首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The activities of cytochrome P450-derived epoxide metabolites of omega-6 polyunsaturated fatty acids (PUFAs) in cellular homeostasis have generated considerable topical interest, but there is less information on the effects of omega-3 PUFA epoxides. Mass spectroscopic data on the epoxides of the omega-3 PUFA eicosapentaenoic acid (EPA) have been reported but the absence of corresponding NMR data currently hinders their biological assessment. In the present study five monoepoxy derivatives of EPA methyl ester were synthesized by treating EPA methyl ester with m-chloroperbenzoic acid. The individual regioisomers were purified by normal-phase chromatography and characterized by LC-MS/MS and a combination of NMR approaches including 1H-, 13C-, 1H-1H-COSY, 1H-13C-HSQC, and 1H-13C-HMBC. The chromatographic properties for these monoepoxides were studied in normal-phase and reversephase-HPLC systems and the MS/MS fragmentation patterns using electrospray ionization were established. This paper also focuses on the NMR characterization of epoxide, olefinic and methylenic moieties and the complete assignments of the isomers.  相似文献   

2.
Epoxide hydrase activity, measured with [3H]styrene oxide as substrate, is present in mammalian liver, kidney, lung, intestine and skin. The hepatic level of the enzyme, measured in vitro with [3H]styrene oxide, benzene oxide or naphthalene-1,2-oxide, is elevated substantially by pretreatment of rats with phenobarbital and to a lesser extent by pretreatment with 3-methylcholanthrene. Metyrapone and 1-(2-isopropylphenyl)-imidazole, two monooxygenase inhibitors, activate epoxide hydrase in vitro, but have no demonstrable effect on the enzyme in vivo. 3,3,3-Trichloropropene oxide, a potent in vitro inhibitor of epoxide hydrase, has no effect on monooxygenase activity measured in vitro with [3H]benzenesulfonanilide. Trichloropropene oxide is extremely toxic. In sub-lethal dosages, it does not significantly inhibit epoxide hydrase activity in vivo, although it and several other epoxides do react with and thereby reduce hepatic levels of glutathione. Cyclohexane oxide, another potent in vitro inhibitor of epoxide hydrase, reduces hepatic glutathione levels to 10% of control values. This relatively non-toxic substance should potentiate the hepatotoxicity of chlorobenzene by inhibiting further metabolism of the toxic chlorobenzene oxide intermediate through either hydration or conjugation with glutathione. Instead, co-administration of cyclohexene oxide and chlorobenzene significantly reduces the rate of metabolism of [14C]chlorobenzene and prevents the hepatic centrilobular necrosis caused by chlorobenzene in rats. Arene oxide-mediated hepatotoxicity apparently is dependent upon a variety of factors including both rates of formation and degradation of arene oxides in tissue. The presently known hydrase inhibitors are not sufficiently selective in their effects on liver cells to permit a quantitative assessment of the relative importance of these factors.  相似文献   

3.
Biocatalytic resolution of 3-(2′-nitrophenoxy)propylene oxide (1a), 3-(3′-nitrophenoxy)propylene oxide (1b) and 3-(4′-nitrophenoxy)propylene oxide (1c) were exploited by using lyophilized cells of yeast Trichosporon loubierii ECU1040 with epoxide hydrolase (EH) activity, which preferentially hydrolyzes (S)-enantiomers of the epoxides (1a–c), yielding (S)-diols and (R)-epoxides. The activity increased as the nitro group in the phenyl ring was shifted from 4′-position (1c) to 2′-position (1a). When the substrate concentration of 1a was increased from 10 to 80 mM, the E-value increased at first, until reaching a peak at 40 mM, and then decreased at higher concentrations (>40 mM). The optically active epoxide (R)-1a was prepared at gram-scale (97% ee, 41% yield). Furthermore, a simple method was developed to predict the enantiomeric excess of substrate (ees) at any time of the whole reaction course based on the ees value determined at a certain reaction time at a relatively lower substrate concentration. This will be helpful for terminating the reaction at a proper time to get both higher optical purity and higher yield of the remaining epoxides.  相似文献   

