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.
Lactobacillus casei Zhang is a new probiotic bacterium isolated from koumiss collected in Inner Mongolia, China. Here, we report the main genome features of L. casei Zhang and the identification of several predicted proteins implicated in interactions with the host.Koumiss, a traditional drink made from mare''s milk by nomadic peoples in China and Mongolia, is believed to be beneficial in the cure of digestive diseases and a wide range of chronic diseases, including tuberculosis, bronchitis, and anemia (3). Lactobacillus casei Zhang is a novel probiotic strain identified by screening of lactic acid bacteria isolated from koumiss samples collected in Inner Mongolia, China, and exhibits high-level resistance to acid and bile stresses, as well as antibacterial, antioxidative, and immunomodulatory properties (6, 7, 11).A whole-genome shotgun strategy was used for sequencing of the genome of L. casei Zhang. pUC18 plasmid libraries with insertions of 1.5 to 2.5 kb and 4 to 6 kb were constructed (8). Gaps were closed by sequencing of PCR products. Base calling and sequence assembly were carried out using the Phred/Phrap/Consed software package (http://www.phrap.org/), and reads giving a total of 6.2-fold coverage were assembled with an error rate of <0.0001. Gene prediction and annotation were performed as described previously (10).The complete genome of L. casei Zhang consists of a 2,861,848-bp circular chromosome and a 36-kb plasmid. The average G+C content of the chromosome is 46.5%, while the plasmid has a lower G+C content (10). The L. casei Zhang genome contains 2,804 predicted coding sequences (CDSs), five rRNA operons, and 59 tRNAs. No functional prophages were identified, except for the previously described prophage remnant (9). Genes for 41 transposases were found in the genome, and this number was much lower than (only about 30%) those of transposase genes in L. casei ATCC 334 and BL23 (1, 4), suggesting that insertion element (IS)-mediated genome diversification was less frequent in L. casei Zhang.Comparative genome analysis revealed that the number of phosphotransferase system (PTS)-related proteins varied significantly in L. casei strains. Almost twice as many PTS components were found in L. casei Zhang and BL23 as in L. casei ATCC 334. In contrast to L. casei ATCC 334, L. casei Zhang was found to have 33 PTS components consisting of 11 complete substrate-specific enzyme II (EII) complexes encoded by six genomic islands. The G+C contents of the six islands ranged from 41 to 47%, similar to the average G+C content of the L. casei Zhang genome. In addition, most of the EII components in L. casei Zhang (81 of 96) were conserved in L. casei BL23, suggesting that a large-scale loss of PTSs occurred in L. casei ATCC 334 during its evolution. Conspicuous redundancy of chromosome-encoded PTSs in L. casei Zhang may offer benefits in the transport and use of a large panel of carbon sources.Genes encoding five putative mucus-binding proteins (LCAZH_0407, LCAZH_2292, LCAZH_2478, LCAZH_2398, and LCAZH_1427) and a cluster of genes encoding bacteriocin biosynthetic proteins (LCAZH_2341 to LCAZH_2348) nearly identical to those in L. casei ATCC 334 and BL23 were identified in L. casei Zhang and may provide this bacterium with some competitive advantages in the gastrointestinal environment (2, 5).In conclusion, the comparative analysis revealed the flexibility of L. casei Zhang in sugar utilization. In addition, some possible hints for its interactions with the host were identified. This genome sequence will be the basis for systematic studies into the mechanism for the probiotic properties of L. casei Zhang.  相似文献   

2.
3.
Multilocus sequence typing with nine selected genes is shown to be a promising new tool for accurate identifications of Brevibacteriaceae at the species level. A developed microarray also allows intraspecific diversity investigations of Brevibacterium aurantiacum showing that 13% to 15% of the genes of strain ATCC 9174 were absent or divergent in strain BL2 or ATCC 9175.Brevibacteriaceae play a major part in the cheese smear community (6, 11). The classification and typing of cheese-related Brevibacteriaceae have been based mainly on molecular methods such as amplified ribosomal DNA restriction enzyme analysis, pulsed-field gel electrophoresis, and ribotyping (8, 10, 12). Recently, the original Brevibacterium linens group was split into two species on the basis of their physiological and biochemical characteristics, the sugar and polyol composition of their teichoic acids, and their 16S rRNA sequence and DNA-DNA hybridization levels. One species remains B. linens and is represented by type strain ATCC 9172. The other, represented by type strain ATCC 9175, has been renamed Brevibacterium aurantiacum. Regarding this new classification, the taxonomic position of cheese-related isolates has to be revisited and potential relationships between phylogenetic affiliation and the potential occurrence of given metabolic characteristics redefined (7). The unfinished genome sequence of B. aurantiacum ATCC 9174 has recently been released by the Joint Genome Institute (http://genome.jgi-psf.org/draft_microbes/breli/breli.home.html). The development of focused phylogenetic approaches using multiple markers in conjunction with whole-genome screening techniques such as comparative genomic hybridization (CGH) has proven to be useful for the detailed characterization of pathogenic species, including food pathogens (3, 5, 9). However, only a few technological species have been investigated at an intraspecies level (2). Our intention was thus to develop modern tools to facilitate the typing of strains of technological interest, for which Brevibacteriaceae could be used as a case study.  相似文献   

4.
Cell surface polysaccharides have an established role as virulence factors in human bacterial pathogens. Less documented are the biosynthesis and biological functions of surface polysaccharides in beneficial bacteria. We identified a gene cluster that encodes the enzymes and regulatory and transporter proteins for the different steps in the biosynthesis of extracellular polysaccharides (EPS) of the well-documented probiotic strain Lactobacillus rhamnosus GG. Subsequent mutation of the welE gene, encoding the priming glycosyltransferase within this cluster, and comparative phenotypic analyses of wild-type versus mutant strains confirmed the specific function of this gene cluster in the biosynthesis of high-molecular-weight, galactose-rich heteropolymeric EPS molecules. The phenotypic analyses included monomer composition determination, estimation of the polymer length of the isolated EPS molecules, and single-molecule force spectroscopy of the surface polysaccharides. Further characterization of the welE mutant also showed that deprivation of these long, galactose-rich EPS molecules results in an increased adherence and biofilm formation capacity of L. rhamnosus GG, possibly because of less shielding of adhesins such as fimbria-like structures.Bacterial surface polysaccharides are considered to be key macromolecules in determining microbe-host interactions, as they display a high degree of variety and diversity among bacterial species in terms of composition, monomer linkages, branching degree, polymer size, production level, etc. (24, 46). Since most bacteria contain more than one type of surface polysaccharides, such as lipopolysaccharides (O antigens), capsular polysaccharides (CPS), exopolysaccharides (EPS), and/or glycan chains as part of glycoproteins, the elucidation of their exact role is complex. Nevertheless, surface polysaccharides are now known to exert important functions at several stages during pathogenesis, including tissue adherence, biofilm formation, and evasion of host defenses such as phagocytosis (9, 24, 33). In addition to their role in pathogens, an important biological role for CPS and glycoproteins has also recently been shown in colonization of the gut by bacteria of the genus Bacteroides (10, 34).Conversely, the role of surface polysaccharides in probiotic-host interactions has not yet been studied in great detail. A probiotic bacterium is defined as “a live microorganism that, when administered or ingested in adequate amounts, confers a health benefit on the host” (18). Members of the genus Lactobacillus are commonly studied for their health-promoting capacities (26, 31, 37). As polysaccharides display a high diversity among lactobacilli, they are thought to be involved in determining strain-specific properties important for probiotic action, such as adhesion, stress resistance, and interactions with specific receptors and effectors of the host defense system (13, 56). Moreover, these EPS molecules are of interest in the dairy industry for conferring textural and rheological properties to fermented products such as yogurt and soft cheese (56). Nevertheless, detailed genetic and functional studies of EPS molecules of lactobacilli are currently limited (26, 56).Lactobacillus rhamnosus GG (ATCC 53103) is one of the probiotic strains with the largest number of proven health benefits (15). Several clinical trials have reported that L. rhamnosus GG can prevent and relieve certain types of diarrhea (22) and atopic disease (25) and reduce inflammation in some milder states of inflammatory bowel diseases (60). However, the cell surface factors or specific characteristics of L. rhamnosus GG that underlie these health benefits are largely unknown.We recently showed by single-molecule force spectroscopy (SMFS) with specific lectin tips that the cell surface of L. rhamnosus GG wild-type cells contains two major types of cell wall-associated polysaccharides (CW-PS) (21). The longest and most abundant polysaccharides are galactose-rich and seem to correspond with the EPS molecules of L. rhamnosus GG, which were previously structurally identified by Landersjö et al. (27) using nuclear magnetic resonance spectroscopy. Additionally, shorter, yet-uncharacterized glucose-rich polysaccharides are present on the L. rhamnosus GG surface (21). In the current study, we describe the identification and annotation of the L. rhamnosus GG gene cluster that encodes the enzymes and transporter and regulatory proteins involved in the biosynthesis of long, galactose-rich EPS molecules. This was experimentally confirmed by the construction of a knockout mutant of the corresponding priming glycosyltransferase and subsequent characterization of the surface polysaccharides of wild-type and mutant strains. We also studied the specific role of these EPS molecules in adherence to mucus and gut epithelial cells and in biofilm formation by L. rhamnosus GG.  相似文献   

5.
