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.
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   

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.
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   

4.
Metabolism of anthracene by a Rhodococcus species   总被引:2,自引:0,他引:2  
A Rhodococcus sp. isolated from contaminated river sediment was investigated to determine if the isolate could degrade high molecular mass polycyclic aromatic hydrocarbons. The Rhodococcus sp. was able to utilize anthracene (53%), phenanthrene (31%), pyrene (13%), and fluoranthene (5%) as sole source of carbon and energy, but not naphthalene or chrysene. In a study of the degradation of anthracene by a Rhodococcus sp., the identification of ring-fission products indicated at least two ring-cleavage pathways. One results in the production of 6,7-benzocoumarin, previously shown to be produced chemically from the product of meta cleavage of 1,2-dihydroxyanthracene, a pathway which has been well established in Gram-negative bacteria. The second is an ortho cleavage of 1,2-dihydroxyanthracene that produces 3-(2-carboxyvinyl)naphthalene-2-carboxylic acid, a dicarboxylic acid ring-fission product. This represents a novel metabolic pathway only identified in Gram-positive bacteria.  相似文献   

5.
Cells of the gram-negative bacterium Ralstonia sp. strain SBUG 290 grown in the presence of biphenyl are able to cooxidize dibenzofuran which has been 1,2-hydroxylated. Meta cleavage of the 1,2-dihydroxydibenzofuran between carbon atoms 1 and 9b produced 2-hydroxy-4-(3′-oxo-3′H-benzofuran-2′-yliden)but-2-enoic acid, which was degraded completely via salicylic acid. The presence of these intermediates indicates a degradation mechanism for dibenzofuran via lateral dioxygenation by Ralstonia sp. strain SBUG 290.  相似文献   

6.
A bacterial strain was isolated from a Mumbai coastal area. It was dosed with anthracene and phenanthrene, and, after 14 days of incubation, it had degraded 90% and 93% of the anthracene and phenanthrene, respectively. The metabolites were extracted and identified by ultraviolet (UV)-visible light absorption, high-performance liquid chromatography, mass spectrometry, and by comparing with actual compounds and data. Neutral extracts from anthracene showed four metabolites, viz 1,2-dihydroxyanthracene, 6,7-benzocoumarin, 1-methoxy-2-hydroxyanthracene, and 9,10 anthraquinone. When Pseudomonas were grown in the presence of phenanthrene, two metabolites, viz 9,10-dihydroxyphenanthrene and 3,4-dihydroxyphenanthrene were identified.  相似文献   

7.
1. Partially purified extracts of a Pseudomonas converted the meta ring-fission product of 4-methylcatechol into a compound having spectroscopic and chemical properties consistent with its being 2-oxohex-4-enoic acid. 2. Catechol and 3-methylcatechol were both metabolized to a compound that appeared to be 2-oxopent-4-enoic acid. 3. Solutions of norvaline and norleucine were prepared from these metabolites. 4. A reaction scheme is presented for the conversion of catechols into hydroxyoxo acids after meta ring-fission.  相似文献   

8.
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   

9.
Two new lignans, identified as 6,7-dihydroxy-4-(4′’-hydroxy-3′’-methoxyphenyl)-2-naphthoic acid, named Oleralignan C (1), and 4-(3,4-dihydroxyphenyl)-6-hydroxy-7-methoxy-2-naphthoic acid, named Oleralignan D (2), were obtained from Portulaca oleracea L. The structures were determined by spectroscopic methods, including UHPLC-ESI-QTOFMS, 1D and 2D NMR. Both Oleralignan C (1) and Oleralignan D (2) inhibited the inflammatory factors, IL-1β and TNF-α in RAW 264.7 cells induced by lipopolysaccharide (LPS). Both compounds also could clear 1,1-diphenyl-2-picryl-hydrazyl (DPPH) radicals indicating their antioxidant potential.  相似文献   

