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 15 毫秒
1.
The genetic systems responsible for naphthalene and phenanthrene catabolism have been analyzed in the five strains of Burkholderia sp. isolated from soil samples (West Siberia) contaminated by heavy residual fuel and in the laboratory collection strain Burkholderia sp. BS3702 isolated from soil samples of the coke gas plant (Vidnoe, Moscow oblast). The results of this work demonstrate that naphthalene and phenanthrene degradation in the above strains is encoded by the sequences not homologous to the classical nah genes of pseudomonades. In the Burkholderia sp. BS3702 strain, the initial stages of phenanthrene degradation and the subsequent stages of salicylate degradation are controlled by the sequences of different evolutionary descent (phn and nag genes).  相似文献   

2.
Sinorhizobium sp. C4 was isolated from a polycyclic aromatic hydrocarbon (PAH)-contaminated site in Hilo, HI, USA. This isolate can utilize phenanthrene as a sole carbon source. Sixteen metabolites of phenanthrene were isolated and identified, and the metabolic map was proposed. Degradation of phenanthrene was initiated by dioxygenation on 1,2- and 3,4-C, where the 3,4-dioxygenation was dominant. Subsequent accumulation of 5,6- and 7,8-benzocoumarins confirmed dioxygenation on multiple positions and extradiol cleavage of corresponding diols. The products were further transformed to 1-hydroxy-2-naphthoic acid and 2-hydroxy-1-naphthoic acid then to naphthalene-1,2-diol. In addition to the typical degradation pathways, intradiol cleavage of phenanthrene-3,4-diol was proposed based on the observation of naphthalene-1,2-dicarboxylic acid. Degradation of naphthalene-1,2-diol proceeded through intradiol cleavage to produce trans-2-carboxycinnamic acid. Phthalic acid, 4,5-dihydroxyphthalic acid, and protocatechuic acid were identified as probable metabolites of trans-2-carboxycinnamic acid, but no trace salicylic acid or its metabolites were found. This is the first detailed study of PAH metabolism by a Sinorhizobium species. The results give a new insight into microbial degradation of PAHs.  相似文献   

3.
Cycloclasticus sp. strain A5 is able to grow with petroleum polycyclic aromatic hydrocarbons (PAHs), including unsubstituted and substituted naphthalenes, dibenzothiophenes, phenanthrenes, and fluorenes. A set of genes responsible for the degradation of petroleum PAHs was isolated by using the ability of the organism to oxidize indole to indigo. This 10.5-kb DNA fragment was sequenced and found to contain 10 open reading frames (ORFs). Seven ORFs showed homology to previously characterized genes for PAH degradation and were designated phn genes, although the sequence and order of these phn genes were significantly different from the sequence and order of the known PAH-degrading genes. The phnA1, phnA2, phnA3, and phnA4 genes, which encode the α and β subunits of an iron-sulfur protein, a ferredoxin, and a ferredoxin reductase, respectively, were identified as the genes coding for PAH dioxygenase. The phnA4A3 gene cluster was located 3.7 kb downstream of the phnA2 gene. PhnA1 and PhnA2 exhibited moderate (less than 62%) sequence identity to the α and β subunits of other aromatic ring-hydroxylating dioxygenases, but motifs such as the Fe(II)-binding site and the [2Fe-2S] cluster ligands were conserved. Escherichia coli cells possessing the phnA1A2A3A4 genes were able to convert phenanthrene, naphthalene, and methylnaphthalene in addition to the tricyclic heterocycles dibenzofuran and dibenzothiophene to their hydroxylated forms. Significantly, the E. coli cells also transformed biphenyl and diphenylmethane, which are ordinarily the substrates of biphenyl dioxygenases.  相似文献   

