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 62 毫秒
1.
Five naphthalene- and salicylate-utilizing Pseudomonas putida strains cultivated for a long time on phenanthrene produced mutants capable of growing on this substrate and 1-hydroxy-2-naphthoate as the sole sources of carbon and energy. The mutants catabolize phenanthrene with the formation of 1-hydroxy-2-naphthoate, 2-hydroxy-1-naphthoate, salicylate, and catechol. The latter products are further metabolized by the meta- and ortho-cleavage pathways. In all five mutants, naphthalene and phenanthrene are utilized with the involvement of plasmid-born genes. The acquired ability of naphthalene-degrading strains to grow on phenanthrene is explained by the fact that the inducible character of the synthesis of naphthalene dioxygenase, the key enzyme of naphthalene and phenanthrene degradation, becomes constitutive.  相似文献   

2.
Five naphthalene- and salicylate-utilizing Pseudomonas putida strains cultivated for a long time on phenanthrene produced mutants capable of growing on this substrate and 1-hydroxy-2-naphthoate as the sole sources of carbon and energy. The mutants catabolize phenanthrene with the formation of 1-hydroxy-2-naphthoate, 2-hydroxy-1-naphthoate, salicylate, and catechol. The latter products are further metabolized by the meta- and ortho-cleavage pathways. In all five mutants, naphthalene and phenanthrene are utilized with the involvement of plasmid-borne genes. The acquired ability of naphthalene-degrading strains to grow on phenanthrene is explained by the fact that the inducible character of the synthesis of naphthalene dioxygenase, the key enzyme of naphthalene and phenanthrene degradation, becomes constitutive.  相似文献   

3.
In the bacterial degradation of polycyclic aromatic hydrocarbons (PAHs), salicylate hydroxylases catalyze essential reactions at the junction between the so-called upper and lower catabolic pathways. Unlike the salicylate 1-hydroxylase from pseudomonads, which is a well-characterized flavoprotein, the enzyme found in sphingomonads appears to be a three-component Fe-S protein complex, which so far has not been characterized. Here, the salicylate 1-hydroxylase from Sphingomonas sp. strain CHY-1 was purified, and its biochemical and catalytic properties were characterized. The oxygenase component, designated PhnII, exhibited an α3β3 heterohexameric structure and contained one Rieske-type [2Fe-2S] cluster and one mononuclear iron per α subunit. In the presence of purified reductase (PhnA4) and ferredoxin (PhnA3) components, PhnII catalyzed the hydroxylation of salicylate to catechol with a maximal specific activity of 0.89 U/mg and showed an apparent Km for salicylate of 1.1 ± 0.2 μM. The hydroxylase exhibited similar activity levels with methylsalicylates and low activity with salicylate analogues bearing additional hydroxyl or electron-withdrawing substituents. PhnII converted anthranilate to 2-aminophenol and exhibited a relatively low affinity for this substrate (Km, 28 ± 6 μM). 1-Hydroxy-2-naphthoate, which is an intermediate in phenanthrene degradation, was not hydroxylated by PhnII, but it induced a high rate of uncoupled oxidation of NADH. It also exerted strong competitive inhibition of salicylate hydroxylation, with a Ki of 0.68 μM. The properties of this three-component hydroxylase are compared with those of analogous bacterial hydroxylases and are discussed in light of our current knowledge of PAH degradation by sphingomonads.  相似文献   

4.
In cell extracts of Pseudaminobacter salicylatoxidans strain BN12, an enzymatic activity was detected which converted salicylate in an oxygen-dependent but NAD(P)H-independent reaction to a product with an absorbance maximum at 283 nm. This metabolite was isolated, purified, and identified by mass spectrometry and (1)H and (13)C nuclear magnetic resonance spectroscopy as 2-oxohepta-3,5-dienedioic acid. This metabolite could be formed only by direct ring fission of salicylate by a 1,2-dioxygenase reaction. Cell extracts from P. salicylatoxidans also oxidized 5-aminosalicylate, 3-, 4-, and 5-chlorosalicylate, 3-, 4-, and 5-methylsalicylate, 3- and 5-hydroxysalicylate (gentisate), and 1-hydroxy-2-naphthoate. The dioxygenase was purified and shown to consist of four identical subunits with a molecular weight of about 45,000. The purified enzyme showed higher catalytic constants with gentisate or 1-hydroxy-2-naphthoate than with salicylate. It was therefore concluded that P. salicylatoxidans synthesized a gentisate 1,2-dioxygenase with an extraordinary substrate range, which also allowed the oxidation of salicylate.  相似文献   