10.
Lu J  Dang Z  Lu G  Yang C  Yi X  Guo C 《Current microbiology》2012,65(3):225-230
The fusant strain (F14), which produced by protoplast fusion between Sphingomonas sp. GY2B (GenBank DQ139343) and Pseudomonas sp. GP3A (GenBank EU233280), was tested for phenanthrene biodegradation at 30 °C and pH of 7.0. The kinetics of phenanthrene biodegradation by F14 was investigated over a wide range of initial concentration (15-1,000 mg l(-1)). The rate and the extent of phenanthrene degradation increased with the increase of concentration up to 230 mg l(-1), which indicated negligible inhibition effect at low concentrations. The non-competitive inhibition model was found to be fit for the process. GC-MS analysis showed that biodegradation of phenanthrene by F14 was via dioxygenation at both 1,2- and 3,4-positions and followed by 2-hydroxy-1-naphthoic acid and 1-hydroxy-2-naphthoic acid. The relative intensity of 2-hydroxy-1-naphthoic acid was approximately 3-4 times higher than that of 1-hydroxy-2-naphthoic acid, indicating the 2-hydroxy-1-naphthoic acid was the predominant product in the phenanthrene degradation by fusant strain F14.  相似文献   

11.
The biodegradation of the polycyclic aromatic hydrocarbon phenantherene by the rhizobacterial strain Ensifer meliloti P221, isolated from the root zone of plant grown in PAH-contaminated soil was studied. Bacterial growth and phenanthrene degradation under the influence of root-exuded organic acids were also investigated. Analysis of the metabolites produced by the strain by using thin-layer chromatography, gas chromatography, high-pressure liquid chromatography, and mass-spectrometry revealed that phenanthrene is bioconverted via two parallel pathways. The first, major pathway is through terminal aromatic ring cleavage (presumably at the C3–C4 bond) producing benzocoumarin and 1-hydroxy-2-naphthoic acid, whose further degradation with the formation of salicylic acid is difficult or is very slow. The second pathway is through the oxidation of the central aromatic ring at the C9–C10 bond, producing 9,10-dihydro-9,10-dihydroxyphenanthrene, 9,10-phenanthrenequinone, and 2,2′-diphenic acid. This is the first time that the dioxygenation of phenanthrene at the C9 and C10 atoms, proven by identification of characteristic metabolites, has been reported for a bacterium of the Ensifer genus.  相似文献   

12.
《Phytochemistry》1999,52(7):1307-1312
A phenylpropanoid, threo-3-chloro-1-(4-hydroxy-3-methoxyphenyl)propane-1,2-diol, was isolated from the berries of Pimenta dioica together with five known compounds, eugenol, 4-hydroxy-3-methoxycinnamaldehyde, 3,4-dimethoxycinnamaldehyde, vanillin and 3-(4-hydroxy-3-methoxyphenyl)propane-1,2-diol. In addition, the stereochemistry of 3-(4-hydroxy-3-methoxyphenyl)propane-1,2-diol was determined. The phenylpropanoids inhibited autoxidation of linoleic acid in a water-alcohol system.  相似文献   

13.
Reactions of peracetylated α-D-glucosaminyl chloride with isomeric hydroxybenzoic and 1-hydroxy-2-naphthoic acids in a solid phase transfer system of potassium carbonate-acetonitrile were studied. It was found that the nature of carboxylic acids, lipophilicity of the phase transfer catalyst, and reaction temperatures affected the reaction composition and product yields. The O-β-glycosyl esters of ortho-hydroxyaromatic acids were first found to form anomeric 1,2-cis derivatives in the presence of potassium carbonate. The structures of the synthesized compounds were confirmed by 1H NMR spectroscopy. As was shown in vivo experiments, the analgesic activities of glycosyl esters of salicylic acid and peracetylated 2-carboxyphenylglucosaminide were comparable with that of aspirin.  相似文献   

