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.
Pseudomonas rhodesiae KK1 was isolated from a former manufactured-gas plant site, due to its ability to grow rapidly in a mixture of polycyclic aromatic hydrocarbons (PAHs). Radiorespirometric analysis revealed that strain KK1 was found to be able to mineralize anthracene, naphthalene and phenanthrene. Notably, phenanthrene-grown cells were able to mineralize anthracene much more rapidly than naphthalene-grown cells. Comparative analysis of amino acid sequences from 17 randomly selected dioxygenases capable of hydroxylating unactivated aromatic nuclei indicated that the enzymes for catabolism of PAHs, such as naphthalene and phenanthrene, might exist redundantly in strain KK1. Northern hybridization for cells grown on naphthalene or phenanthrene, using the putative naphthalene or phenanthrene dioxygenase gene fragment as a probe, suggested that the enzyme for naphthalene catabolism might share some homology in deduced amino acid sequences with phenanthrene dioxygenases. Also, it was found that three lipids (17:0 cyclo, 18:1 omega7c, 19:0 cyclo) increased in response to both naphthalene and phenanthrene, while the shift of other lipids varied from substrate to substrate.  相似文献   

2.
Naphthalene and phenanthrene are transformed by enzymes encoded by the pah gene cluster of Pseudomonas putida OUS82. The pahA and pahB genes, which encode the first and second enzymes, dioxygenase and cis-dihydrodiol dehydrogenase, respectively, were identified and sequenced. The DNA sequences showed that pahA and pahB were clustered and that pahA consisted of four cistrons, pahAa, pahAb, pahAc, and pahAd, which encode ferredoxin reductase, ferredoxin, and two subunits of the iron-sulfur protein, respectively.  相似文献   

3.
[13C6]salicylate, [U-13C]naphthalene, and [U-13C]phenanthrene were synthesized and separately added to slurry from a bench-scale, aerobic bioreactor used to treat soil contaminated with polycyclic aromatic hydrocarbons. Incubations were performed for either 2 days (salicylate, naphthalene) or 7 days (naphthalene, phenanthrene). Total DNA was extracted from the incubations, the "heavy" and "light" DNA were separated, and the bacterial populations associated with the heavy fractions were examined by denaturing gradient gel electrophoresis (DGGE) and 16S rRNA gene clone libraries. Unlabeled DNA from Escherichia coli K-12 was added to each sample as an internal indicator of separation efficiency. While E. coli was not detected in most analyses of heavy DNA, a low number of E. coli sequences was recovered in the clone libraries associated with the heavy DNA fraction of [13C]phenanthrene incubations. The number of E. coli clones recovered proved useful in determining the relative amount of light DNA contamination of the heavy fraction in that sample. Salicylate- and naphthalene-degrading communities displayed similar DGGE profiles and their clone libraries were composed primarily of sequences belonging to the Pseudomonas and Ralstonia genera. In contrast, heavy DNA from the phenanthrene incubations displayed a markedly different DGGE profile and was composed primarily of sequences related to the Acidovorax genus. There was little difference in the DGGE profiles and types of sequences recovered from 2- and 7-day incubations with naphthalene, so secondary utilization of the 13C during the incubation did not appear to be an issue in this experiment.  相似文献   

4.
Nineteen polycyclic aromatic hydrocarbon (PAH)-degrading bacteria were isolated from environmental samples in Kuwait, Indonesia, Thailand, and Japan by enrichment with either naphthalene or phenanthrene as a sole carbon source. Sequence analyses of the 16-S rRNA gene indicated that at least seven genera (Ralstonia, Sphingomonas, Burkholderia, Pseudomonas, Comamonas, Flavobacterium, and Bacillus) were present in this collection. Determination of the ability of the isolates to use PAH and its presumed catabolic intermediates suggests that the isolates showed multiple phenotypes in terms of utilization and degradation pathways. The large subunit of the terminal oxygenase gene (phnAc) from Burkholderia sp. strain RP007 hybridized to 32% (6/19) of the isolates, whilst gene probing using the large subunit of terminal oxygenase gene (pahAc) from Pseudomonas putida strain OUS82 revealed no pahAc-like genes amongst the isolates. Using three degenerated primer sets (pPAH-F/NR700, AJ025/26, and RieskeF/R), targeting a conserved region with the genes encoding the large subunit of terminal oxygenase successfully amplified material from 6 additional PAH-degrading isolates. Sequence analyses showed that the large subunit of terminal oxygenase in 4 isolates was highly homologous to the large subunit of naphthalene dioxygenase gene from Ralstonia sp. strain U2. However, we could not obtain any information on the oxygenase system involved in the naphthalene and/or phenathrene degradation by 7 other strains. These results suggest that PAH-degrading bacteria are diverse, and that there are still many unidentified PAH-degrading bacteria.  相似文献   