5.
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   

6.
Burkholderia sp. TNFYE-5 was isolated from soil for the ability to grow on phenanthrene as sole carbon and energy source. Unlike most other phenanthrene-degrading bacteria, TNFYE-5 was unable to grow on naphthalene. Growth substrate range experiments coupled with the ring-cleavage enzyme assay data suggest that TNFYE-5 initially metabolizes phenanthrene to 1-hydroxy-2-naphthoate with subsequent degradation through the phthalate and protocatechuate and beta-ketoadipate pathway. A metabolite in the degradation of naphthalene by TNFYE-5 was isolated by high-pressure liquid chromatography (HPLC) and was identified as salicylate by UV-visible spectral and gas chromatography-mass spectrometry analyses. Thus, the inability to degrade salicylate is apparently one major reason for the incapability of TNFYE-5 to grow on naphthalene.  相似文献   

7.
The gene coding for a dioxygenase with the ability to cleave salicylate by a direct ring fission mechanism to 2-oxohepta-3,5-dienedioic acid was cloned from Pseudaminobacter salicylatoxidans strain BN12. The deduced amino acid sequence encoded a protein with a molecular mass of 41,176 Da, which showed 28 and 31% sequence identity, respectively, to a gentisate 1,2-dioxygenase from Pseudomonas alcaligenes NCIMB 9867 and a 1-hydroxy-2-naphthoate 1,2-dioxygenase from Nocardioides sp. KP7. The highest degree of sequence identity (58%) was found to a presumed gentisate 1,2-dioxygenase from Corynebacterium glutamicum. The enzyme from P. salicylatoxidans BN12 was heterologously expressed in Escherichia coli and purified as a His-tagged enzyme variant. The purified enzyme oxidized in addition to salicylate, gentisate, 5-aminosalicylate, and 1-hydroxy-2-naphthoate also 3-amino- and 3- and 4-hydroxysalicylate, 5-fluorosalicylate, 3-, 4-, and 5-chlorosalicylate, 3-, 4-, and 5-bromosalicylate, 3-, 4-, and 5-methylsalicylate, and 3,5-dichlorosalicylate. The reactions were analyzed by high pressure liquid chromatography/mass spectrometry, and the reaction products were tentatively identified. For comparison, the putative gentisate 1,2-dioxygenase from C. glutamicum was functionally expressed in E. coli and shown to convert gentisate but not salicylate or 1-hydroxy-2-naphthoate.  相似文献   

8.
Plasmid vector and allelic exchange mutagenesis systems were established for the genetic analysis of a phenanthrene-degrading mycobacterial strain, Mycobacterium sp. EPa45. Successful application of these systems revealed the necessity of the EPa45 phdI gene for the degradation of 1-hydroxy-2-naphthoate, which has been proposed to be an intermediate product in the degradation pathway of phenanthrene.  相似文献   

9.
A putative gentisate 1,2-dioxygenase was encoded in the dibenzothiophene degradation gene cluster (dbd) from Xanthobacter polyaromaticivorans 127W. The deduced amino acid sequence showed high sequence similarity with gentisate dioxygenases from Pseudomonas alcaligenes (AAD49427, 65% identical), Bradyrhizobium japonicum (NP_766750, 64%), and P. aeruginosa (ZP_00135722, 54%), and moderate similarity with 1-hydroxy-2-naphthoate dioxygenase from Nocardioides sp. KP7 (BAA31235, 33%) and salicylate dioxygenase from Pseudaminobacter salicylatoxidans (AAQ91293, 33%). The enzyme, GDOxp, was heterologously produced in Escherichia coli and purified to homogeneity. GDOxp formed a tetramer and exhibited high dioxygenase activity against 1,4-dihydroxy 2-naphthoate as well as gentisate, suggesting unusually broad substrate specificity. GDOxp easily released ferrous ion under unfavorable temperature and pH conditions to become an inactive monomer protein. An inactive monomer protein can reconstitute a tetramer structure and restore enzyme activity in a cooperative manner upon the addition of ferrous ion. Chymotryptic digestion and protein truncation experiments suggested that the N-terminal region is important for the tetramerization of GDOxp.  相似文献   