Telomeres are structural and functional chromosome regions that are essential for the cell cycle to proceed normally. They are, however, difficult to map genetically and to identify in genome-wide sequence programs because of their structure and repetitive nature. We studied the telomeric and subtelomeric organization in the basidiomycete Pleurotus ostreatus using a combination of molecular and bioinformatics tools that permitted us to determine 19 out of the 22 telomeres expected in this fungus. The telomeric repeating unit in P. ostreatus is TTAGGG, and the numbers of repetitions of this unit range between 25 and 150. The mapping of the telomere restriction fragments to linkage groups 6 and 7 revealed polymorphisms compatible with those observed by pulsed field gel electrophoresis separation of the corresponding chromosomes. The subtelomeric regions in Pleurotus contain genes similar to those described in other eukaryotic systems. The presence of a cluster of laccase genes in chromosome 6 and a bipartite structure containing a Het-related protein and an alcohol dehydrogenase are especially relevant; this bipartite structure is characteristic of the Pezizomycotina fungi Neurospora crassa and Aspergillus terreus. As far as we know, this is the first report describing the presence of such structures in basidiomycetes and the location of a laccase gene cluster in the subtelomeric region, where, among others, species-specific genes allowing the organism to adapt rapidly to the environment usually map.Pleurotus ostreatus (Jacq.: Fr) Kumm. (Dikarya, Basidiomycota, Agaricomycotina, Agaricales) (52) is an active lignin degrader that lives as a saprophyte on dead or decaying wood. P. ostreatus (oyster mushroom) has been industrially cultivated for food production because of its flavor and its nutritional (49) and health-stimulating (8) properties. In addition, it produces various secondary metabolites of medical interest (33). P. ostreatus ligninolytic activity and enzymes have been used in the bioconversion of agricultural wastes (1); in the biodegradation of organopollutants, xenobiotics, and industrial contaminants (12); and in paper pulp bleaching (65), among other applications (10).The whole genome sequence of P. ostreatus is currently being assembled at the Joint Genome Institute (California). P. ostreatus is the first edible and the second lignin-degrading basidiomycete to be sequenced. The sequences of other basidiomycetes, such as Phanerochaete chrysosporium (48), Cryptococcus neoformans (44), Ustilago maydis (38), and Laccaria bicolor (47) have been published, and others (Postia placenta, Heterobasidion annosum, Agaricus bisporus, Serpula lacrymans, etc.) are in progress.Telomeres are the protective DNA-protein complexes found at chromosome termini (6, 13, 76). In most eukaryotes, telomeric DNA consists of tandem arrays of 5- to 8-bp direct repeats where specific telomere-capping proteins bind to ensure chromosomal-end integrity. Telomeres are essential for genome stability, and their shortening (attrition) can lead to chromosome instability, replicative senescence, and apoptosis (43), while their loss causes activation of DNA damage responses (45, 66), cell cycle arrest (28), and chromosome fusions, such as nonreciprocal translocations (7, 32). Moreover, high recombination rates are frequent near telomeres (50).Telomeres and subtelomeric regions are usually gene reservoirs that permit organisms to quickly adapt to new ecological niches (60). Two types of genes participate in this adaptive process: species-specific (18) and contingency genes (5). Species-specific genes are shorter than the core genes of the genomes in which they are present, contain fewer exons, exhibit a subtelomeric bias, and arise by duplication, diversification, and differential gene loss. The avirulence genes of some phytopathogenic fungi are contingency genes that appear near telomeres (15). Furthermore, it has recently been found in Fusarium species that pathogenicity-related genes cooccur with telomeric regions. In this case, chromosomal rearrangements (fusions) have maintained these structures. The Fusarium graminearum genome revealed a link between localized polymorphism and pathogen specialization (11). Among the genes frequently found in subtelomeric regions in Magnaporthe oryzae and Aspergillus sp., the presence of transposons, telomere-linked RecQ helicases, clusters of secondary-metabolite genes, cytochrome oxidases, hydrolases, molecular transporters, and genes encoding secreted proteins, among others, has been reported (18, 56).RecQ helicases are highly conserved in evolution and are required for genome stability. Genes coding for these enzymes have been described in prokaryotes and eukaryotes (4, 9, 39, 71). There are a minimum of five RecQ helicase-like genes in humans, and three of them (BLM, WRN, and RECQL4) are mutated in the Bloom, Werner, and Rothmund-Thomson recessive autosomal syndromes, which exhibit genomic instability leading ultimately to cancer (9). Fungal RecQ helicase-like genes have been previously found associated with chromosome ends (23, 35, 56, 61).In genome-sequencing projects, telomeres and subtelomeric regions are rarely present or assembled because of problems derived from their repetitive nature; therefore, it is necessary to perform direct cloning of the subtelomeric regions. The rice pathogen M. oryzae (56) is one of the few fungi with telomeric and subtelomeric regions characterized. Telomere-associated markers provide an accurate assessment of linkage group (LG) completeness and a better estimate of genetic size and help in establishing the synteny of LGs, especially in those organisms for which genetic-linkage maps are not available (34). Moreover, these markers inform us about the genome organization and the occurrence of species-specific and contingency genes (5, 18), as well as about the chromosome rearrangements that could have occurred in the evolution of the genome.In this work, we mapped and studied the telomeric and subtelomeric regions of most of the P. ostreatus chromosomes, and we describe the main genes present in them. The study was carried out with a combination of genetic, molecular, and bioinformatics tools. The results obtained show the high complexity of these regions and confirm the presence of RecQ helicase-like, heterokaryotic incompatibility (het), and short-chain dehydrogenase genes that have also been found in other fungi. In addition, a laccase gene cluster is described for the first time in the subtelomeric region of chromosome 6. This study is the first step toward analyzing the effects that the subtelomeric positions of some fungal-species-specific genes (such as the laccases can be in white rot lignocellulolytic fungi) could have in the adaptation to new growing substrates and in the generation of large families of apparently redundant elements.  相似文献   

6.