14.
Anaerobic degradation of naphthalene, 2-methylnaphthalene, and tetralin (1,2,3,4-tetrahydronaphthalene) was investigated with a sulfate-reducing enrichment culture obtained from a contaminated aquifer. Degradation studies with tetralin revealed 5,6,7,8-tetrahydro-2-naphthoic acid as a major metabolite indicating activation by addition of a C1 unit to tetralin, comparable to the formation of 2-naphthoic acid in anaerobic naphthalene degradation. The activation reaction was specific for the aromatic ring of tetralin; 1,2,3,4-tetrahydro-2-naphthoic acid was not detected. The reduced 2-naphthoic acid derivatives tetrahydro-, octahydro-, and decahydro-2-naphthoic acid were identified consistently in supernatants of cultures grown with either naphthalene, 2-methylnaphthalene, or tetralin. In addition, two common ring cleavage products were identified. Gas chromatography-mass spectrometry (GC-MS) and high-resolution GC-MS analyses revealed a compound with a cyclohexane ring and two carboxylic acid side chains as one of the first ring cleavage products. The elemental composition was C11H16O4 (C11H16O4-diacid), indicating that all carbon atoms of the precursor 2-naphthoic acid structure were preserved in this ring cleavage product. According to the mass spectrum, the side chains could be either an acetic acid and a propenic acid, or a carboxy group and a butenic acid side chain. A further ring cleavage product was identified as 2-carboxycyclohexylacetic acid and was assumed to be formed by β-oxidation of one of the side chains of the C11H16O4-diacid. Stable isotope-labeling growth experiments with either 13C-labeled naphthalene, per-deuterated naphthalene-d8, or a 13C-bicarbonate-buffered medium showed that the ring cleavage products derived from the introduced carbon source naphthalene. The series of identified metabolites suggests that anaerobic degradation of naphthalenes proceeds via reduction of the aromatic ring system of 2-naphthoic acid to initiate ring cleavage in analogy to the benzoyl-coenzyme A pathway for monoaromatic hydrocarbons. Our findings provide strong indications that further degradation goes through saturated compounds with a cyclohexane ring structure and not through monoaromatic compounds. A metabolic pathway for anaerobic degradation of bicyclic aromatic hydrocarbons with 2-naphthoic acid as the central intermediate is proposed.  相似文献   

15.
Cells of the gram-negative bacterium Ralstonia sp. strain SBUG 290 grown in the presence of biphenyl are able to cooxidize dibenzofuran which has been 1,2-hydroxylated. Meta cleavage of the 1, 2-dihydroxydibenzofuran between carbon atoms 1 and 9b produced 2-hydroxy-4-(3'-oxo-3'H-benzofuran-2'-yliden)but-2-enoic acid, which was degraded completely via salicylic acid. The presence of these intermediates indicates a degradation mechanism for dibenzofuran via lateral dioxygenation by Ralstonia sp. strain SBUG 290.  相似文献   

