Gly residues and produce an intermolecular isopeptide bond linking the Thr to a free amino group provided by a specific target. In attaching proteins to the cell wall, the target amino group is provided by the lipid II peptidoglycan precursor (30, 36, 40). In joining pilus subunits, the target is the ɛ-amino group in the side chain of a specific Lys residue in the second subunit (14, 18, 19). Current models of pilus biogenesis envisage repeated transpeptidation reactions adding additional subunits to the base of the growing pilus, until the terminal subunit is eventually linked covalently via an intermolecular isopeptide bond to the cell wall (28, 41, 45).The major subunit (sometimes called the backbone or shaft subunit) extends along the length of the pilus and appears to play a structural role, while minor subunits have been detected either at the tip, the base, and/or at occasional intervals along the shaft, depending on the species (4, 23, 24, 32, 47). In S. pneumoniae and S. agalactiae one of the minor subunits acts as an adhesin, while the second appears to act as a linker between the base of the assembled pilus and the cell wall (7, 15, 22, 34, 35). It was originally suggested that both minor subunits of C. diphtheriae pili could act as adhesins (27). However, recent data showed one of these has a wall linker role (26, 44) and may therefore not function as an adhesin.S. pyogenes strain SF370 pili are composed of a major (backbone) subunit, termed Spy0128, plus two minor subunits, called Spy0125 and Spy0130 (1, 32). All three are required for efficient adhesion to target cells (1). Studies employing purified recombinant proteins have shown that both of the minor subunits, but not the major subunit, bind to Detroit cells (29), suggesting both might act as pilus-presented adhesins. Here we report studies employing a combination of recombinant proteins, specific antisera, and allelic replacement mutants which show that only Spy0125 is the pilus-presented adhesin and that Spy0130 has a distinct role in linking pili to the cell wall.  相似文献   

19.
Magnetosomes are unique bacterial organelles comprising membrane-enveloped magnetic crystals produced by magnetotactic bacteria. Because of several desirable chemical and physical properties, magnetosomes would be ideal scaffolds on which to display highly complicated biological complexes artificially. As a model experiment for the functional expression of a multisubunit complex on magnetosomes, we examined the display of a chimeric bacterial RNase P enzyme composed of the protein subunit (C5) of Escherichia coli RNase P and the endogenous RNA subunit by expressing a translational fusion of C5 with MamC, a known magnetosome protein, in the magnetotactic bacterium Magnetospirillum gryphiswaldense. As intended, the purified C5 fusion magnetosomes, but not wild-type magnetosomes, showed apparent RNase P activity and the association of a typical bacterial RNase P RNA. Our results demonstrate for the first time that magnetosomes can be employed as scaffolds for the display of multisubunit complexes.Magnetosomes are unique organelles comprising membrane-enveloped magnetic crystals of iron minerals (Fe3O4 or Fe3S4) produced by magnetotactic bacteria (1, 11). The bacteria employ magnetosomes to sense the environmental magnetic field, probably in order to recognize their favorite environments. Compared with chemically or physically synthesized magnetic nanoparticles, magnetosomes have a variety of desirable features, including their genetically controlled uniform size and morphology, characteristic crystal habits, and their coverage by a biological membrane that can be addressed by functionalization (1, 4, 11). Based on these features, magnetosomes would be ideal scaffolds on which to display biological molecules artificially.Until now, several heterologous target proteins have been examined for artificial display on magnetosomes (1, 11). For example, reporter proteins such as luciferase and green fluorescent protein were employed to analyze the targeting, expression, and stability of chimeric proteins displayed on magnetosomes (14, 18, 23, 30, 41). For more-practical applications, general antibody-binding proteins (protein A and protein G) were displayed to capture desired antibodies (16, 17, 25, 33, 34, 37, 41). Such antibody-captured magnetosomes are applicable for the magnetic separation of target molecules and cells. Displays of G protein-coupled receptors (the D1 dopamine receptor and the ligand binding domain of the estrogen receptor) were also examined for screening of drugs targeting these receptors (38, 39, 40).There are two major strategies for the construction of functionalized magnetosomes: subsequent chemical modifications of purified magnetosomes (3) and in vivo expression of modified magnetosome proteins (1, 19). The latter approach is confined to biological molecules that can be expressed as a genetic fusion with a magnetosome protein inside a magnetotactic bacterium. By this approach, the target-displaying magnetosomes can be constructed inside cells or under physiological conditions in the presence of a variety of chaperons, are recoverable under mild conditions employing a magnetic field, and provide control by genetic means. Thus, the approach is highly promising for the display of a naïve target such as a multisubunit complex. To date, however, experimental evidence that magnetosomes can be employed as scaffolds for the display of such targets is still lacking. In order to demonstrate this potential of magnetosomes, here, we examined the display of a holoenzyme of bacterial RNase P, one of the simplest complexes composed of a single RNA and a single protein subunit (10, 12), by expressing a fusion of a protein component of the RNase P and a magnetosome membrane protein.  相似文献   

20.
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   

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