We have sequenced the genome and identified the structural proteins and lipids of the novel membrane-containing, icosahedral virus P23-77 of Thermus thermophilus. P23-77 has an ∼17-kb circular double-stranded DNA genome, which was annotated to contain 37 putative genes. Virions were subjected to dissociation analysis, and five protein species were shown to associate with the internal viral membrane, while three were constituents of the protein capsid. Analysis of the bacteriophage genome revealed it to be evolutionarily related to another Thermus phage (IN93), archaeal Halobacterium plasmid (pHH205), a genetic element integrated into Haloarcula genome (designated here as IHP for integrated Haloarcula provirus), and the Haloarcula virus SH1. These genetic elements share two major capsid proteins and a putative packaging ATPase. The ATPase is similar with the ATPases found in the PRD1-type viruses, thus providing an evolutionary link to these viruses and furthering our knowledge on the origin of viruses.Three-dimensional structures of the major capsid proteins, as well as the architecture of the virion and the sequence similarity of putative genome packaging ATPases, have revealed unexpected evolutionary connection between virus families. Viruses infecting hosts residing in different domains of life (Bacteria, Archaea, and Eukarya) share common structural elements and possibly also ways to package the viral genome (8, 13, 41). It has been proposed that the set of genes responsible for virion assembly is a hallmark of the virus and is designated as the innate viral “self,” which may retain its identity through evolutionary times (5). Based on this, it is proposed that viruses can be classified into lineages that span the different domains of life. Therefore, the studies of new virus isolates might provide insights into the events that led to the origin of viruses and maybe even the origin of life itself (34, 40). However, viruses are known to be genetic mosaics (28), and these structural lineages therefore do not reflect the evolutionary history of all genes in a given virus. For example, the genome replication strategies vary significantly even in the currently established lineages (41) and, consequently, a structural approach does not point out to a specific form of replication in the ancestor. Nevertheless, as the proposal for a viral self is driven from information on viral structures and pathways of genome encapsidation, the ancestral form of the self was likely to be composed of a protective coat and the necessary mechanisms to incorporate the genetic material within the coat.Viruses structurally related to bacteriophage PRD1, a phage infecting gram-negative bacteria, have been identified in all three domains of life, and the lineage hypothesis was first proposed based on structural information on such viruses. Initially, PRD1 and human adenovirus were proposed to originate from a common ancestor mainly due to the same capsid organization (T=25) and the major coat protein topology, the trimeric double β-barrel fold (12). In addition, these viruses share a common vertex organization and replication mechanism (20, 31, 53, 63). PRD1 is an icosahedral virus with an inner membrane, whereas adenovirus lacks the membrane. Later, many viruses with similar double β-barrel fold in the major coat protein have been discovered and included to this viral lineage. For example, the fold is present in Paramecium bursaria Chlorella virus 1 (56) of algae, Bam35 (45) of gram-positive bacteria, PM2 (2) of gram-negative marine bacteria, and Sulfolobus turreted icosahedral virus (STIV) (38) of an archaeal host. Moreover, genomic analyses have revealed a common set of genes in a number of nucleocytoplasmic large DNA viruses. Chilo iridescent virus and African swine fever virus 1 are related to Paramecium bursaria Chlorella virus 1 and most probably share structural similarity to PRD1-type viruses (13, 30, 31, 68). The largest known viruses, represented by mimivirus and poxvirus, may also belong to this lineage (29, 77). Two euryarchaeal proviruses, TKV4 and MVV, are also proposed to belong to this lineage based on bioinformatic searches (42). The proposed PRD1-related viruses share the same basic architectural principles despite major differences in the host organisms and particle and genome sizes (1, 2, 38, 56). PM2, for example, has a genome of only 10 kbp, whereas mimivirus (infecting Acanthamoeba polyphaga) double-stranded DNA (dsDNA) genome is 1.2 Mbp in size (59).How many virion structure-based lineages might there be? This obviously relates to the number of protein folds that have the properties needed to make viral capsids. It has been noted that, in addition to PRD1-type viruses, at least tailed bacterial and archaeal viruses, as well as herpesviruses, share the same coat protein fold. Also, certain dsRNA viruses seem to have structural and functional similarities, although their hosts include bacteria and yeasts, as well as plants and animals (6, 18, 19, 27, 55, 60, 74). Obviously, many structural principles to build a virus capsid exist, and it has been suggested that especially geothermally heated environments have preserved many of the anciently formed virus morphotypes (35).Thermophilic dsDNA bacteriophage P23-77 was isolated from an alkaline hot spring in New Zealand on Thermus thermophilus (17) ATCC 33923 (deposited as Thermus flavus). P23-77 was shown to have an icosahedral capsid and possibly an internal membrane but no tail (81). Previously, another Thermus virus, IN93, with a similar morphology has been described (50). IN93 was inducible from a lysogenic strain of Thermus aquaticus TZ2, which was isolated from hot spring soil in Japan. Recently, P23-77 was characterized in more detail (33). It has an icosahedral protein coat, organized in a T=28 capsid lattice (21). The presence of an internal membrane was confirmed, and lipids were shown to be constituents of the virion. Ten structural proteins were identified, with apparent molecular masses ranging from 8 to 35 kDa. Two major protein species with molecular masses of 20 and 35 kDa were proposed to make the capsomers, one forming the hexagonal building blocks and the other the two towers that decorate the capsomer bases (33). Surprisingly, P23-77 is structurally closest to the haloarchaeal virus SH1, which is the only other example of a T=28 virion architecture (32, 33). In both cases it was proposed that the capsomers are made of six single β-barrels opposing the situation with the other structurally related viruses where the hexagonal capsomers are made of three double β-barrel coat protein monomers (8).In the present study we analyze the dsDNA genome of P23-77. Viral membrane proteins and those associated with the capsid were identified by virion dissociation studies. The protein chemistry data and genome annotation are consistent with the results of the disruption studies. A detailed analysis of the lipid composition of P23-77 and its T. thermophilus host was carried out. The data collected here reveal additional challenges in attempts to generate viral lineages based on the structural and architectural properties of the virion.  相似文献   

7.
Lactobacillus crispatus is a common member of the beneficial microbiota present in the vertebrate gastrointestinal and human genitourinary tracts. Here, we report the genome sequence of L. crispatus ST1, a chicken isolate displaying strong adherence to vaginal epithelial cells.Lactobacillus crispatus can persist in the vertebrate gastrointestinal tract and is among the most prevalent species of the Lactobacillus-dominated human vaginal microbiota (2, 9, 13, 14). It belongs to the so-called acidophilus group (3), which has attracted interest because some of its species are important factors in the production of fermented foods (12) and some can, at least transiently, colonize the human host (2, 9, 13, 14). Moreover, some specific strains, mainly L. acidophilus NCFM and L. johnsonii NCC 533, have received prominence as intestinal-health-promoting microbes (4). Although the genomes of seven members of the acidophilus complex have been sequenced to date (12), the genome sequences of L. crispatus and other predominant lactobacillar species in the urogenital flora have mostly remained obscure. Vaginal lactobacilli can have an important role in controlling the health of the host (2, 14). They can, for example, positively influence and stabilize the host''s vaginal microbiota via the production of compounds that are acidic or exert a direct inhibiting action toward pathogenic bacteria (2, 14). In addition to the antimicrobial compounds, the competitive exclusion of pathogens is another mechanism by which the host''s microbiota can be balanced (2). L. crispatus ST1 was originally isolated from the crop of a chicken, and PCR profiling of L. crispatus isolates has verified it to be an abundant colonizer of the chicken crop (6, 8). It also displays a strong protein-dependent adhesion to the epithelial cells of the human vagina and has been shown to inhibit the adhesion of avian pathogenic Escherichia coli (6, 7).The genome was sequenced (18× coverage) using a 454 pyrosequencer with GS FLX chemistry (Roche). The contig order was confirmed and gaps were filled by sequencing PCR fragments from the genomic DNA template using ABI 3730 and Big Dye chemistry (Applied Biosystems). Genomic data were processed using the Staden Package (11) and gsAssembler (Roche). Coding sequences (CDSs) were predicted using Glimmer3 (5) followed by manual curation of the start sites. The remaining intergenic regions were reanalyzed for missed CDSs by using BlastX (1). Annotation transfer was performed based on a BlastP search, followed by Blannotator analysis using default settings (http://ekhidna.biocenter.helsinki.fi/poxo/blannotator) and manual verification. Orthologous groups between the different lactobacillar proteomes were identified using OrthoMCL (10).The genome of L. crispatus ST1 consists of a single circular chromosome 2.04 Mbp in size, with an overall G+C content of 37%, without any plasmids. There are 64 tRNA genes, 4 rRNA operons, and 2 CRISPR loci. Out of the 2,024 predicted CDSs, a putative function was assigned to 77%, whereas 10% of the CDSs were annotated as conserved and 13% as novel. Based on the orthologous grouping, 302 (15%) of the CDSs encoded by ST1 have no detectable homologs in any of the Lactobacillus proteomes published to date.  相似文献   

8.
