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.
The regulation of naphthalene and 1-naphthol metabolism in a Rhodococcus sp. (NCIMB 12038) has been investigated. The microorganism utilizes separate pathways for the degradation of these compounds, and they are regulated independently. Naphthalene metabolism was inducible, but not by salicylate, and 1-naphthol metabolism, although constitutive, was also repressed during growth on salicylate. The biochemistry of naphthalene degradation in this strain was otherwise identical to that found in Pseudomonas putida, with salicylate as a central metabolite and naphthalene initially being oxidized via a naphthalene dioxygenase enzyme to cis-(1R,2S)-1,2-dihydroxy-1,2-dihydronaphthalene (naphthalene cis-diol). A dioxygenase enzyme was not expressed under growth conditions which facilitate 1-naphthol degradation. However, biotransformations with indene as a substrate suggested that a monooxygenase enzyme may be involved in the degradation of this compound. Indole was transformed to indigo by both naphthalene-grown NCIMB 12038 and by cells grown in the absence of an inducer. Therefore, the presence of a naphthalene dioxygenase enzyme activity was not necessary for this reaction. Thus, the biotransformation of indole to indigo may be facilitated by another type of enzyme (possibly a monooxygenase) in this organism.  相似文献   

2.
The physiological role of NahW, the second salicylate hydroxylase of Pseudomonas stutzeri AN10, has been analysed by gene mutation and further complementation. When grown on naphthalene as a unique carbon and energy source, the nahW mutant showed a strong decrease in salicylate hydroxylase activity when compared with the wild-type strain, exhibited lower specific growth rates and accumulated salicylate in culture supernatants. Similarly, lower specific growth rates and salicylate accumulation were observed for the nahW mutant when growth on naphthalene supplemented with succinate or pyruvate. When P. stutzeri AN10 was grown in Luria–Bertani medium in the presence of salicylate, or was cultivated on minimal medium supplemented with salicylate as a unique carbon and energy source, an increase in the lag phase and a decrease in the specific growth rate were observed on increasing the salicylate concentrations, suggesting a plausible toxic effect. This toxic effect of salicylate was much more evident for the nahW mutant than for the wild-type strain. Complementation of the nahW mutant restored all growth parameters. These results indicate that NahW may have two functions in P. stutzeri AN10: (1) to improve its capacity to degrade naphthalene and (2) effectively convert the salicylate produced during naphthalene degradation to tricarboxylic acid cycle intermediates, preventing its toxic effect.  相似文献   

3.
Rhodococcus sp. strain TFB is a metabolic versatile bacterium able to grow on naphthalene as the only carbon and energy source. Applying proteomic, genetic and biochemical approaches, we propose in this paper that, at least, three coordinated but independently regulated set of genes are combined to degrade naphthalene in TFB. First, proteins involved in tetralin degradation are also induced by naphthalene and may carry out its conversion to salicylaldehyde. This is the only part of the naphthalene degradation pathway showing glucose catabolite repression. Second, a salicylaldehyde dehydrogenase activity that converts salicylaldehyde to salicylate is detected in naphthalene‐grown cells but not in tetralin‐ or salicylate‐grown cells. Finally, we describe the chromosomally located nag genes, encoding the gentisate pathway for salicylate conversion into fumarate and pyruvate, which are only induced by salicylate and not by naphthalene. This work shows how biodegradation pathways in Rhodococcus sp. strain TFB could be assembled using elements from different pathways mainly because of the laxity of the regulatory systems and the broad specificity of the catabolic enzymes.  相似文献   

4.
A bacterium, designated F199, utilized toluene, naphthalene, dibenzothiophene, salicylate, benzoate, p-cresol, and all isomers of xylene as a sole carbon and energy source. This bacterium was isolated from Middendorf sediments, a Cretaceous age formation that underlies the Southeast Coastal Plain in South Carolina, at a depth of approximately 410 m. F199 is a gram-positive, irregular-shaped bacterium that has a varied cell morphology that is dependent on culture medium type and growth stage. F199 required microaerobic conditions (40 to 80 μM O2) for growth on hydrocarbons, glucose, acetate, and lactate in mineral salts medium but not for growth on rich media. [14C]naphthalene mineralization by F199 was induced by either naphthalene or toulene; however, [14C]toluene mineralization by this strain was induced by toluene but not naphthalene. F199 was also found to harbor two plasmids larger than 100 kb. Restricted F199 plasmid and genomic DNA did not hybridize with toluene (pWW0) or naphthalene (NAH7) catabolic plasmid DNA probes. The presence in the Middendorf formation of bacteria with the capacity for degrading a variety of aromatic compounds suggests that indigenous microorganisms may have potential for in situ degradation of organic contaminants.  相似文献   