10.
Sphingomonas sp. strain RW1, when grown in salicylate-salts medium, synthesized the enzymes for the degradation of dibenzofuran. The reaction subsequent tometa cleavage of the first benzene ring was found to be catalyzed by two isofunctional hydrolases, H1 and H2, which were purified by chromatography on anion exchange, hydrophobic interaction and gel filtration media. Each enzyme was able to hydrolze 2-hydroxy-6-oxo-6-(2-hydroxyphenyl)hexa-2,4-dienoate and 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate to produce salicylate and benzoate, respectively. SDS/PAGE of each purified enzyme showed a single band ofM r 31 000 (H1) or 29 000 (H2). The N-terminal amino acid sequences of the two proteins showed 50% homology.Abbreviations DHB 2,3-dihydroxybiphenyl - DSM German Culture Collection (Braunschweig) - FPLC protein liquid chromatograph(y) - HOHPDA 2-hydroxy-6-oxo-6-(2-hydroxyphenyl)hexa-2,4-dienoate - HOPDA 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate - THB 2,2,3-trihydroxybiphenyl  相似文献   

11.
The chrysene-degrading bacterium Pseudoxanthomonas sp. PNK-04 was isolated from a coal sample. Three novel metabolites, hydroxyphenanthroic acid, 1-hydroxy-2-naphthoic acid and salicylic acid, were identified by TLC, HPLC and MS. Key enzyme activities, namely 1-hydroxy-2-naphthoate hydroxylase, 1,2-dihydroxynaphthalene dioxygenase, salicylaldehyde dehydrogenase and catechol-1,2-dioxygenase, were noted in the cell-free extract. These results suggest that chrysene is catabolized via hydroxyphenanthroic acid, 1-hydroxy-2-naphthoic acid, salicylic acid and catechol. The terminal aromatic metabolite, catechol, is then catabolized by catechol-1,2-dioxygenase to cis,cis-muconic acid, ultimately forming TCA cycle intermediates. Based on these studies, the proposed catabolic pathway for chrysene degradation by strain PNK-04 is chrysene → hydroxyphenanthroic acid → 1-hydroxy-2-naphthoic acid → 1,2-dihydroxynaphthalene → salicylic acid → catechol →cis,cis-muconic acid.  相似文献   

12.
2-Naphthoate catabolic pathway in Burkholderia strain JT 1500.   总被引:1,自引:0,他引:1       下载免费PDF全文
Burkholderia strain (JT 1500), able to use 2-naphthoate as the sole source of carbon, was isolated from soil. On the basis of growth characteristics, oxygen uptake experiments, enzyme assays, and detection of intermediates, a degradation pathway of 2-naphthoate is proposed. The features of this pathway are convergent with those for phenanthrene. We propose a pathway for the conversion of 2-naphthoate to 1 mol (each) of pyruvate, succinate, and acetyl coenzyme A and 2 mol of CO2. During growth in the presence of 2-naphthoate, six metabolites were detected by thin-layer chromatography, high-performance liquid chromatography, and spectroscopy. 1-Hydroxy-2-naphthoate accumulated in the culture broth during growth on 2-naphthoate. Also, the formation of 2'-carboxybenzalpyruvate, phthalaldehydate, phthalate, protocatechuate, and beta-carboxy-cis,cis-muconic acid was demonstrated. (1R,2S)-cis-1,2-Dihydro-1,2-dihydroxy-2-naphthoate was thus considered an intermediate between 2-naphthoate and 1-hydroxy-2-naphthoate, but it was not transformed by whole cells or their extracts. We conclude that this diol is not responsible for the formation of 1-hydroxy-2-naphthoate from 2-naphthoate but that one of the other three diastereomers is not eliminated as a potential intermediate for a dehydration reaction.  相似文献   