4.
The theory that Shigella is derived from multiple independent origins of Escherichia coli (Pupo et al. 2000) has been challenged by recent findings that the virulence plasmids (VPs) and the chromosomes share a similar evolutionary history (Escobar-Paramo et al. 2003), which suggests that an ancestral VP entered an E. coli strain only once, which gave rise to Shigella spp. In an attempt to resolve these conflicting theories, we constructed three phylogenetic trees in this study: a robust chromosomal tree using 23 housekeeping genes from 46 strains of Shigella and enteroinvasive E. coli (EIEC), a chromosomal tree using 4 housekeeping genes from 19 EcoR strains and 46 Shigella/EIEC strains, and a VP tree using 5 genes outside of the VP cell-entry region from 38 Shigella/EIEC strains. Both chromosomal trees group Shigella into three main clusters and five outliers, and strongly suggest that Shigella has multiple origins within E. coli. Most strikingly, the VP tree shows that the VPs from two main Shigella clusters, C1 and C2, are more closely related, which contradicts the chromosomal trees that place C2 and C3 next to each other but C1 at a distance. Additionally, we have identified a complete tra operon of the F-plasmid in the genome sequence of an EIEC strain and found that two other EIEC strains are also likely to possess a complete tra operon. All lines of evidence support an alternative multiorigin theory that transferable diverse ancestral VPs entered diverse origins of E. coli multiple times during a prolonged period of time, resulting in Shigella species with diverse genomes but similar pathogenic properties. Electronic Supplementary Material Electronic Supplementary material is available for this article at and accessible for authorised users. [Reviewing Editor: Dr. Martin Kreitman] Jian Yang and Huan Nie contributed equally to this work.  相似文献   

5.
A phenanthrene-assimilating bacterium which belongs to the genus Aeromonas was isolated from soil. The cells which adapted to phenanthrene required a growth lag time on a naphthalene medium. The cells oxidized l-hydroxy-2-naphthoate (1H2NA), 2-carboxybenzaldehyde (2CBAL), o-phthalate (OPA) and protocatechuate (PCA) but did not oxidize salicylaldehyde (SAL), salicylate (SA) and catechol (CAT) which are intermediates in naphthalene catabolism. Using the cell-free extract, the same results were obtained in oxidative capacity. The intact cells metabolized 1H2NA and 2CBAL without the lag time, giving 2CBAL and PCA, respectively. The ammonium sulfate-treated extract prepared from the cells grown in phenanthrene medium, converted 1H2NA to 2CBAL and 2CBAL to OPA. It was suggested that the Aeromonas sp. degraded phenanthrene through OPA.  相似文献   

6.
Seo JS  Keum YS  Hu Y  Lee SE  Li QX 《Biodegradation》2007,18(1):123-131
Burkholderia sp. C3 was isolated from a polycyclic aromatic hydrocarbon (PAH)-contaminated site in Hilo, Hawaii, USA, and studied for its degradation of phenanthrene as a sole carbon source. The initial 3,4-C dioxygenation was faster than 1,2-C dioxygenation in the first 3-day culture. However, 1-hydroxy-2-naphthoic acid derived from 3,4-C dioxygenation degraded much slower than 2-hydroxy-1-naphthoic acid derived from 1,2-C dioxygenation. Slow degradation of 1-hydroxy-2-naphthoic acid relative to 2-hydroxy-1-naphthoic acid may trigger 1,2-C dioxygenation faster after 3 days of culture. High concentrations of 5,6-␣and 7,8-benzocoumarins indicated that meta-cleavage was the major degradation mechanism of phenanthrene-1,2- and -3,4-diols. Separate cultures with 2-hydroxy-1-naphthoic acid and 1-hydroxy-2-naphthoic acid showed that the degradation rate of the former to naphthalene-1,2-diol was much faster than that of the latter. The two upper metabolic pathways of phenanthrene are converged into naphthalene-1,2-diol that is further metabolized to 2-carboxycinnamic acid and 2-hydroxybenzalpyruvic acid by ortho- and meta-cleavages, respectively. Transformation of naphthalene-1,2-diol to 2-carboxycinnamic acid by this strain represents the first observation of ortho-cleavage of two rings-PAH-diols by a Gram-negative species.  相似文献   