5.
The majority of polycyclic aromatic hydrocarbons (PAHs) sorb strongly to soil organic matter posing a complex barrier to biodegradation. Biosurfactants can increase soil-sorbed PAHs desorption, solubilisation, and dissolution into the aqueous phase, which increases the bioavailability of PAHs for microbial metabolism. In this study, biosurfactants, carbon sources, and metabolic pathway inducers were tested as stimulators of microorganism degradation. Phenanthrene served as a model PAH and Pseudomonas putida ATCC 17484 was used as the phenanthrene degrading microorganism for the liquid solutions and soil used in this investigation. Bench-scale trials demonstrated that the addition of rhamnolipid biosurfactant increases the apparent aqueous solubility of phenanthrene, and overall degradation by at least 20% when combined with salicylate or glucose in liquid solution, when compared to solutions that contained salicylate or glucose with no biosurfactant. However, salicylate addition, with no biosurfactant addition, increased the total degradation of phenanthrene 30% more than liquid systems with only biosurfactant addition. In soil slurries, small amounts of biosurfactant (0.25 g/L) showed a significant increase in total removal when only biosurfactant was added. In soil slurries containing salicylate, the effects of biosurfactant additions were negligible as there was greater than 90% removal, regardless of the biosurfactant concentration. The results of experiments performed in this study provide further evidence that an in situ enhancement strategy for phenanthrene degradation could focus on providing additional carbon substrates to induce metabolic pathway catabolic enzyme production, if degradation pathway intermediates are known.  相似文献   

6.
The objectives of this work were (1) to demonstrate how the chemostat approach could be modified to allow determination of kinetic parameters for a sparingly soluble, volatile substrate such as naphthalene and (2) to examine the influence of the interactions of various nutrients on possible growth-inhibitory effects of naphthalene. Pseudomonas putida G7 was used as a model naphthalene-degrading microorganism. Naphthalene was found to be toxic to P. putida G7 in the absence of a nitrogen source or oxygen. The death rate of cells grown on minimal medium plus naphthalene and then exposed to naphthalene under anoxic conditions was higher than that observed under oxic conditions in the absence of a nitrogen source. The presence of necessary nutrients for the biodegradation of PAH compounds is indicated to be important for the survival of microorganisms that are capable of PAH degradation. The amounts of ammonia and oxygen necessary for naphthalene biodegradation and for suppression of naphthalene toxicity were calculated from growth yield coefficients. A chemostat culture of P. putida G7 using naphthalene as a carbon and energy source was accomplished by using a feed augmented with a methanol solution of naphthalene so as to provide sufficient growth to allow accurate evaluation of kinetic parameters. When naphthalene was the growth-limiting substrate, the degradation of naphthalene followed Monod kinetics. Maximum specific growth rate (micrometer) and Monod constant (Ks) were 0.627 +/- 0.007 h-1 and 0.234 +/- 0.0185 mg/L, respectively. The evaluation of biodegradation parameters will allow a mathematical model to be applied to predict the long-term behavior of PAH compounds in soil when combined with PAH transport parameters.  相似文献   

7.
A naphthalene (Nap) and salicylate (Sal) degrading microorganism, Pseudomonas putida RKJ1, is chemotactic towards these compounds. This strain carries a 83 kb plasmid. A 25 kb EcoRI fragment of the plasmid contains the genes responsible for Nap degradation through Sal. RKJ5, the plasmid-cured derivative of RKJ1, is neither capable of degradation nor is chemotactic towards Nap or Sal. The recombinant plasmid pRKJ3, which contained a 25 kb EcoRI fragment, was transferred back into the plasmid-free wild-type strain RKJ5, and the transconjugant showed both degradation and chemotaxis. The recombinant plasmid pRKJ3 was also transferred into motile, plasmid-free P. putida KT2442. The resulting transconjugant (RKJ15) showed chemotaxis towards both Nap and Sal. Two mutant strains carrying deletions in pRKJ3 (in KT2442) with phenotypes Nap- Sal+ and Nap- Sal-, were also tested for chemotaxis. It was found that the Nap- Sal+ mutant strain showed chemotaxis towards Sal only, whereas the Nap- Sal- mutant strain is non-chemotactic towards both the compounds. These results suggest that the metabolism of Nap and Sal may be required for the chemotactic activity.  相似文献   