13.
Aims:  The metabolism of phenanthrene and anthracene by a moderate thermophilic Nocardia otitidiscaviarum strain TSH1 was examined.
Methods and Results:  When strain TSH1 was grown in the presence of anthracene, four metabolites were identified as 1,2-dihydroxy-1,2-dihydroanthracene, 3-(2-carboxyvinyl)naphthalene-2-carboxylic acid, 2,3-dihydroxynaphthalene and benzoic acid using gas chromatography-mass spectrometry (GC-MS), reverse phase-high performance liquid chromatography (RP-HPLC) and thin-layer chromatography (TLC). Degradation studies with phenanthrene revealed 2,2'-diphenic acid, phthalic acid, 4-hydroxyphenylacetic acid, o -hydroxyphenylacetic acid, benzoic acid, a phenanthrene dihydrodiol, 4-[1-hydroxy(2-naphthyl)]-2-oxobut-3-enoic acid and 1-hydroxy-2-naphthoic acid (1H2NA), as detectable metabolites.
Conclusions:  Strain TSH1 initiates phenanthrene degradation via dioxygenation at the C-3 and C-4 or at C-9 and C-10 ring positions. Ortho -cleavage of the 9,10-diol leads to formation of 2,2'-diphenic acid. The 3,4-diol ring is cleaved to form 1H2NA which can subsequently be degraded through o -phthalic acid pathway. Benzoate does not fit in the previously published pathways from mesophiles. Anthracene metabolism seems to start with a dioxygenation at the 1 and 2 positions and ortho -cleavage of the resulting diol. The pathway proceeds probably through 2,3-dicarboxynaphthalene and 2,3-dihydroxynaphthalene. Degradation of 2,3-dihydroxynaphthalene to benzoate and transformation of the later to catechol is a possible route for the further degradation of anthracene.
Significance and Impact of the Study:  For the first time, metabolism of phenanthrene and anthracene in a thermophilic Nocardia strain was investigated.  相似文献   

14.
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   

15.
A 25-kb DNA SalI fragment cloned from the chromosomal DNA of Pseudomonas putida OUS82, which utilizes phenanthrene (Phn+) and naphthalene (Nah+), carried all of the genes necessary for upper naphthalene catabolism. Cosmid recombinant pIP7 complemented both the Nah- and Phn- defects of OUS8211 (Trp-Nah-Phn-Sal+[salicylate utilizing]Hna+[1-hydroxy-2-naphthoate utilizing]) and only the Phn- defect of OUS8212 (Trp-Nah-Phn-Sal-Hna+). The results indicate that strain OUS82 uses different pathways after o-hydroxycarboxylic aromatics in the catabolism of naphthalene and phenanthrene.  相似文献   

16.
Pseudomonas paucimobilis Q1 originally isolated as biphenyl degrading organism (Furukawa et al. 1983), was shown to grow with naphthalene. After growth with biphenyl or naphthalene the strain synthesized the same enzyme for the ring cleavage of 2,3-dihydroxybiphenyl or 1,2-dihydroxynaphthalene. The enzyme, although characterized as 2,3-dihydroxybiphenyl dioxygenase (Taira et al. 1988), exhibited considerably higher relative activity with 1,2-dihydroxynaphthalene. These results demonstrate that this enzyme can function both in the naphthalene and biphenyl degradative pathway.Abbreviations DHBP dihydroxybiphenyl - DHBPDO 2,3-dihydroxybiphenyl dioxygenase - DHDHNDH 1,2-dihydroxy-1,2-dihydronaphthalene dehydrogenase - DHN 1,2-dihydroxynaphthalene - DHNDO 1,2-dihydroxynaphthalene dioxygenase - HBP cis-2-hydroxybenzalpyruvate - HOPDA 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoate - PCB polychlorinated biphenyl - 2NS naphthalene-2-sulfonic acid  相似文献   