5.
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   

6.
[13C6]salicylate, [U-13C]naphthalene, and [U-13C]phenanthrene were synthesized and separately added to slurry from a bench-scale, aerobic bioreactor used to treat soil contaminated with polycyclic aromatic hydrocarbons. Incubations were performed for either 2 days (salicylate, naphthalene) or 7 days (naphthalene, phenanthrene). Total DNA was extracted from the incubations, the “heavy” and “light” DNA were separated, and the bacterial populations associated with the heavy fractions were examined by denaturing gradient gel electrophoresis (DGGE) and 16S rRNA gene clone libraries. Unlabeled DNA from Escherichia coli K-12 was added to each sample as an internal indicator of separation efficiency. While E. coli was not detected in most analyses of heavy DNA, a low number of E. coli sequences was recovered in the clone libraries associated with the heavy DNA fraction of [13C]phenanthrene incubations. The number of E. coli clones recovered proved useful in determining the relative amount of light DNA contamination of the heavy fraction in that sample. Salicylate- and naphthalene-degrading communities displayed similar DGGE profiles and their clone libraries were composed primarily of sequences belonging to the Pseudomonas and Ralstonia genera. In contrast, heavy DNA from the phenanthrene incubations displayed a markedly different DGGE profile and was composed primarily of sequences related to the Acidovorax genus. There was little difference in the DGGE profiles and types of sequences recovered from 2- and 7-day incubations with naphthalene, so secondary utilization of the 13C during the incubation did not appear to be an issue in this experiment.  相似文献   

7.
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   

8.
Naphthalene utilizing bacteria were isolated from several sites above and below the discharge from a coking plant. The distribution of the bacteria was influenced by the effluent. Of these isolates, 11·4% obtained from the effluent discharge site, contained plasmids. No conjugal transfer of naphthalene utilizing ability was observed in over 1000 matings. Curing and transformation experiments demonstrated that one plasmid pNB33 (101 kb) was concerned with naphthalene catabolism.  相似文献   

9.
The impact of surfactants on naphthalene and phenanthrene biodegradation and vice versa after surfactant flushing were evaluated using two anionic surfactants: sodium dodecyl sulfate (SDS) and sodium dodecyl benzene sulfonate (SDBS); and two nonionic surfactants: POE (20) sorbitan monooleate (T-maz-80) and octylphenol poly(ethyleneoxy) ethanol (CA-620). Naphthalene and phenanthrene biodegradation varied differently in the presence of different surfactants. Naphthalene biodegradation was not impacted by the presence of SDS. In the presence of T-maz-80 and CA-620, naphthalene biodegradation occurred at a lower rate (0.14 d-1 for T-maz-80 and 0.19 d-1 for CA-620) as compared to un-amended control (0.29 d-1). Naphthalene biodegradation was inhibited by the presence of SDBS. In the presence of SDS, phenanthrene biodegradation occurred at a lower rate (0.10 d-1 as compared to un-amended control of 0.17 d-1) and the presence of SDBS, CA-620 and T-maz-80 inhibited phenanthrene biodegradation. The surfactants also responded differently to the presence of naphthalene and phenanthrene. In the presence of naphthalene, SDS biodegradation was inhibited; SDBS and T-maz-80 depleted at a lower rate (0.41d-1 and 0.12 d-1 as compared to 0.48 d-1 and 0.22 d-1). In the absence of naphthalene, CA-620 was not degradable, while in the presence of naphthalene, CA-620 began to degrade at a comparatively low rate (0.12 d-1). In the presence of phenanthrene, SDS biodegradation occurred at a lower rate (1.2 d-1 as compared to 1.68 d-1) and a similar trend was observed for T-maz-80. The depletion of SDBS and CA-620 did not change significantly. The choice of SDS for naphthalene-contaminated sites would not adversely affect the natural attenuation of naphthalene, in addition, naphthalene was preferentially utilized to SDS by naphthalene-acclimated microorganisms. Therefore, SDS was the best choice. T-maz-80 was also found to be usable in naphthalene-contaminated sites. For phenanthrene contaminated sites, SDS was the only choice.  相似文献   

10.
11.
A Beijerinckia species, capable of oxidizing phenanthrene, biphenyl and other polycyclic aromatic hydrocarbons, was shown to contain two plasmids that were designated pKGl and pKG2. The molecular masses of plasmids pKG1 and pKG2, as determined by electron microscopy, were approximately 147 X 10(6) and 20.8 X 10(6) daltons, respectively. Growth of the organism on benzoate led to the isolation of strains that had lost the ability to grow with phenanthrene and biphenyl. All of the Phn-, Bph- strains had also lost the smaller plasmid, pKG2. The results presented suggest that plasmid pKG2 is responsible for the synthesis of enzymes involved in the degradation of phenanthrene and biphenyl.  相似文献   