The main siderophores produced by streptomycetes are desferrioxamines. Here we show that Streptomyces sp. ATCC 700974 and several Streptomyces griseus strains, in addition, synthesize a hitherto unknown siderophore with a catechol-peptide structure, named griseobactin. The production is repressed by iron. We sequenced a 26-kb DNA region comprising a siderophore biosynthetic gene cluster encoding proteins similar to DhbABCEFG, which are involved in the biosynthesis of 2,3-dihydroxybenzoate (DHBA) and in the incorporation of DHBA into siderophores via a nonribosomal peptide synthetase. Adjacent to the biosynthesis genes are genes that encode proteins for the secretion, uptake, and degradation of siderophores. To correlate the gene cluster with griseobactin synthesis, the dhb genes in ATCC 700974 were disrupted. The resulting mutants no longer synthesized DHBA and griseobactin; production of both was restored by complementation with the dhb genes. Heterologous expression of the dhb genes or of the entire griseobactin biosynthesis gene cluster in the catechol-negative strain Streptomyces lividans TK23 resulted in the synthesis and secretion of DHBA or griseobactin, respectively, suggesting that these genes are sufficient for DHBA and griseobactin biosynthesis. Griseobactin was purified and characterized; its structure is consistent with a cyclic and, to a lesser extent, linear form of the trimeric ester of 2,3-dihydroxybenzoyl-arginyl-threonine complexed with aluminum under iron-limiting conditions. This is the first report identifying the gene cluster for the biosynthesis of DHBA and a catechol siderophore in Streptomyces.Iron is an essential element for the growth and proliferation of nearly all microorganisms. In the presence of oxygen, the soluble ferrous iron is readily oxidized to its ferric form, which exists predominantly as a highly insoluble hydroxide complex at neutral pH. To overcome iron limitation, many bacteria synthesize and secrete low-molecular-weight, high-affinity ferric iron chelators, called siderophores (38, 53). Following the chelation of Fe3+ in the medium, the iron-siderophore complex is actively taken up by its cognate ABC transport system, and Fe3+ is subsequently released by reduction to Fe2+ and/or by hydrolysis of the siderophore (28, 32, 36). The three main classes of siderophores contain catecholates, hydroxamates, or (α-hydroxy-)carboxylates as iron-coordinating ligands, but mixed siderophores and siderophores containing other functional groups, such as diphenolates, imidazoles, and thiazolines, have also been found (16, 38).Siderophores containing peptide moieties are synthesized by proteins belonging to the nonribosomal peptide synthetase (NRPS) family (16, 38). These multimodular enzymes function as enzymatic assembly lines in which the order of the modules usually determines the order of the amino acids incorporated into the peptide (19, 34). Each module contains the complete information for an elongation step combining the catalytic functions for the activation of the amino acid by the adenylation (A) domain, the tethering of the corresponding adenylate to the terminal thiol of the enzyme-bound 4′-phosphopantetheinyl (4′-PP) cofactor by the peptidyl carrier protein (PCP) domain, and the formation of the peptide bond by the condensation (C) domain (26, 34, 52). At the end, the product is released by the C-terminal thioesterase (TE) domain by hydrolysis or by cyclization via intramolecular condensation. Each adenylation domain recognizes a specific amino acid, and its substrate specificity can be predicted by its sequence. An NRPS specificity-conferring code consisting of 10 nonadjacent amino acid residues in the A domain has been proposed (49). Exceptions to the “colinearity-rule” (19) have been discovered. For example, in the biosynthesis of the siderophores enterobactin and bacillibactin, all the modules in the NRPS are used iteratively, and the TE domain stitches the chains together into a cyclic product (35, 45). Enterobactin is the trilactone of 2,3-dihydroxybenzoyl-serine, and bacillibactin is the lactone of 2,3-dihydroxybenzoyl-glycyl-threonine.The typical siderophores produced by streptomycetes are desferrioxamines (24), and the genes encoding the enzymes for their biosynthesis have been identified (5). Recently, structurally different siderophores have been reported to be coproduced with desferrioxamines in some species, e.g., coelichelin in Streptomyces coelicolor (9, 30) and enterobactin in Streptomyces tendae (18). The genes encoding the proteins for the biosynthesis of enterobactin in S. tendae remain unknown.Here we describe the gene cluster for the biosynthesis of a new siderophore, named griseobactin, produced by Streptomyces sp. strain ATCC 700974 and some strains of Streptomyces griseus. By sequencing two cosmids isolated from a Streptomyces sp. strain ATCC 700974 genomic library, we assigned the encoded proteins to enzymes that convert chorismate to 2,3-dihydroxybenzoate (DHBA), and to proteins involved in nonribosomal peptide biosynthesis and in the export, uptake, and utilization of siderophores. Knockout mutagenesis and heterologous expression confirmed the requirement of this gene cluster for the biosynthesis of griseobactin. This is the first report on the identification of the genes responsible for DHBA and catechol siderophore biosynthesis in Streptomyces.  相似文献   

9.