17.
A putative gentisate 1,2-dioxygenase was encoded in the dibenzothiophene degradation gene cluster (dbd) from Xanthobacter polyaromaticivorans 127W. The deduced amino acid sequence showed high sequence similarity with gentisate dioxygenases from Pseudomonas alcaligenes (AAD49427, 65% identical), Bradyrhizobium japonicum (NP_766750, 64%), and P. aeruginosa (ZP_00135722, 54%), and moderate similarity with 1-hydroxy-2-naphthoate dioxygenase from Nocardioides sp. KP7 (BAA31235, 33%) and salicylate dioxygenase from Pseudaminobacter salicylatoxidans (AAQ91293, 33%). The enzyme, GDOxp, was heterologously produced in Escherichia coli and purified to homogeneity. GDOxp formed a tetramer and exhibited high dioxygenase activity against 1,4-dihydroxy 2-naphthoate as well as gentisate, suggesting unusually broad substrate specificity. GDOxp easily released ferrous ion under unfavorable temperature and pH conditions to become an inactive monomer protein. An inactive monomer protein can reconstitute a tetramer structure and restore enzyme activity in a cooperative manner upon the addition of ferrous ion. Chymotryptic digestion and protein truncation experiments suggested that the N-terminal region is important for the tetramerization of GDOxp.  相似文献   

18.
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   

19.
Pseudomonas maltophilia CSV89, a bacterium isolated from soil in our laboratory, grows on 1-naphthoic acid as the sole source of carbon and energy. To elucidate the pathway for degradation of 1-naphthoic acid, the metabolites were isolated from spent medium, purified by TLC, and characterized by gas chromatography-mass spectrometry. The involvement of various metabolites as intermediates in the pathway was established by demonstrating relevant enzyme activities in cell-free extracts, oxygen uptake and transformation of metabolites by the whole cells. The results obtained from such studies suggest that the degradation of 1-naphthoic acid is initiated by double hydroxylation of the aromatic ring adjacent to the one bearing the carboxyl group, resulting in the formation of 1,2-dihydroxy-8-carboxynaphthalene. The resultant diol was oxidized via 3-formyl salicylate, 2-hydroxyisophthalate, salicylate and catechol to TCA cycle intermediates.  相似文献   

20.
Four polycyclic aromatic hydrocarbon (PAH)- degrading bacteria, namely Arthrobacter sulphureus RKJ4, Acidovorax delafieldii P4-1, Brevibacterium sp. HL4 and Pseudomonas sp. DLC-P11, capable of utilizing phenanthrene as the sole source of carbon and energy, were tested for its degradation using radiolabelled phenanthrene. [9-14C]Phenanthrene was incubated with microorganisms containing 100 mg/l unlabelled phenanthrene and the evolution of 14CO2 was monitored: within 18 h of incubation, 30.1, 35.6, 26.5 and 2.1% of the recovered radiolabelled carbon was degraded to 14CO2 by RKJ4, P4-1, HL4 and DLC-P11, respectively. When mixtures of other PAHs such as fluorene, fluoranthene and pyrene, in addition to phenanthrene, were added as additional carbon sources, there was a 36.1 and 20.6% increase in 14CO2 production from [9-14C]phenanthrene in the cases of RKJ4 and HL4, respectively, whereas P4-1 and DLC-P11 did not show any enhancement in 14CO2 production. Although, a combination of many bacteria enhances the degradation of organic compounds, no enhancement in the degradation of [9-14C]phenanthrene was observed in mixed culture involving all four microorganisms together. However, when different PAHs, as indicated above, were used in mixed culture, there was a 68.2% increase in 14CO2 production. In another experiment, the overall growth rate of P4-1 on phenanthrene could be enhanced by adding the non-ionic surfactant Triton X-100, whereas RKJ4, HL4 and DLC-P11 did not show any enhancement in growth. Pathways for phenanthrene degradation were also analysed by thin-layer chromatography, gas chromatography and gas chromatography-mass spectrometry. Common intermediates such as o-phthalic acid and protocatechuic acid were detected in the case of RKJ4 and o-phthalic acid was detected in the case of P4-1. A new intermediate, 1-naphthol, was detected in the cases of HL4 and DLC-P11. HL4 degrades phenanthrene via 1-hydroxy-2-naphthoic acid, 1-naphthol and salicylic acid, whereas DLC-P11 degrades phenanthrene via the formation of 1-hydroxy-2-naphthoic acid, 1-naphthol and o-phthalic acid. Both transformation sequences are novel and have not been previously reported in the literature. Mega plasmids were found to be present in RKJ4, HL4 and DLC-P11, but their involvement in phenanthrene degradation could not be established. Received: 25 May 1999 / Received revision: 16 July 1999 / Accepted: 1 August 1999  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