4.
A two-phase membrane bioreactor was developed to continuously produce enantiopure epoxides using the epoxide hydrolase activity of Rhodotorula glutinis. An aqueous/organic cascade, hydrophilic, hollow-fiber membrane bioreactor was used: (1) to carry out large-scale resolution of epoxides, (2) to continuously extract residual enantiopure epoxides from the aqueous phase, and (3) to separate inhibitory formed diol from the yeast cells contained in the aqueous phase. Dodecane was employed to dissolve-feed epoxide as well as to extract residual epoxide. 1,2-Epoxyhexane was used as a model substrate. By use of this membrane bioreactor, enantiopure (S)-1,2-epoxyhexane (>98% enantiomeric excess) was obtained with a volumetric productivity of 3.8 g l−1 h−1. The continuous-production system was operated for 12 days and resulted in 38 g enantiopure (S)-1,2-epoxyhexane. Received: 14 February 2000 / Received revision: 15 June 2000 / Accepted: 18 June 2000  相似文献   

5.
Docosahexaenoic acid (22:6(n-3)) was recently shown to be metabolized by liver microsomes to vicinal diol regioisomers. To identify the diols, and to compare their biological actions with those of epoxide precursors, we developed a chemical method to synthesize microgram to milligram amounts of epoxides and corresponding diols. In brief, methylated docosahexaenoate was reacted for 15 min with 0.1 eq m-chloroperoxybenzoic acid. After normal- and reverse-phase high performance liquid chromatography, six products were isolated. The underivatized or hydrogenated products were characterized and identified using capillary gas-liquid chromatography and mass spectrometry. The products were identified as 19,20-, 16,17-, 13,14-, 10,11-, 7,8-, and 4,5-epoxy-docosapentaenoate. Per incubation, the total epoxide yield from 22:6(n-3) was 8.6%. By reincubating unused substrate 10-20 times (cycling), the total epoxide could be increased to 55-70%. As found for epoxides of arachidonic and eicosapentaenoic acids, the yield of individual regioisomers increased as the distance between the targeted double bond and carbomethoxy group increased. Each epoxide regioisomer was hydrolyzed to its corresponding vicinal diol. The gas-liquid chromatographic retention times and mass spectra of the diol products were found to match those of metabolites produced by cytochrome P-450 monooxygenases.  相似文献   

6.
A cell-free extract prepared from Fusarium solani pisi grown on cutin, catalyzed the hydration of 18-hydroxy-9,10-epoxyoctadecanoic acid to 9,10,18-trihydroxyoctadecanoic acid while extracts from glucose-grown cells contained <6% of this activity. The product was identified by Chromatographic techniques and by radio gas-liquid chromatography of its periodate oxidation products. This epoxide hydrase activity had a pH optimum at 9.0 and it was located mainly in the 100,000g supernatant fraction. Rate of hydration of the epoxy acid was linear up to 15 min and up to a protein concentration of 30 μg/ml. This fungal epoxide hydrase has a molecular weight of 35,000, as determined by Sephadex G-100 gel filtration. It was partially purified by ammonium sulfate fractionation and gel filtration. The apparent Km and V of the enzyme was 2 × 10?4m and 222 nmoles/min/mg, respectively. Parachloromercuribenzoate strongly inhibited the enzyme, while N-ethylmaleimide was a less potent inhibitor. 1,1,1,-Trichloropropylene-2,3-oxide at 10?3m gave 50% inhibition of the hydration of 18-hydroxy-9,10-epoxyoctadecanoic acid. Kinetic analysis showed that trichloropropylene oxide was a competitive inhibitor. 18-Acetoxy-9,10-epox-yoctadecanoic acid, methyl 18-acetoxy-9,10-epoxyoctadecanoate, 9,10-epoxyoctadecanoic acid, and styrene oxide were not readily hydrated by this fungal epoxide hydrase showing that it has a stringent substrate specificity. Analysis of the enzymatic hydration product on boric acid-impregnated silica gel plates showed that the product obtained from the cis epoxide was exclusively erythro while acid hydrolysis of this epoxide gave rise to the expected threo product. This enzyme is novel in that it catalyzes cis hydration of epoxide while the other epoxide hydrases heretofore isolated catalyzed trans hydration of epoxides.  相似文献   