Staphylococcus saprophyticus is a common cause of uncomplicated urinary tract infections in women. S. saprophyticus strain ATCC 15305 carries two staphylococcal cassette chromosome genetic elements, SCC15305RM and SCC15305cap. The SCC15305cap element carries 13 open reading frames (ORFs) involved in capsular polysaccharide (CP) biosynthesis, and its G+C content (26.7%) is lower than the average G+C content (33.2%) for the whole genome. S. saprophyticus strain ATCC 15305 capD, capL, and capK (capDSsp, capLSsp, and capKSsp) are homologous to genes encoding UDP-FucNAc biosynthesis, and gtaB and capISsp show homology to genes involved in UDP-glucuronic acid synthesis. S. saprophyticus ATCC 15305 CP, visualized by immunoelectron microscopy, was extracted and purified using anionic-exchange and size exclusion chromatography. Analysis of the purified CP by 1H and 13C nuclear magnetic resonance (NMR) spectroscopy and gas-liquid chromatography revealed two types of branched tetrasaccharide repeating units composed of the following: Sug represents two stereoisomers of 2-acetamido-2,6-dideoxy-hexos-4-ulose residues, one of which has an arabino configuration. The encapsulated ATCC 15305 strain was resistant to complement-mediated opsonophagocytic killing by human neutrophils, whereas the acapsular mutant C1 was susceptible. None of 14 clinical isolates reacted with antibodies to the ATCC 15305 CP. However, 11 of the 14 S. saprophyticus isolates were phenotypically encapsulated based on their resistance to complement-mediated opsonophagocytic killing and their failure to hemagglutinate when cultivated aerobically. Ten of the 14 clinical strains carried homologues of the conserved staphylococcal capD gene or the S. saprophyticus gtaB gene, or both. Our results suggest that some strains of S. saprophyticus are encapsulated and that more than one capsular serotype exists.Approximately 13 million women develop urinary tract infections (UTIs) annually in the United States, with a recurrence rate between 25% and 44% (45). Staphylococcus saprophyticus is second only to Escherichia coli as a cause of uncomplicated UTI in young women (45, 46). A novobiocin-resistant member of the coagulase-negative staphylococci (60), S. saprophyticus has rarely exhibited resistance to other antibiotics (25). However, a recent report (19) indicated that methicillin-resistant S. saprophyticus isolates have emerged in Japan. The gastrointestinal tract and the vagina are the major reservoirs of S. saprophyticus (18, 30) and the likely sources of recurrent infection (20, 37, 49). Approximately 40% of patients with S. saprophyticus UTI present with acute pyelonephritis (22, 30). These patients experience symptoms more severe than those of patients infected by E. coli (24), and they are more likely to develop recurrent infections (21).A number of potential virulence factors have been identified in S. saprophyticus. Gatermann et al. showed that in a rodent model of ascending UTI, the production of urease contributes to S. saprophyticus growth and pathogenicity in the bladder (10, 12). Other putative virulence factors of S. saprophyticus include a surface-associated lipase (11, 51, 53), the collagen binding protein SdrI (52), and a cell wall-anchored hemagglutinin protein that mediates the binding of S. saprophyticus to sheep erythrocytes, fibronectin, and human uroepithelial cells (14, 29, 34, 35). The hemagglutinin was dubbed UafA in the sequenced ATCC 15305 strain, and deletion of the uafA gene resulted in reduced S. saprophyticus hemagglutination (HA) and adherence to human bladder carcinoma cells (29). Kuroda et al. noted that UafA-mediated adherence of S. saprophyticus to the T24 cell line was inhibited by the presence of the ATCC 15305 polysaccharide capsule (29).Staphylococcal species produce a variety of extracellular glycopolymers that contribute to the surface properties and virulence of the bacterium, such as capsular polysaccharides (CP), teichoic acids, and poly-N-acetylglucosamine (PNAG). CP production renders Staphylococcus aureus resistant to opsonophagocytic killing; alanine modifications of teichoic acids promote bacterial resistance to antimicrobial peptides (40); and PNAG is involved in biofilm formation (4). Recently, the secretion of another anionic polymer (poly-γ-dl-glutamic acid) by certain other coagulase-negative staphylococci was reported (28). Polyglutamic acid production is enhanced under high-salt conditions and may contribute to the survival of Staphylococcus epidermidis on human skin.S. saprophyticus strain 15305 does not produce PNAG or polyglutamic acid (28, 29), but this uropathogenic species is encapsulated. CP are lacking in isolates of S. epidermidis, the most common of the coagulase-negative species, but genomic evidence indicates that Staphylococcus haemolyticus (7, 57), S. saprophyticus (29), and Staphylococcus carnosus (47) carry capsule loci with genetic similarity to the Staphylococcus aureus cap5 (cap8) gene locus. In this study, we purified and characterized the CP produced by S. saprophyticus ATCC 15305 and investigated the CP phenotype of S. saprophyticus clinical isolates.  相似文献   

10.
Microsporidia are a group of highly adapted obligate intracellular parasites that are now recognized as close relatives of fungi. Their adaptation to parasitism has resulted in broad and severe reduction at (i) a genomic level by extensive gene loss, gene compaction, and gene shortening; (ii) a biochemical level with the loss of much basic metabolism; and (iii) a cellular level, resulting in lost or cryptic organelles. Consistent with this trend, the mitochondrion is severely reduced, lacking ATP synthesis and other typical functions and apparently containing only a fraction of the proteins of canonical mitochondria. We have investigated the mitochondrial protein import apparatus of this reduced organelle in the microsporidian Encephalitozoon cuniculi and find evidence of reduced and modified machinery. Notably, a putative outer membrane receptor, Tom70, is reduced in length but maintains a conserved structure chiefly consisting of tetratricopeptide repeats. When expressed in Saccharomyces cerevisiae, EcTom70 inserts with the correct topology into the outer membrane of mitochondria but is unable to complement the growth defects of Tom70-deficient yeast. We have scanned genomic data using hidden Markov models for other homologues of import machinery proteins and find evidence of severe reduction of this system.Microsporidia are a eukaryotic group highly adapted as obligate intracellular parasites (31, 50). They infect a diverse range of vertebrate and invertebrate animal hosts. In humans they are the cause of a number of diseases (e.g., gastroenteritis, encephalitis, and hepatitis), having their greatest impact on immunocompromised individuals, notably in children with human immunodeficiency virus (14, 31). Microsporidia are most closely related to fungi, although their high level of specialization as intracellular parasites obscured this relationship for a long time (18, 25, 30). Gene phylogenies now firmly connect these two groups, although it remains uncertain whether microsporidia are sisters to the fungi or represent a lineage derived from within fungal diversity (21, 28).A clear adaptive response to parasitism in microsporidia has been a reduction in cellular complexity. This was first recognized at an ultrastructural level with the apparent lack of peroxisomes, flagella, stacked Golgi bodies, and mitochondria (31). This reductive evolution is mirrored at a genomic level, with microsporidia containing the smallest eukaryotic genomes described to date (28, 29). The complete genomic sequence from the human microsporidian parasite Encephalitozoon cuniculi reveals a genome of only ∼2.9 Mb containing approximately 2,000 genes, in contrast to the 6,000 genes found in the genome of the model fungus Saccharomyces cerevisiae. The minimal genome of E. cuniculi has been achieved through three mechanisms in concert: (i) gene loss, resulting in broad loss of biochemical pathways and capabilities, including much basic energy metabolism and numerous anabolic pathways; (ii) gene compaction with an average intergenic space of ∼130 bp; and (iii) gene shortening, with E. cuniculi genes being on average 14% shorter than their homologues in fungi such as S. cerevisiae (28, 45). Thus, microsporidian evolution has apparently been shaped by a very strong trend to eliminate superfluous molecular and biochemical complexity.Despite earlier suppositions that microsporidia lacked mitochondria, genome and expressed sequence tag data from microsporidia suggested the presence of several proteins typically targeted to this organelle (3, 19, 20, 24, 28, 38). Immunolocalization of a mitochondrial Hsp70 to small double membrane-bound organelles in Trachipleistophora hominis provided strong evidence for the existence of a mitochondrion in microsporidia, albeit a simplified organelle that lacks cisternae (48). Annotation of genomic data from E. cuniculi provided compelling matches for only 22 proteins implicated in mitochondrial function, suggesting that the metabolism of this relict mitochondrion (or mitosome) is also significantly reduced compared to that of canonical mitochondria (28). Further, no mitochondrial genome has been retained; thus, biogenesis of this organelle is wholly dependent on nucleus-encoded proteins. Based on these 22 proteins, a major role for the mitosome is iron-sulfur cluster assembly (22, 28). No genes have been found for ATP synthesis via oxidative phosphorylation, suggesting loss of this activity in mitosomes (28, 46). While it is likely that further mitosome-targeted proteins will be identified, it is clear that compared to mitochondria from fungal relatives, which are known to import ∼1,000 proteins (40, 44), microsporidian mitosomes represent organelles with highly reduced proteomes, a feature consistent with other traits of cellular reduction.The highly reduced state of the microsporidian mitosome, requiring only a fraction of the protein diversity of other mitochondria, presents an interesting case for studying organelle biogenesis—particularly the machinery for protein import of nucleus-encoded proteins. Mitochondrial protein import has been best characterized in fungi, and in these systems most proteins are imported via four major import complexes: a TOM (translocase of the outer mitochondrial membrane), a SAM (sorting and assembly machinery), and one of two TIMs (translocase of the inner mitochondrial membrane), TIM23 or TIM22 (see Fig. Fig.5A)5A) (5, 36). These complexes are broadly conserved throughout fungi as well as animals (15). Mitochondrial proteins can take one of several routes to the mitochondrion via this apparatus (5, 36). Broadly, soluble matrix proteins are recognized at the TOM complex by the receptor protein Tom20 through the binding of N-terminal presequences with characteristic features (1, 5, 7, 8, 36). These proteins are passed through the pore protein Tom40 of the TOM to the TIM23 complex and then driven into the mitochondrial matrix by way of the presequence translocase-associated motor (PAM) complex, where their presequences are subsequently removed. Some membrane proteins can also be released into the inner membrane from the TIM23 complex. Mitochondrial proteins that apparently lack such an extension, notably including many of the membrane proteins, are recognized by internal sequence elements. Tom70 has a greater role in recognizing these internal signals and thus the import of hydrophobic proteins (4, 11, 32, 39, 47). Such hydrophobic proteins are often bound by cytosolic molecular chaperones (Hsp70 and/or Hsp90) en route to the mitochondrion, and Tom70 is known to independently bind to both the chaperone and the substrate protein (7, 23, 33, 52). While a measure of substrate overlap between Tom20 and Tom70 occurs, the division of responsibility between these two receptors has likely evolved in response to the wide range of substrate proteins that must be imported into mitochondria and the need to handle this complexity.Open in a separate windowFIG. 