7.
8.
The human gut houses a complex group of bacterial genera, including both opportunistic pathogens and commensal micro-organisms. These are regularly exposed to antibiotics, and their subinhibitory concentrations play a pivotal role in shaping the microbial responses. This study was aimed to investigate the effects exerted by sub-MICs of nalidixic acid (NA) on the growth rate, bacterial motility, biofilm formation and expression of outer membrane proteins (OMPs) in a commensal strain of E. coli. The NA-sensitive strain was sequentially passaged under sub-MICs of NA. E-test was used to determine the MIC values of NA. Results indicated significant changes in the growth profile of commensal E. coli upon exposure to NA at sub-MICs. Differential expression of OMPs was observed in cells treated with sub-MICs of NA. Bacterial motility was reduced under 1/2 MIC of NA. Interestingly, successive passaging under 1/2 MIC of NA led to the emergence of resistant E. coli with an increased MIC value of 64 µg ml−1 in just 24 days. The NA-resistant variant was confirmed by comparing its 16S rRNA sequence to that of the sensitive commensal strain. Mutations in the Quinolone Resistance-Determining Regions (QRDRs) of chromosomal gyrA, and Topoisomerase IV-encoding parC genes were detected in NA-resistant E. coli. Our results demonstrate how antibiotics play an important role as signalling molecules or elicitors in driving the pathogenicity of commensal bacteria in vitro.  相似文献   

9.
The polycyclic aromatic hydrocarbon phenanthrene was mineralized in two stages by soil, estuarine water, and sediment microbial populations. At high concentrations, phenanthrene was degraded, with the concomitant production of biomass and accumulation of Folin-Ciocalteau-reactive aromatic intermediates. Subsequent consumption of these intermediates resulted in a secondary increase in biomass. Analysis of intermediates by high-performance liquid chromatography, thin-layer chromatography, and UV absorption spectrometry showed 1-hydroxy-2-naphthoic acid (1H2NA) to be the predominant product. A less pronounced two-stage mineralization pattern was also observed by monitoring 14CO2 production from low concentrations (0.5 mg liter−1) of radiolabeled phenanthrene. Here, mineralization of 14C-labeled 1H2NA could explain the incremental 14CO2 produced during the later part of the incubations. Accumulation of 1H2NA by isolates obtained from enrichments was dependent on the initial phenanthrene concentration. The production of metabolites during polycyclic aromatic hydrocarbon biodegradation is discussed with regard to its possible adaptive significance and its methodological implications.  相似文献   

10.
11.
The first biosynthetic system for lactate (LA)-based polyesters was previously created in recombinant Escherichia coli (Taguchi et al. 2008). Here, we have begun efforts to upgrade the prototype polymer production system to a practical stage by using metabolically engineered Gram-positive bacterium Corynebacterium glutamicum as an endotoxin-free platform. We designed metabolic pathways in C. glutamicum to generate monomer substrates, lactyl-CoA (LA-CoA), and 3-hydroxybutyryl-CoA (3HB-CoA), for the copolymerization catalyzed by the LA-polymerizing enzyme (LPE). LA-CoA was synthesized by D-lactate dehydrogenase and propionyl-CoA transferase, while 3HB-CoA was supplied by β-ketothiolase (PhaA) and NADPH-dependent acetoacetyl-CoA reductase (PhaB). The functional expression of these enzymes led to a production of P(LA-co-3HB) with high LA fractions (96.8 mol%). The omission of PhaA and PhaB from this pathway led to a further increase in LA fraction up to 99.3 mol%. The newly engineered C. glutamicum potentially serves as a food-grade and biomedically applicable platform for the production of poly(lactic acid)-like polyester.  相似文献   

12.