8.
Rhodococcus sp. strain B4, isolated from a soil sample contaminated with polycyclic aromatic hydrocarbons, grows with naphthalene as the sole source of carbon and energy. Salicylate and gentisate were identified as intermediates in the catabolism of naphthalene. In contrast to the well-studied catabolic pathway encoded by the NAH7 plasmid of Pseudomonas putida, salicylate does not induce the genes of the naphthalene-degradative pathway in Rhodococcus sp. strain B4. The key enzymes of naphthalene degradation in Rhodococcus sp. strain B4 have unusual cofactor requirements. The 1,2-dihydroxynaphthalene oxygenase activity depends on NADH and the salicylate 5-hydroxylase requires NADPH, ATP, and coenzyme A.  相似文献   

9.
Naphthalene oxidation by a parent and a mutant strain of Pseudomonas putida was studied. The parent strain contained a plasmid NPL-1 which controlled oxidation of naphthalene to salicylic acid and was capable of oxidizing salicylate. The mutant strain did not oxidize salicylate because of a mutation in salicylate hydroxylase; it contained also a mutant plasmid NPL-41 which determined constitutive synthesis of naphthalene oxygenase. Salicylic acid which accumulated as a product of naphthalene catabolism in the cultural broth of the wild strain was found to undergo further oxidation by the population of growing cells. The content of salicylic acid in the cultural broth of the mutant strain reached maximum and then remained constant. An anion-exchange resin was tested in order to prevent the inhibition of naphthalene oxygenase by salicylate and to increase the yield of salicylic acid. The transmissible character of the mutant plasmid NPL-41 makes it possible, with the aid of conjugation, to construct Pseudomonas strains which would oxidize naphthalene to salicylic acid without further degradation of this compound.  相似文献   

10.
Streptomyces peucetius var. caesius produces a family of secondary metabolites called anthracyclines. Production of these compounds is negatively affected in the presence of glucose, galactose, and lactose, but the greatest effect is observed under conditions of excess glucose. Other carbon sources, such as arabinose or glutamate, show either no effect or stimulate production. Among the carbon sources that negatively affect anthracycline production, glucose is consumed in greater concentrations. We determined glucose and galactose transport in S. peucetius var. caesius and in a mutant of this strain whose anthracycline production is insensitive to carbon catabolite repression (CCR). In the original strain, incorporation of glucose and galactose was stimulated when the microorganism was grown in media containing these sugars, although we also observed basal galactose incorporation. Both the induced and the basal incorporation of galactose were suppressed when the microorganism was grown in the presence of glucose. Furthermore, adding glucose directly during the transport assay also inhibited galactose incorporation. In the mutant strain, we observed a reduction in both glucose (48%) and galactose (81%) incorporation compared to the original. Galactose transport in this mutant showed reduced sensitivity to the negative effect of glucose; however, it was still sensitive to inhibition. The deficient transport of these sugars, as well as CCR sensitivity to glucose in this mutant was corrected when the mutant was transformed with the SCO2127 region of the Streptomyces coelicolor genome. Our results support a role for glucose as the most easily utilized carbon source capable of exerting the greatest repression on anthracycline biosynthesis. In consequence, glucose also prevented the repressive effect of galactose by suppressing its incorporation. This suggests the participation of an integral regulatory system, which is initiated by an increase in incorporation of repressive sugars and their metabolism as a prerequisite for establishing the phenomenon of CCR in S. peucetius var. caesius.  相似文献   