5.Schematics of the protein import machinery and pathways in yeast mitochondria (A) and E. cuniculi mitosome (B) based on identified homologues of the general fungal/animal pathways. Protein components of the yeast system were all represented by HMMs used to search the microsporidian genomic data and represent the major presequence-dependent and presequence-independent pathways. Homologues identified in E. cuniculi indicate a severely reduced import apparatus utilizing elements of the presequence-independent pathway.For microsporidia little is known of the protein import apparatus for their relict mitochondrion, the mitosome. Has the very reduced organelle proteome, in concert with a genome-wide trend of the loss of redundant or superfluous genes, resulted in a smaller and/or derived import apparatus? In this study we have investigated the microsporidian mitosome protein import apparatus from E. cuniculi in order to evaluate how this apparatus has responded to the reduction in the number of proteins required to be imported and the overall radical reduction in the number and size of proteins encoded in the nuclear genome. A putative homologue of the outer membrane receptor protein Tom70 is of particular interest as the only receptor for the TOM complex and, given the known structure of Tom70 proteins, provides a highly informative example of how proteins can be shortened in the course of genome reduction.  相似文献   

11.
12.
13.
Lactobacillus rhamnosus GG is a well-established Gram-positive probiotic strain, whose health-benefiting properties are dependent in part on prolonged residence in the gastrointestinal tract and are likely dictated by adherence to the intestinal mucosa. Previously, we identified two pilus gene clusters (spaCBA and spaFED) in the genome of this probiotic bacterium, each of which contained the predicted genes for three pilin subunits and a single sortase. We also confirmed the presence of SpaCBA pili on the cell surface and attributed an intestinal mucus-binding capacity to one of the pilin subunits (SpaC). Here, we report cloning of the remaining pilin genes (spaA, spaB, spaD, spaE, and spaF) in Escherichia coli, production and purification of the recombinant proteins, and assessment of the adherence of these proteins to human intestinal mucus. Our findings indicate that the SpaB and SpaF pilin subunits also exhibit substantial binding to mucus, which can be inhibited competitively in a dose-related manner. Moreover, the binding between the SpaB pilin subunit and the mucosal substrate appears to operate through electrostatic contacts and is not related to a recognized mucus-binding domain. We conclude from these results that it is conceivable that two pilin subunits (SpaB and SpaC) in the SpaCBA pilus fiber play a role in binding to intestinal mucus, but for the uncharacterized and putative SpaFED pilus fiber only a single pilin subunit (SpaF) is potentially responsible for adhesion to mucus.The human intestinal microbiota is comprised of more than 1,000 species of commensal and probiotic bacteria, including several members of the Gram-positive genus Lactobacillus (42, 52). Many strains of lactobacilli have a variety of health-promoting effects in humans and consequently have been used commercially as probiotics in foods and nutritional supplements (for a review, see reference 48). Often a necessary precondition for colonization of the human gastrointestinal (GI) tract by probiotic bacteria is preferential adherence to the intestinal mucosa, which in turn prolongs and stabilizes intestinal residence, possibly triggering a variety of defensive host cell immune responses and excluding pathogenic bacteria by competitive inhibition or steric hindrance (48). The outermost layer of the intestinal mucosa, which is a secreted and hydrated mucus gel that acts as a protective barrier and filter, consists primarily of a heterogeneous mixture of highly glycosylated membrane-associated and secreted glycoproteins called mucins (36). Although many studies have demonstrated that various probiotic Lactobacillus spp. adhere initially to the mucus gel layer, relatively few details about the overall molecular mechanism of mucosal adhesion are known (for a review, see reference 23). Nonetheless, several studies have reported that the adherence of Lactobacillus cells to the mucosal barrier is frequently due to a surface protein-mediated interaction. For example, Rojas et al. (44) determined that the ability of Lactobacillus fermentum 104R (reclassified as Lactobacillus reuteri 104R) to bind to porcine small intestinal mucus and gastric mucin was facilitated by a cell surface-localized mucus adhesion-promoting protein (MapA). Similarly, Macías-Rodríguez et al. (25) described two adhesion-associated proteins specific for porcine intestinal mucus-related substrates that are attached noncovalently to the cell surface of L. fermentum BCS87. Also, Roos and Jonsson (45) demonstrated adherence between the surface-associated Mub (mucus binding) protein from L. reuteri 1063 and intestinal mucus components derived from porcine and poultry sources. In addition, Pretzer et al. (38) identified a large multidomain surface protein in Lactobacillus plantarum WCFS1 with binding specificity for the mannose moieties in mucins. Interestingly, Kinoshita et al. (19) discovered that glyceraldehyde 3-phosphate dehydrogenase (GAPDH), an enzyme normally associated with glycolysis, is localized on the surface of L. plantarum LA318 cells and adheres tightly to human colonic mucin.Until quite recently, only indirect or circumstantial evidence suggested that pilus-like structures extend from the surface of probiotic lactobacilli (28, 39). However, in a previous study (18) we demonstrated that Lactobacillus rhamnosus GG, a well-studied and widely used probiotic strain (48), is a piliated microbe. Pili are slender, elongated, heteromeric, proteinaceous surface appendages that are present in numerous other Gram-positive bacteria and often mediate adherence between pathogenic and nonpathogenic species and their host cell targets (for reviews, see references 20, 26, 40, and 49) but have now emerged as possible facilitators of adhesion for probiotic colonization of the GI tract (18). Prototypically, the pilus fiber is composed of one major pilin that forms the pilus backbone and two minor pilin subunits (26, 40, 49), one subunit that has a role in signaling the cessation of pilus polymerization (27, 30) and is deposited at the pilus base and at irregular intervals along the pilus backbone and another subunit with an adhesive property that is often localized at the pilus tip (1, 41). The current model of pilus assembly in Corynebacterium diphtheriae (27) suggests that these pilin subunits are connected covalently to one another through isopeptidyl bonds by a membrane-bound transpeptidase (pilin-specific sortase) to produce polymerized pili, which are then attached covalently to the cell wall by a different transpeptidase (the housekeeping sortase) that is capable of recognizing all C-terminal LPXTG-like substrates. The genes encoding these pilus proteins, as well as the pilin-specific sortase, are clustered at the same locus in the genome (54).In a recent study (18), we discovered that in the L. rhamnosus GG genome the genes encoding two different pilus fibers are in the spaCBA and spaFED gene clusters and, based on a genomic comparison with another L. rhamnosus strain (LC705), that the spaCBA cluster is present in only L. rhamnosus GG. Moreover, in our previous work (18) the predicted genes for the major pilin subunit forming the pilus backbone (SpaA and SpaD), one ancillary minor pilin subunit (SpaB and SpaE) that (based on a model for pilus biogenesis) is likely located at the pilus base and decorates the pilus backbone (27), and another larger adherent minor pilin subunit (SpaC and SpaF) were identified in L. rhamnosus GG on the basis of amino acid identity with pilins from two enterococcal species. In addition, we also detected in the sequences of the predicted spaCBA and spaFED gene products the anticipated consensus motifs and domains characteristic of a pilin primary structure, including the Sec-dependent secretion signal, the sortase recognition site, the YPKN pilin-like motif, and the E box (18). Subsequently, expression and localization of intact SpaCBA pili on the cell surface of L. rhamnosus GG were confirmed by immunoblotting and immunogold-labeled electron microscopy using antiserum specific for the SpaC pilin (18). Adhesion interactions between the L. rhamnosus GG strain and intestinal mucosal surfaces have been reported and characterized in previous studies (15, 31, 33, 46, 55-57). However, in our recent study (18), SpaCBA pilus-mediated binding of L. rhamnosus GG cells to human intestinal mucus was revealed in adhesion experiments performed with both L. rhamnosus GG pretreated with SpaC antiserum and an L. rhamnosus GG spaC insertion mutant. More specifically, we demonstrated that there was significant binding between recombinant SpaC pilin protein and intestinal mucus and thus identified a mucus-binding capacity for one of the minor pilin components localized at the tip and along the backbone of the SpaCBA pilus (18). To expand on these findings, here we describe a study in which each of the remaining predicted pilin subunits (SpaA, SpaB, SpaD, SpaE, and SpaF) encoded by genes in the spaCBA and spaFED gene clusters was overproduced in a recombinant form, purified to apparent homogeneity, and characterized to determine its adherence to human intestinal mucus.  相似文献   

14.