Background  

Highly virulent enterohemorrhagic Escherichia coli O157:H7 strains possess three sodC genes encoding for periplasmic Cu, Zn superoxide dismutases: sodC, which is identical to the gene present in non-pathogenic E. coli strains, and sodC-F1 and sodC-F2, two nearly identical genes located within lambdoid prophage sequences. The significance of this apparent sodC redundancy in E. coli O157:H7 has not yet been investigated.  相似文献   

13.
14.
Aims: Investigating mechanisms of lethality enhancement when Escherichia coli O157:H7, and selected E. coli mutants, were exposed to tert‐butylhydroquinone (TBHQ) during ultra‐high pressure (UHP) treatment. Methods and Results: Escherichia coli O157:H7 EDL‐933, and 14 E. coli K12 strains with mutations in selected genes, were treated with dimethyl sulfoxide solution of TBHQ (15–30 ppm), and processed with UHP (400 MPa, 23 ± 2°C for 5 min). Treatment of wild‐type E. coli strains with UHP alone inactivated 2·4–3·7 log CFU ml?1, whereas presence of TBHQ increased UHP lethality by 1·1–6·2 log CFU ml?1; TBHQ without pressure was minimally lethal (0–0·6 log reduction). Response of E. coli K12 mutants to these treatments suggests that iron–sulfur cluster‐containing proteins ([Fe–S]‐proteins), particularly those related to the sulfur mobilization (SUF system), nitrate metabolism, and intracellular redox potential, are critical to the UHP–TBHQ synergy against E. coli. Mutations in genes maintaining redox homeostasis and anaerobic metabolism were associated with UHP–TBHQ resistance. Conclusions: The redox cycling activity of cellular [Fe–S]‐proteins may oxidize TBHQ, potentially leading to the generation of bactericidal reactive oxygen species. Significance and Impact of the Study: A mechanism is proposed for the enhanced lethality of UHP by TBHQ against E. coli O157:H7. The results may benefit food processors using UHP–based preservation, and biologists interested in piezophilic micro‐organisms.  相似文献   

15.
A phenanthrene-mineralizing Pseudomonas sp., designated UG14, was isolated from creosote-contaminated soil. It contained two plasmids, of approximately 77 kb and 76 kb, the smaller of which contained DNA sequences that hybridized with probes specific for ndoB and xylE, genes involved in catabolism of aromatic hydrocarbons. At initial phenanthrene concentrations of 10, 50, 200 and 1000 mg/l broth, 27%, 19%, 7.7% and 3.3%, respectively, of the [9-14C]phenanthrene was recovered as 14CO2 after 36 days' incubation at 30°C. Most 14C-label was converted to a water-soluble metabolite tentatively identified as 1-hydroxy-2-naphthoic acid. Rhamnolipid biosurfactants produced by P. aeruginosa UG2 enhanced mineralization of 50 mg phenanthrene/l by Pseudomonas sp. UG14. With the biosurfactant at 0, 25 and 250 mg rhamnose equivalents/l, 6.5%, 8.2% and 9.8%, respectively, of the phenanthrene was mineralized after 35 days.M.A. Providenti, H. Lee and J.T. Trevors are with the Department of Environmental Biology, University of Guelph, Guelph, Ontario, N1G 2W1, Canada; C.W. Greer is with the National Research Council Canada, Biotechnology Research Institute, 6100 Royalmount Ave, Montreal, Quebec, H4P 2R2, Canada.  相似文献   

16.
17.
Summary Four genes, nagR, A, B and E, clustered in the nag locus of Escherichia coli K12 and Klebsiella pneumoniae, were cloned and physically mapped, and the corresponding gene products involved in amino sugar metabolism identified. Expression of the nag genes was also analysed using a series of lacZ fusions. In both bacteria, the genes are arranged in two divergent operons and controlled by a common NagR repressor. The corresponding gene nagR was found to map in the first operon together with the promoter proximal gene nagB, encoding the enzyme d-glucosamine isomerase (deaminase) (NagB) and the middle gene nagA, coding for N-acetyl-glucosamine deacetylase (NagA). Polar mutations in nagB and nagA prevent the efficient expression of nagR and cause constitutive expression of all nag genes. This includes the gene nagE encoding Enzyme IINag of the phosphoenolpyruvate-dependent carbohydrate phosphotransferase system (PTS), encoded in the second divergently transcribed operon. No further gene is found in this operon which in both organisms is directly adjacent to the gene glnS. It is interesting that the NagR repressor also affects the mannose PTS (genes manX, Y, Z), the second transport system involved in amino sugar uptake and phosphorylation.  相似文献   