11.
The process of naphthalene degradation by indigenous, introduced, and transconjugant strains was studied in laboratory soil microcosms. Conjugation transfer of catabolic plasmids was demonstrated in naphthalene-contaminated soil. Both indigenous microorganisms and an introduced laboratory strain BS394 (pNF142::TnMod-OTc) served as donors of these plasmids. The indigenous bacterial degraders of naphthalene isolated from soil were identified as Pseudomonas putida and Pseudomonas fluorescens. The frequency of plasmid transfer in soil was 10(-5)-10(-4) per donor cell. The activity of the key enzymes of naphthalene biodegradation in indigenous and transconjugant strains was studied. Transconjugant strains harboring indigenous catabolic plasmids possessed high salicylate hydroxylase and low catechol-2,3-dioxygenase activities, in contrast to indigenous degraders, which had a high level of catechol-2,3-dioxygenase activity and a low level of salicylate hydroxylase. Naphthalene degradation in batch culture in liquid mineral medium was shown to accelerate due to cooperation of the indigenous naphthalene degrader P. fluorescens AP1 and the transconjugant strain P. putida KT2442 harboring the indigenous catabolic plasmid pAP35. The role of conjugative transfer of naphthalene biodegradation plasmids in acceleration of naphthalene degradation was demonstrated in laboratory soil microcosms.  相似文献   

12.
Zeng G  Fu H  Zhong H  Yuan X  Fu M  Wang W  Huang G 《Biodegradation》2007,18(3):303-310
Strengthened biodegradation is one of the key means to treat surfactant pollution in environment, and microorganism and surfactant have significant effects on degradation. In this paper, co-degradation of CTAB, Triton X-100, SDS and rhamnolipid with glucose by Pseudomonas aeruginosa, Bacillus subtilis and compost microorganisms in liquid culture media, as well as the degradation of rhamnolipid in compost were investigated. The results showed that CTAB was recalcitrant to degrade by the three microorganisms and it also inhibited microorganisms from utilizing readily degradable carbon source. Non-ionic surfactant Triton X-100 could also hardly be degraded, but it was not toxic to microorganisms and would not inhibit the growth of the microorganisms. Anion surfactant SDS had no toxicity to microorganisms and could be co-degraded as carbon source with glucose. Biosurfactant rhamnolipid was a kind of particular surfactant, which had no toxicity and could be degraded by Bacillus subtilis and compost microorganisms, while it could not be utilized by its producing bacterium Pseudomonas aeruginosa. Among these three bacteria, the compost consortium had the strongest degradation capacity on the tested surfactants due to their microorganisms’ diversity. In compost matrix rhamnolipid could be degraded during composting, but not preferentially utilized.  相似文献   

13.
The evolution of microbial catabolic enzymes cannot keep pace with the rapid introduction of novel compounds into the environment. These new synthetic compounds that are slowly biodegradable or non-biodegradable are known as recalcitrant compounds, and range from simple halogenated hydrocarbons to complex polymers. Recalcitrant compounds can be made biodegradable by developing microorganisms capable of degrading the compound and by treating the compound to make it more conducive to mirobial attack. Many factors contribute to recalcitrance. The organism may lack the necessary genetic information. The organism can acquire this information by plasmid transfer or de novo enzyme synthesis. Plasmids have been characterized that degrade or transform antibiotics, pesticides, and hydrocarbons. By the use of chemostat techniques or chemical mutagens, organisms have been shown to synthesize de novo enzymes. The compound may be too large to enter the cell, or a transport system may not exist to transport it across the membrane. The compound may be insoluble, either as a solid or a liquid, and the microorganism may lack the proper nutrients. Recalcitrant compounds can be oxygenated prior to degradation, in the presence of a readily assimilable carbon source. In the absence of the assimilable carbon source, the recalcitrant compound is not degraded, or only very slowly. Examples of such co-oxidative metabolism are alkane and lignin degradation. Polymers, particularly synthetic ones, are prime examples of difficult-to-degrade compounds. The initial rate of polymer degradation follows a Freundlich or modified Langmuir isotherm rather than Michaelis-Menten kinetics. Microorganisms can irreversibly bind to solid surfaces by various methods. Soil microorganisms have been found to degrade styrene monomers and dimers. Polystyrene has been shown to be biodegradable by 14CO2 evolution but at a very slow rate. In car tyres, styrene as a copolymer of butadiene is co-metabolized in the presence of other assimilable carbon sources.  相似文献   