12.
Fifteen bacterial strains capable of utilizing naphthalene, phenanthrene, and biphenyl as the sole sources of carbon and energy were isolated from soils and bottom sediments contaminated with waste products generated by chemical and salt producing plants. Based on cultural, morphological, and chemotaxonomic characteristics, ten of these strains were identified as belonging to the genera Rhodococcus, Arthrobacter, Bacillus, and Pseudomonas. All ten strains were found to be halotolerant bacteria capable of growing in nutrient-rich media at NaCl concentrations of 1-1.5 M. With naphthalene as the sole source of carbon and energy, the strains could grow in a mineral medium with 1 M NaCl. Apart from being able to grow on naphthalene, six of the ten strains were able to grow on phenanthrene; three strains, on biphenyl; three strains, on octane; and one strain, on phenol. All of the strains were plasmid-bearing. The plasmids of the Pseudomonas sp. strains SN11, SN101, and G51 are conjugative, contain genes responsible for the degradation of naphthalene and salicylate, and are characterized by the same restriction fragment maps. The transconjugants that gained the plasmid from strain SN11 acquired the ability to grow at elevated NaCl concentrations. Microbial associations isolated from the same samples were able to grow at a NaCl concentration of 2.5 M.  相似文献   

13.
Like other highly urbanized and industrialized estuaries, the Seine estuary (France) has, for decades, received high inputs of polycyclic aromatic hydrocarbons (PAHs). In order to estimate the bioremediation potentials and to identify the bacterial species involved in hydrocarbon degradation, we used microcosms containing seawater from the Seine estuary supplemented with either naphthalene, phenanthrene, fluorene or pyrene. In the microcosms enriched with naphthalene or phenanthrene, hydrocarbon biodegradation was significant within 9 weeks (43% or 46%, respectively), as shown by analyses in GC-MS. In similar microcosms incubated also with naphthalene or phenanthrene, analysis of the 16S rRNA gene sequences (DNA and cDNA) with denaturing gradient gel electrophoresis and clone libraries indicated that the PAH-degrading communities were dominated by Cycloclasticus spp., confirming their universal key role in degradation of low-molecular-weight PAHs in marine environments. However, in contrast to previous studies, we found that Pseudomonas spp. also degraded naphthalene and phenanthrene in seawater; this occurred only after 21 days, as was confirmed by real-time PCR. Although this genus has been abundantly described in the literature as a good PAH-degrading bacterial group in soil or in sediment, to our knowledge, this is the first evidence of a significant fitness in PAH degradation in seawater.  相似文献   

14.
The well-characterized plasmid-encoded naphthalene degradation pathway in Pseudomonas putida PpG7(NAH7) was used to investigate the role of the NAH plasmid-encoded pathway in mineralizing phenanthrene and anthracene. Three Pseudomonas strains, designated 5R, DFC49, and DFC50, were recovered from a polynuclear aromatic hydrocarbon-degrading inoculum developed from a manufactured gas plant soil slurry reactor. Plasmids pKA1, pKA2, and pKA3, approximately 100 kb in size, were isolated from these strains and characterized. These plasmids have homologous regions of upper and lower NAH7 plasmid catabolic genes. By conjugation experiments, these plasmids, including NAH7, have been shown to encode the genotype for mineralization of [9-14C]phenanthrene and [U-14C]anthracene, as well as [1-14C]naphthalene. One strain, Pseudomonas fluorescens 5RL, which has the complete lower pathway inactivated by transposon insertion in nahG, accumulated a metabolite from phenanthrene and anthracene degradation. This is the first direct evidence to indicate that the NAH plasmid-encoded catabolic genes are involved in degradation of polynuclear aromatic hydrocarbons other than naphthalene.  相似文献   

15.
Rhodococcus sp. strain NCIMB 12038 utilizes naphthalene as a sole source of carbon and energy, and degrades naphthalene via salicylate and gentisate. To identify the genes involved in this pathway, we cloned and sequenced a 12-kb DNA fragment containing a gentisate catabolic gene cluster. Among the 13 complete open reading frames deduced from this fragment, three (narIKL) have been shown to encode the enzymes involved in the reactions of gentisate catabolism. NarI is gentisate 1,2-dioxygenase which converts gentisate to maleylpyruvate, NarL is a mycothiol-dependent maleylpyruvate isomerase which catalyzes the isomerization of maleylpyruvate to fumarylpyruvate, and NarK is a fumarylpyruvate hydrolase which hydrolyzes fumarylpyruvate to fumarate and pyruvate. The narX gene, which is divergently transcribed with narIKL, has been shown to encode a functional 3-hydroxybenzoate 6-monooxygenase. This led us to discover that this strain is also capable of utilizing 3-hydroxybenzoate as its sole source of carbon and energy. Both NarL and NarX were purified to homogeneity as His-tagged proteins, and they were determined by gel filtration to exist as a trimer and a monomer, respectively. Our study suggested that the gentisate degradation pathway was shared by both naphthalene and 3-hydroxybenzoate catabolism in this strain.  相似文献   