18.
Three filamentous fungi were examined for the ability to biotransform phenanthrene to oxidative (phase I) and conjugative (phase II) metabolites. Phenanthrene metabolites were purified by high-performance liquid chromatography (HPLC) and identified by UV/visible absorption, mass, and1H NMR spectra.Aspergillus niger ATCC 6275,Syncephalastrum racemosum UT-70, andCunninghamella elegans ATCC 9245 initially transformed [9-14C]phenanthrene to produce metabolites at the 9,10-, 1,2-, and 3,4- positions. Subsequently, sulfate conjugates of phase I metabolites were formed byA. niger, S. racemosum, andC. elegans. Minor glucuronide conjugates of 9-phenanthrol and phenanthrenetrans-9,10-dihydrodiol were formed byS. racemosum andA. niger, respectively. In addition,C. elegans produced the glucose conjugates 1-phenanthryl -d-glucopyranoside and 2-hydroxy-1-phenanthryl -d-glucopyranoside, a novel metabolite. [9-14C]Phenanthrene metabolites were not detected in organic extracts from biotransformation experiments with the yeasts,Candida lipolytica 37-1,Candida tropicalis ATCC 32113, andCandida maltosa R-42.  相似文献   

19.
Keratinase from Pseudomonas aeruginosa KS-1 was expressed constitutively as an extracellular protein in Escherichia coli with high specific activity of 3.7 kU/mg. It was purified fourfold as a 33 kDa monomeric protein by Q-Sepharose ion exchange chromatography with a recovery of 95%. It is a serine protease with optimal activity at pH 9 and 50°C. It was stable from pH 4 to 12 for 1 h with a t1/2 of 12 min at 70°C. It hydrolyzed haemoglobin > fibrin > feather keratin > azo-casein > casein > meat protein > gelatin. Among synthetic substrates, it efficiently hydrolyzed N-Suc-ala-ala-pro-phe-pNA, N-Suc-ala-ala-ala-pNA, N-Suc-ala-ala-pro-leu-pNA and also plasmin substrate, d-Val-Leu-Lys-pNA  相似文献   

20.
The presence of 11 genes encoding subtilisin-like serine proteases was demonstrated by cloning from the genome of alkaliphilic Bacillus sp. strain KSM-LD1. This strain exoproduces the oxidatively stable alkaline protease LD-1 (Saeki et al. Curr Microbiol, 47:337–340, 2003). Among the 11 genes, six genes encoding alkaline proteases (SA, SB, SC, SD, SE, and LD-1) were expressed in Bacillus hosts. However, the other five genes for subtilisin-like proteases (SF, SG, SH, SI, and SJ) were expressed in neither Bacillus hosts nor Escherichia coli. The deduced amino acid sequences of SA, SB, SC, SF, SG, SH, SI, and SJ showed similarity to those of other subtilisin-like proteases from Bacillus strains with only 38 to 86% identity. The deduced amino acid sequence of SD was completely identical to that of an oxidatively stable alkaline protease from Bacillus sp. strain SD521, and that of SE was almost identical to that of a high-molecular mass subtilisin from Bacillus sp. strain D-6 with 99.7% identity. There are four to nine subtilisin-like serine protease genes in the reported genomes of Bacillus strains. At least 11 genes for the enzymes present in the genome of Bacillus sp. strain KSM-LD1, and this is the greatest number identified to date.  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