14.
Li S  Li X  Zhao H  Cai B 《Microbiological research》2011,166(8):643-653
The classical salicylaldehyde dehydrogenases found in naphthalene-degrading bacteria are denoted as NahF. In addition to NahF, NahV, and its corresponding gene nahV, were found here in multiple naphthalene-degrading bacteria isolated from industrial wastewater polluted with polycyclic aromatic hydrocarbons (PAHs). In this study, we described for the first time the biological function and regulation model of NahV for the mineralization of naphthalene by P. putida ND6 via the construction of nahF-, nahV- and regulatory gene nahR-deficient strains. The two mutants of salicylaldehyde dehydrogenase genes and wild-type Pseudomonas ND6 were compared with respect to growth rate, naphthalene degradation efficiency, protein expression level, and salicylaldehyde dehydrogenase activity. The data showed that the presence of NahV conferred a physiological advantage on P. putida ND6 for the catabolism of naphthalene in the presence of NahF. NahV could facilitate naphthalene degradation by increasing total salicylaldehyde dehydrogenase activity when both dehydrogenases are present and it could replace the function of NahF when nahF gene is deleted or mutated, thus ensuring mutants could survive in naphthalene-containing environments. To investigate regulation model of NahV, we detected the expression levels and salicylaldehyde dehydrogenase activity in the wild-type and the nahR mutant strains following cultivation in the presence of glucose±salicylate. The data demonstrated that just like the classical salicylaldehyde dehydrogenases, NahF, NahV was induced by salicylate in the presence of NahR.  相似文献   

15.
16.
The enzymes of naphthalene metabolism are induced in Pseudomonas putida ATCC 17484, PpG7, NCIB 9816, and PG and in Pseudomonas sp. ATCC 17483 during growth on naphthalene or salicylate; 2-aminobenzoate is a gratuitous inducer of these enzymes. The meta-pathway enzymes of catechol metabolism are induced in ATCC 17483 and PPG7 during growth on naphthalene or salicylate or during growth in the presence of 2-aminobenzoate, but in ATCC 17484 and NCIB 9816 the ortho-pathway enzymes of catechol metabolism are induced during growth on naphthalene or salicylate. 2-Aminobenzoate does not induce any enzymes of catechol metabolism in the latter two organisms. In Pseudomonas PG the meta-pathway enzymes are present at high levels under all conditions of growth, but this organism and PpG7 can induce ortho-pathway enzymes during naphthalene or salicylate metabolism. Salicylate appears to be the inducer of the enzymes of naphthalene metabolism in all of the organisms studied and, where they are inducible, of the meta-pathway enzymes, but the properties of Pseudomonas PG suggest that separate, regulatory systems may exist.  相似文献   

17.
Pseudomonas sp. strain PP2 isolated in our laboratory efficiently metabolizes phenanthrene at 0.3% concentration as the sole source of carbon and energy. The metabolic pathways for the degradation of phenanthrene, benzoate and p-hydroxybenzoate were elucidated by identifying metabolites, biotransformation studies, oxygen uptake by whole cells on probable metabolic intermediates, and monitoring enzyme activities in cell-free extracts. The results obtained suggest that phenanthrene degradation is initiated by double hydroxylation resulting in the formation of 3,4-dihydroxyphenanthrene. The diol was finally oxidized to 2-hydroxymuconic semialdehyde. Detection of 1-hydroxy-2-naphthoic acid, alpha-naphthol, 1,2-dihydroxy naphthalene, and salicylate in the spent medium by thin layer chromatography; the presence of 1,2-dihydroxynaphthalene dioxygenase, salicylaldehyde dehydrogenase and catechol-2,3-dioxygenase activity in the extract; O(2) uptake by cells on alpha-naphthol, 1,2-dihydroxynaphthalene, salicylaldehyde, salicylate and catechol; and no O(2) uptake on o-phthalate and 3,4-dihydroxybenzoate supports the novel route of metabolism of phenanthrene via 1-hydroxy-2-naphthoic acid --> [alpha-naphthol] --> 1,2-dihydroxy naphthalene --> salicylate --> catechol. The strain degrades benzoate via catechol and cis,cis-muconic acid, and p-hydroxybenzoate via 3,4-dihydroxybenzoate and 3-carboxy- cis,cis-muconic acid. Interestingly, the culture failed to grow on naphthalene. When grown on either hydrocarbon or dextrose, the culture showed good extracellular biosurfactant production. Growth-dependent changes in the cell surface hydrophobicity, and emulsification activity experiments suggest that: (1) production of biosurfactant was constitutive and growth-associated, (2) production was higher when cells were grown on phenanthrene as compared to dextrose and benzoate, (3) hydrocarbon-grown cells were more hydrophobic and showed higher affinity towards both aromatic and aliphatic hydrocarbons compared to dextrose-grown cells, and (4) mid-log-phase cells were significantly (2-fold) more hydrophobic than stationary phase cells. Based on these results, we hypothesize that growth-associated extracellular biosurfactant production and modulation of cell surface hydrophobicity plays an important role in hydrocarbon assimilation/uptake in Pseudomonas sp. strain PP2.  相似文献   