16.
Cultures of Mycobacterium vanbaalenii strain PYR-1 grown in mineral salts medium and nutrients in the presence of benz[a]anthracene metabolized 15% of the added benz[a]anthracene after 12days of incubation. Neutral and acidic ethyl acetate extractable metabolites were isolated and characterized by high performance liquid chromatography (HPLC) and uv–visible absorption, gas chromatography/mass (GC/MS) and nuclear magnetic resonance (NMR) spectral analysis. Trimethylsilylation of the metabolitesfollowed by GC/MS analysis facilitated identification of metabolites. The characterization of metabolites indicated that M. vanbaalenii initiated attack of benz[a]anthracene at the C-1,2-, C-5,6-, C-7,12- and C-10,11-positions to form dihydroxylated and methoxylated intermediates. The major site of enzymatic attack was in the C-10, C-11 positions. Subsequent ortho- and meta-cleavage of each of the aromatic rings led to the accumulation of novel ring-fission metabolites in the medium. The major metabolites identified were 3-hydrobenzo[f]isobenzofuran-1-one (3.2%), 6-hydrofuran[3,4-g]chromene-2,8-dione (1.3%), benzo[g]chromene-2-one (1.7%), naphtho[2,1-g]chromen-10-one (48.1%), 10-hydroxy-11-methoxybenz[a]anthracene (9.3%), and 10,11-dimethoxybenz[a]anthracene (36.4%). Enzymatic attack at the C-7 and C-12 positions resulted in the formation of benz[a]anthracene-7,12-dione, 1-(2-hydroxybenzoyl)-2-naphthoic acid, and 1-benzoyl-2-naphthoic acid. A phenyl-naphthyl metabolite, 3-(2-carboxylphenyl)-2-naphthoic acid, was formed when M. vanbaalenii was incubated with benz[a]anthracene cis-5,6-dihydrodiol, indicating ortho-cleavage of 5,6-dihydroxybenz[a]anthracene. A minor amount of 5,6-dimethoxybenz[a]anthracene was also formed. The data extend and propose novel pathways for the bacterial metabolism of benz[a]anthracene.  相似文献   

17.
Two diastereoisomers, 5R,6R-5-hydroxy-6(9α)-oxido-11α,15S-dihydroxyprost-13-enoic acid (7) and 5S,6S-5-hydroxy-6(9α)-oxido-11α,15S-dihydroxyprost-13-enoic acid (10) were synthesized for evaluation as possible biosynthetic intermediates in the enzymatic transformation of PGH2 or PGG2 into PGI2. The synthetic sequence entails the stereospecific reduction of the 9-keto function in PGE2 methyl ester after protecting the C-11 and C-15 hydroxyls as tbutyldimethylsilyl ethers. The resulting PGF derivative was epoxidized exclusively at the C-5 (6) double bond to yield a mixture of epoxides, which underwent facile rearrangement with SiO2 to yield the 5S,6S and 5R,6R-5-hydroxy-6(9α)-oxido cyclic ethers. It was found that dog aortic microsomes were unable to transform radioactive 9β-5S,6S[3H] or 9β-5R,6R[3H]-5-hydroxy-6(9α)-oxido cyclic ethers into PGI2. Also, when either diastereoisomer was included in the incubation mixture, neither isomer diluted the conversion of [1-14C]arachidonic acid into [1-14C]PGI2.  相似文献   

18.
A new lignan 1-hydroxy-2-(4-hydroxy-3-methoxyphenyl)-6-(3-acetyl-4-hydroxy-5-methoxyphenyl)-3,7-dioxabicyclo[3.3.0]octane, the secoiridoid 2H-pyran-4-acetic acid,3-hydroxymethyl-2,3-dihydro-5-(methoxycarbonyl)-2-methyl-, methyl ester, the phenylglycoside 4-[beta-D-xylopyranosyl-(1-->6)]-beta-D-glucopyranosyl-1,4-dihydroxy-2-methoxybenzene and the lactone 3-[1-(hydroxymethyl)-1-propenyl] delta-glutarolactone were isolated and identified on the basis of spectroscopic data including two-dimensional NMR, as components of olive oil mill waste-waters. The known aromatic compounds catechol, 4-hydroxybenzoic acid, protocatechuic acid, vanillic acid, 4-hydroxy-3,5-dimethoxybenzoic acid, 4-hydroxyphenylacetic acid, 3,4-dihydroxyphenylacetic acid, tyrosol, hydroxytyrosol, 2-(4-hydroxy-3-methoxy)phenylethanol, 2-(3,4-dihydroxy)phenyl-1,2-ethandiol, p-coumaric acid, caffeic acid, ferulic acid, sinapic acid, 1-O-[2-(3,4-dihydroxy)phenylethyl]-(3,4-dihydroxy)phenyl-1,2-ethandiol, 1-O-[2-(4-hydroxy)phenylethyl]-(3,4-dihydroxy)phenyl-1,2-ethandiol, D(+)-erythro-1-(4-hydroxy-3-methoxy)-phenyl-1,2,3-propantriol, p-hydroxyphenethyl-beta-D-glucopyranoside,2(3,4-dihydroxyphenyl)ethanol 3beta-D-glucopyranoside, and 2(3,4-dihydroxyphenyl)ethanol 4beta-D-glucopyranoside were also confirmed as constituents of the waste-waters.  相似文献   