Acinetobacter species encounter cycles of feast and famine in nature. We show that populations of Acinetobacter baylyi strain ADP1 remain dynamic for 6 weeks in batch culture. We created a library of lacZ reporters inserted into SalI sites in the genome and then isolated 30 genes with lacZ insertions whose expression was induced by starvation during long-term stationary phase compared with their expression during exponential growth. The genes encode metabolic, gene expression, DNA maintenance, envelope, and conserved hypothetical proteins.Acinetobacter species are ubiquitous soil organisms. Starvation during long-term stationary phase (LTSP) can serve as a laboratory model for natural competitive conditions such as those found in soils (4). This model has been used to study Escherichia coli, and here, we have applied it to Acinetobacter baylyi strain ADP1 (8).During long-term batch culture, an initially clonal population of Escherichia coli experiences five growth stages: lag, exponential, and stationary phases and then death phase and LTSP (4). Prior to LTSP, most of the cells die and serve as nutrition for starving survivors (6, 13). In LTSP, the cell population remains almost steady, declining slowly over years (reviewed in reference 4): for each newly dead cell, slightly less than one new cell is “born.”Much of what is known about starvation physiology during LTSP has been determined through study of the growth advantage in stationary phase (GASP) phenotype. The phenotype arises from genetic changes that occur when cells experience LTSP. During LTSP, the population may have a mutation frequency approaching 1 in 600 base pairs per genome (5).Some physiological changes that take place during LTSP have been described, as have some genes necessary for the development of GASP (13, reviewed in reference 12). Some mutant strains that exhibit GASP have mutations that enhance catabolic efficiency for processing amino acids (14-16). Another nutrient consumed is DNA, which requires genes homologous to strain ADP1''s competence genes (6). Additionally, mutations that knock out SOS polymerases interfere with the formation of GASP mutants (11).  相似文献   

15.
16.
17.
A gene that codes for a novel intracellular poly(3-hydroxybutyrate) (PHB) depolymerase, designated PhaZ1, has been identified in the genome of Bacillus megaterium. A native PHB (nPHB) granule-binding assay showed that purified soluble PhaZ1 had strong affinity for nPHB granules. Turbidimetric analyses revealed that PhaZ1 could rapidly degrade nPHB granules in vitro without the need for protease pretreatment of the granules to remove surface proteins. Notably, almost all the final hydrolytic products produced from the in vitro degradation of nPHB granules by PhaZ1 were 3-hydroxybutyric acid (3HB) monomers. Unexpectedly, PhaZ1 could also hydrolyze denatured semicrystalline PHB, with the generation of 3HB monomers. The disruption of the phaZ1 gene significantly affected intracellular PHB mobilization during the PHB-degrading stage in B. megaterium, as demonstrated by transmission electron microscopy and the measurement of the PHB content. These results indicate that PhaZ1 is functional in intracellular PHB mobilization in vivo. Some of these features, which are in striking contrast with those of other known nPHB granule-degrading PhaZs, may provide an advantage for B. megaterium PhaZ1 in fermentative production of the biotechnologically valuable chiral compound (R)-3HB.Polyhydroxyalkanoates (PHAs) are a group of polyesters that are produced by numerous bacteria as carbon and energy storage materials in response to nutritional stress (13, 27, 29). Poly(3-hydroxybutyrate) (PHB) is the most common and intensively studied PHA. Intracellular native PHB (nPHB) granules are composed of a hydrophobic PHB core and a surface layer consisting of proteins and phospholipids (13). The PHB of intracellular nPHB granules is in an amorphous state. When intracellular nPHB granules are exposed to extracellular environments due to cell death and lysis, the amorphous PHB is transformed into a denatured semicrystalline state. nPHB granules subjected to physical damage or solvent extraction to remove the surface layer can also crystallize into denatured PHB (dPHB) (13, 15). Artificial PHB (aPHB) granules, in which PHB is in an amorphous state, can be prepared from semicrystalline dPHB and detergents (1, 11, 23, 31).Various extracellular PHB depolymerases (PhaZs) that are secreted by many PHB-degrading bacteria have been demonstrated to specifically degrade dPHB (13, 14, 37). One exception is that PhaZ7, an extracellular PHB depolymerase secreted by Paucimonas lemoignei, displays unusual substrate specificity for amorphous PHB, with 3-hydroxybutyrate (3HB) oligomers as the main products of enzymatic hydrolysis (7). PhaZ7 exhibits no enzymatic activity toward dPHB. So far, a growing number of intracellular PHB depolymerases have been characterized. The intracellular PHB depolymerase PhaZa1 of Ralstonia eutropha (also called Cupriavidus necator) H16 has recently been established to be especially important for the intracellular mobilization of accumulated PHB (42). The main in vitro hydrolytic products of PhaZa1 degradation of amorphous aPHB are 3HB oligomers (31). PhaZd1, another intracellular PHB depolymerase of R. eutropha H16, shows no significant amino acid similarity to PhaZa1. The in vitro hydrolytic products of PhaZd1 degradation of amorphous aPHB are also 3HB oligomers. A 3HB monomer is rarely detected as a hydrolytic product (1). The intracellular PHB depolymerase PhaZ of Paracoccus denitrificans was reported previously to degrade protease-treated nPHB granules in vitro, with the release of 3HB dimers and oligomers as the main hydrolytic products (6). Recently, we have identified a novel intracellular PHB depolymerase from Bacillus thuringiensis serovar “israelensis” (39). The B. thuringiensis PhaZ shows no significant amino acid similarity to any known PHB depolymerase. This PhaZ has strong amorphous PHB-hydrolyzing activity and can release a considerable amount of 3HB monomers by the hydrolysis of trypsin-treated nPHB granules (39). It is of note that purified PhaZd1 from R. eutropha, PhaZ from P. denitrificans, and PhaZ from B. thuringiensis need pretreatment of nPHB granules with protease to remove surface proteins for PHB degradation (1, 6, 39). They show only very little or no activity toward nPHB granules without trypsin pretreatment. It has been demonstrated previously that these intracellular PHB depolymerases cannot hydrolyze dPHB (1, 31, 39).(R)-3HB, a biotechnologically valuable chiral compound, has been widely used for syntheses of antibiotics, vitamins, and pheromones (3, 30, 38). One way to produce (R)-3HB is heterologous coexpression of a PHB synthetic operon and a gene encoding an amorphous PHB-degrading PhaZ in Escherichia coli (3, 18, 25, 33, 38). A common problem encountered by this method is that oligomeric and dimeric forms of 3HB often constitute a major portion of the products of enzymatic hydrolysis, thus requiring further hydrolysis by 3HB oligomer hydrolase or heating under alkaline conditions to generate 3HB monomers (3, 18, 25, 33).Bacillus megaterium genes involved in the biosynthesis of nPHB granules have been cloned from strain ATCC 11561 and characterized previously (19, 21, 22). A gene encoding the extracellular PHB depolymerase PhaZ from B. megaterium was recently cloned from strain N-18-25-9 (34). However, little is known about B. megaterium genes involved in the intracellular mobilization of PHB. In this study, we have identified in B. megaterium ATCC 11561 an intracellular PHB depolymerase that could rapidly degrade nPHB granules in vitro without the need for trypsin pretreatment of the nPHB granules. Moreover, almost all the in vitro hydrolytic products released from the degradation of amorphous PHB by this PhaZ were 3HB monomers. This PhaZ could also hydrolyze dPHB with the generation of 3HB monomers. Thus, it appears to be a novel intracellular PHB depolymerase and may have promising potential for biotechnological application in the production of enantiomerically pure (R)-3HB monomers.  相似文献   

18.