18.
Bacterial strain M213 was isolated from a fuel oil-contaminated soil in Idaho, USA, by growth on naphthalene as a sole source of carbon, and was identified as Rhodococcus opacus M213 by 16S rDNA sequence analysis and growth on substrates characteristic of this species. M213 was screened for growth on a variety of aromatic hydrocarbons, and growth was observed only on simple 1 and 2 ring compounds. No growth or poor growth was observed with chlorinated aromatic compounds such as 2,4-dichlorophenol and chlorobenzoates. No growth was observed by M213 on salicylate, and M213 resting cells grown on naphthalene did not attack salicylate. In addition, no salicylate hydroxylase activity was detected in cell free lysates, suggesting a pathway for naphthalene catabolism that does not pass through salicylate. Enzyme assays indicated induction of catechol 1,2-dioxygenase and catechol 2,3-dioxygenase on different substrates. Total DNA from M213 was screened for hybridization with a variety of genes encoding catechol dioxygenases, but hybridization was observed only with catA (encoding catechol 1,2-dioxygenase) from R. opacus 1CP and edoD (encoding catechol 2,3-dioxygenase) from Rhodococcus sp. I1. Plasmid analysis indicated the presence of two plasmids (pNUO1 and pNUO2). edoD hybridized to pNUO1, a very large (approximately 750 kb) linear plasmid.  相似文献   

19.
Pseudomonas putida CSV86, a plasmid-free strain possessing capability to transfer the naphthalene degradation property, has been explored for its metabolic diversity through genome sequencing. The analysis of draft genome sequence of CSV86 (6.4 Mb) revealed the presence of genes involved in the degradation of naphthalene, salicylate, benzoate, benzylalcohol, p-hydroxybenzoate, phenylacetate and p-hydroxyphenylacetate on the chromosome thus ensuring the stability of the catabolic potential. Moreover, genes involved in the metabolism of phenylpropanoid and homogentisate, as well as heavy metal resistance, were additionally identified. Ability to grow on vanillin, veratraldehyde and ferulic acid, detection of inducible homogentisate dioxygenase and growth on aromatic compounds in the presence of heavy metals like copper, cadmium, cobalt and arsenic confirm in silico observations reflecting the metabolic versatility. In silico analysis revealed the arrangement of genes in the order: tRNAGly, integrase followed by nah operon, supporting earlier hypothesis of existence of a genomic island (GI) for naphthalene degradation. Deciphering the genomic architecture of CSV86 for aromatic degradation pathways and identification of elements responsible for horizontal gene transfer (HGT) suggests that genetic bioaugmentation strategies could be planned using CSV86 for effective bioremediation.  相似文献   

20.
Pseudomonas sp. CSV86, an Indian soil isolate, degrades wide range of aromatic compounds like naphthalene, benzoate and phenylpropanoids, amongst others. Isolate displays the unique and novel property of preferential utilization of aromatics over glucose and co-metabolizes them with organic acids. Interestingly, as compared to other Pseudomonads, strain CSV86 harbours only high-affinity glucokinase pathway (and absence of low-affinity oxidative route) for glucose metabolism. Such lack of gluconate loop might be responsible for the novel phenotype of preferential utilization of aromatics. The genome analysis and comparative functional mining indicated a large genome (6.79 Mb) with significant enrichment of regulators, transporters as well as presence of various secondary metabolite production clusters, suggesting its eco-physiological and metabolic versatility. Strain harbours various integrative conjugative elements (ICEs) and genomic islands, probably acquired through horizontal gene transfer events, leading to genome mosaicity and plasticity. Naphthalene degradation genes are arranged as regulonic clusters and found to be part of ICECSV86nah. Various eco-physiological properties and absence of major pathogenicity and virulence factors (risk group-1) in CSV86 suggest it to be an ideal candidate for bioremediation. Further, strain can serve as an ideal chassis for metabolic engineering to degrade various xenobiotics preferentially over simple carbon sources for efficient remediation.  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