19.
From the rhizomes of Smilax corbularia Kunth. (Smilacaceae), 11 compounds, (2R,3R)-2″-acetyl astilbin, (2R,3R)-3″-acetyl astilbin, (2R,3R)-4″-acetyl astilbin, (2R,3R)-3″-acetyl engeletin, (2R,3S)-4″-acetyl isoastilbin, 2-(4-hydroxyphenyl)-3,4,9,10-tetrahydro-3,5-dihydroxy-10-(3,4-dihydroxyphenyl)-(2R,3R,10R)-2H,8H-benzo [1,2-b:3,4-b′] dipyran-8-one, 2-(4-hydroxyphenyl)-3,4,9,10-tetrahydro-3,5-dihydroxy-10-(3,4-dihydroxyphenyl)-(2R,3R,10S)-2H, 8H-benzo [1,2-b:3,4-b′] dipyran-8-one, 3,4-dihydro-7-hydroxy-4-(3,4-dihydroxyphenyl)-5-[(1E)-2-(4-hydroxyphenyl) ethenyl]-2H-1-benzopyran-2-one, 3,4-dihydro-7-hydroxy-4-(3,4-dihydroxy-phenyl)-5-[(1E)-2-(3,4-dihydroxyphenyl) ethenyl]-2H-1-benzopyran-2-one, 3,4-dihydro-7-hydroxy-4-(4-hydroxy-3-methoxyphenyl)-5-[(1E)-2-(4-hydroxyphenyl) ethenyl]-2H-1-benzopyran-2-one, and 5,7,3′,4′-tetrahydroxy-3-phenylcoumarin along with 34 known compounds were isolated and characterized as 19 flavonoids, 14 catechin derivatives, 6 stilbene derivatives, and 6 miscellaneous substances. All isolates had their estrogenic and anti-estrogenic activities determined using the estrogen-responsive human breast cancer cell lines MCF-7 and T47D. The major constituents were recognized as flavanonol rhamnosides by the suppressive effect on estradiol induced cell proliferation at a concentration of 1 μM. Meanwhile, flavanonol rhamnoside acetates demonstrated estrogenic activity in both MCF-7 and T47D cells at a concentration of 100 μM, and they enhanced the effects of co-treated E2 on T47D cell proliferation at concentrations of more than 0.1 μM.  相似文献   

20.
The double bonds of 2-hydroxy-6-oxo-6-phenylhexa-2,4-dienoic acid (HOPDA) were stabilized by methylation to establish which of the double bonds of the meta ring-fission compound of biphenyl was reduced by the HOPDA reducing enzyme. HOPDA reducing enzyme III converted 2-methoxy-6-oxo-6-phenylhexa-2,4-dienoic acid methyl ester into 2-methoxy-6-oxo-6-phenylhexa-2-enoic acid methyl ester. To discover the metabolic pathway of HOPDA, partially purified enzyme fractions were used. The eluate from a 2nd column of DEAE-cellulose transformed HOPDA to γ-benzoylbutyric acid, 2,6-dioxo-6-phenylhexanoic acid, and γ-benzoylbutyraldehyde. Fractions passed through the 1st column of DEAE-cellulose formed γ-benzoylbutyric acid and 2-hydroxy-6-oxo-6-phenylhexanoic acid from HOPDA. Based on these data and previous reports, a new metabolic divergence of biphenyl and related compounds was proposed.  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