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   

19.
Human milk oligosaccharides (HMOs) are the third-largest solid component of milk. Their structural complexity renders them nondigestible to the host but liable to hydrolytic enzymes of the infant colonic microbiota. Bifidobacteria and, frequently, Bifidobacterium longum strains predominate the colonic microbiota of exclusively breast-fed infants. Among the three recognized subspecies of B. longum, B. longum subsp. infantis achieves high levels of cell growth on HMOs and is associated with early colonization of the infant gut. The B. longum subsp. infantis ATCC 15697 genome features five distinct gene clusters with the predicted capacity to bind, cleave, and import milk oligosaccharides. Comparative genomic hybridizations (CGHs) were used to associate genotypic biomarkers among 15 B. longum strains exhibiting various HMO utilization phenotypes and host associations. Multilocus sequence typing provided taxonomic subspecies designations and grouped the strains between B. longum subsp. infantis and B. longum subsp. longum. CGH analysis determined that HMO utilization gene regions are exclusively conserved across all B. longum subsp. infantis strains capable of growth on HMOs and have diverged in B. longum subsp. longum strains that cannot grow on HMOs. These regions contain fucosidases, sialidases, glycosyl hydrolases, ABC transporters, and family 1 solute binding proteins and are likely needed for efficient metabolism of HMOs. Urea metabolism genes and their activity were exclusively conserved in B. longum subsp. infantis. These results imply that the B. longum has at least two distinct subspecies: B. longum subsp. infantis, specialized to utilize milk carbon, and B. longum subsp. longum, specialized for plant-derived carbon metabolism.The newborn infant not only tolerates but requires colonization by commensal microbes for its own development and health (3). The relevance of the gut microbiome in health and disease is reflected by its influence in a number of important physiological processes, from physical maturation of the developing immune system (28) to the altered energy homeostasis associated with obesity (51, 52).Human milk provides all the nutrients needed to satisfy the neonate energy expenditure and a cadre of molecules with nonnutritional but biologically relevant functions (6). Neonatal health is likely dependent on the timely and complex interactions among bioactive components in human milk, the mucosal immune system, and specialized gut microbial communities (30). Human milk contains complex prebiotic oligosaccharides that stimulated the growth of select bifidobacteria (24, 25) and are believed to modulate mucosal immunity and protect the newborn against pathogens (23, 33, 41). These complex oligosaccharides, which are abundantly present in human milk (their structures are reviewed by Ninonuevo et al. [31] and LoCascio et al. [24]), arrive intact in the infant colon (5) and modulate the composition of neonatal gastrointestinal (GI) microbial communities.Bifidobacteria and, frequently, Bifidobacterium longum strains often predominate the colonic microbiota of exclusively breast-fed infants (10, 11). Among the three subspecies of B. longum, only B. longum subsp. infantis grows robustly on human milk oligosaccharides (HMOs) (24, 25). The availability of the complete genome sequences of B. longum subsp. infantis ATCC 15697 (40) and two other B. longum subsp. longum strains (22, 39) made possible the analysis of whole-genome diversity across the B. longum species. Analysis of the B. longum subsp. infantis ATCC 15697 genome has identified regions predicted to enable the metabolism of HMOs (40); however, their distribution across the B. longum spp. remains unknown. We predict that these regions are exclusively conserved in B. longum strains adapted to colonization of the infant gut microbiome and are therefore capable of robust growth on HMOs. In this work, whole-genome microarray comparisons (comparative genomic hybridizations [CGHs]) were used to associate genotypic biomarkers among 15 B. longum strains exhibiting various HMO utilization phenotypes and host associations.  相似文献   

20.
Lactococcus lactis IL1403 harbors a putative sortase A (SrtA) and 11 putative sortase substrates that carry the canonical LPXTG signature of such substrates. We report here on the functionality of SrtA to anchor five LPXTG substrates to the cell wall, thus suggesting that SrtA is the housekeeping sortase in L. lactis IL1403.The GRAS (generally recognized as safe) status of lactic acid bacteria (LAB) has catalyzed a myriad of promising applications using these bacteria as a vehicle for in situ delivery of bioactive proteins such as antigens or digestive enzymes in the gastrointestinal tract of the human host (4, 26). In the context of therapeutic applications of LAB, a major fundamental goal is to determine whether they can be engineered to deliver bioactive proteins to the right bacterial and host locations. We previously designed a protein-targeting system in LAB that addressed proteins to the desired bacterial site (i.e., cytoplasm, cell wall, or external medium), as validated using a model protein reporter and various antigens (14, 15). Studies investigating the use of LAB as vaccine delivery vehicles suggested that the cell-wall-anchored protein form may possess superior ability to induce a strong immune response (3, 14). Among the various surface display systems described in Gram-positive bacteria (13), a dedicated surface protein anchoring system catalyzed by sortases was first described and characterized in Staphylococcus aureus (29). It covalently anchors proteins via their C-terminal cell wall anchor (CWA) domain to the bacterial peptidoglycan. SrtA-like sortases process proteins bearing an LPXTG C-terminal motif and are considered to be the housekeeping sortase that anchors most proteins harboring a sorting signal (32). Other sortases were subsequently shown to anchor proteins bearing the same or other motifs (11, 16).Surprisingly, while the roles of sortases and LPXTG proteins are well documented in pathogens, few reports have examined these functions in other bacteria. A report suggests a relationship between sortase activity and adhesion of the LAB Lactobacillus salivarius, although direct involvement of sortase was not demonstrated (47). Recently, sortase activity was correlated to assembly of pili and adhesion properties in Lactobacillus rhamnosus (21). To further characterize sortase in LAB, we chose an industrially important member of this bacterial group, Lactococcus lactis, to study sortase A functionality in anchoring its putative substrates on the cell wall.  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