16.
This paper describes the validation and application of a simple flask-based (14)C-respirometer system designed to assess mineralisation of (14)C-labelled substrates under defined conditions. Validation of this respirometer system indicated stoichiometric CO(2) trapping up to a maximum of 400 micromol of CO(2) (in a single trap). Polycyclic aromatic hydrocarbon (PAH)-degrading bacteria were used to measure growth-linked biodegradation of [(14)C]naphthalene to (14)CO(2). A (14)C activity balance of 101.7+/-8.9% (n=6), after 74 h incubation time and 10 respirometer-opening events, indicated the suitability of the system for monitoring substrate mineralisation. This respirometric apparatus was then successfully applied to assess: (i) the PAH catabolism of microbes in a field contaminated soil, where naphthalene and phenanthrene were rapidly mineralised and (ii) soil-associated organic contaminant bioavailability, where increased soil-phenanthrene contact time resulted in a reduction in phenanthrene mineralisation in the soil. The described respirometer system differs from existing respirometer systems in that the CO(2) trap can be removed and replaced quickly and easily. The system is efficient, reproducible, adaptable to many situations, easy to construct and simple to use, it therefore affords advantages over existing systems.  相似文献   

17.
Pseudomonas sp. strain B13 and Pseudomonas putida OUS82 were genetically tagged with the green fluorescent protein and the Discosoma sp. red fluorescent protein, and the development and dynamics occurring in flow chamber-grown two-colored monospecies or mixed-species biofilms were investigated by the use of confocal scanning laser microscopy. Separate red or green fluorescent microcolonies were formed initially, suggesting that the initial small microcolonies were formed simply by growth of substratum attached cells and not by cell aggregation. Red fluorescent microcolonies containing a few green fluorescent cells and green fluorescent microcolonies containing a few red fluorescent cells were frequently observed in both monospecies and two-species biofilms, suggesting that the bacteria moved between the microcolonies. Rapid movement of P. putida OUS82 bacteria inside microcolonies was observed before a transition from compact microcolonies to loose irregularly shaped protruding structures occurred. Experiments involving a nonflagellated P. putida OUS82 mutant suggested that the movements between and inside microcolonies were flagellum driven. The results are discussed in relation to the prevailing hypothesis that biofilm bacteria are in a physiological state different from planktonic bacteria.  相似文献   

18.
A bacterial strain capable of utilizing phenanthrene as sole source of carbon was isolated from soil and identified as a Bacillus sp. The organism also utilized naphthalene, biphenyl, anthracene, and other aromatic compounds as growth substrates. The organism degraded phenanthrene through the intermediate formation of 1-hydroxy-2-naphthoic acid, which was further metabolized via o-phthalate by a protocatechuate pathway, as evidenced by oxygen uptake and enzymatic studies. Received: 1 December 1999 / Accepted: 5 January 2000  相似文献   

19.
The key genes nahAc and xylE of the naphthalene catabolism of fluorescent Pseudomonas spp. in the total soil DNA samples were detected by the polymerase chain reaction (PCR) technique. The collection of fluorescent Pseudomonas spp. was screened for the occurrence of these genes. The results obtained show the possibility of using this approach in the goal-directed search for plasmid-containing naphthalene-degrading fluorescent pseudomonads in soil. The distribution of the naphthalene catabolism genes in soils contaminated with creosote and petroleum products was also studied.  相似文献   

20.
A yeast strain AEH was isolated from oil contaminated soil and identified by analysis of 18S and 26S ribosomal DNA sequences asPichia anomala. Strain AEH was capable of degrading naphthalene, phenanthrene and chrysene, singly, and benzo(a)pyrene in combination. The yeast degraded 5.36 mg naphthalene l?1 within 2 days, and 5.04 mg phenanthrene l?1 and 1.54 mg chrysene 1?1 within 10 days. When a mixture of the four polycyclic aromatic hydrocarbons (PAHs) was treated at a concentration between 2.98 mg l?1 and 6.89 mg l?1, degradation rates were delayed for naphthalene and phenanthrene (3.79 mg l?1 and, 4.20 mg l?1 within 10 days, respectively), but enhanced for chrysene and benzo(a)pyrene (3.37 mg l?1 and, 1.91 mg l?1 within 10 days, respectively). In a binary system, all of the other 3 PAHs could be utilized as the carbon source for the cometabolic degradation of benzo(a)pyrene with naphthale ne as the best one.  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