7.
Using styrene oxide as substrate, most of the epoxide hydrolase (EH) activity monitored in the bulb mite Rhizoglyphus robini was associated with the microsomal compartment. The microsomal and cytosolic EHs did not display any significant preference in hydrating trans stilbene oxide (TSO) and cis stilbene oxide (CSO). The microsomal EH, which has a Km value of 5×10-5M and pH optimum of 7.8, was sensitive to ethanol and its activity was inhibited to a moderate extent by 4-fluorochalcone oxide, TSO, CSO and trans-chalcone oxide at a level of 10-4M. Microsomal EH was considerably induced (4–5-fold) in mites feeding garlic and onion, or ingesting TSO-impregnated filter papers. Other epoxides like CSO, 2,4-dichlorostilbene oxide, methyl chalcone oxide and heptachlor epoxide displayed moderate induction levels (1.4–2.6-fold). Of the toxicants assayed only sodium phenobarbital was a potent inducer. Lindane, malathion and DDT did not stimulate EH activity and 3-methyl-cholanthrene was even inhibitory. A decrease in EH activity was observed with a number of phytochemicals tested such as sinigrin, flavone, menthol, trans--carotene, chalcone, allyl sulphide and trans-cinnamic acid.  相似文献   

8.
The hydration of cis-β-methylstyrene oxide, cis-2,3-octene oxide, and their 18O-enriched forms by epoxide hydrase of rat liver microsomes has been investigated. Both cis epoxides underwent quantitative enzymatic hydration yielding exclusively the corresponding threo diols, indicating that complete stereochemical inversion at a single oxirane carbon had occurred. Mass spectral analysis of diols formed enzymatically from the 18O enriched epoxides indicated they were formed with great regioselectivity, 89% and 85% of the 18O being located at the benzylic carbon of the styrene diol and at C-3 of the octane diol, respectively.  相似文献   

9.
Isolates representing Cryptococcus laurentii and Cryptococcus podzolicus, originating from soil of a heathland indigenous to South Africa, were screened for the presence of enantioselective epoxide hydrolases for 2,2-disubstituted epoxides. Epoxide hydrolase activity for the 2,2-disubstituted epoxide (+/-)-2-methyl-2-pentyl oxirane was found to be abundantly present in all isolates. The stereochemistry of the products formed by the epoxide hydrolase enzymes from isolates belonging to the two species (11 isolates representing C. laurentii and 23 isolates representing C. podzolicus) was investigated. The enantiopreferences of the epoxide hydrolases for 2,2-disubstituted epoxides of these two species were found to be opposite. All strains of C. laurentii preferentially hydrolysed the (S)-epoxides while all C. podzolicus isolates preferentially hydrolysed the (R)-epoxides of (+/-)-2,2-disubstituted epoxides. These findings indicate that the stereochemistry of the products formed from 2,2-disubstituted epoxides by the epoxide hydrolase enzymes of these yeasts should be evaluated as additional taxonomic criterion within the genus Cryptococcus. Also, the selectivity of some epoxide hydrolases originating from isolates of C. podzolicus was high enough to be considered for application in biotransformations for the synthesis of enantiopure epoxides and vicinal diols.  相似文献   

10.
The mutagenicities of 17 closely related oxiranes were determined in 4 tester strains (Salmonella typhimurium TA98, TA100, TA1535, TA1537). The test compounds comprised all possible oxides of benzene and its partially hydrogenated congeners. In TA100 and TA1535, 12 of the tested oxiranes were weak to moderate mutagens. 4 of these were also active in TA98. No mutagenicity was observed with the remaining 5 compounds in any of the 4 strains.The presence of a double bond in formal conjugation with the epoxide ring increased the mutagenicity relative to that of the saturated oxirane. Interestingly, additional epoxide rings within the same molecule did not markedly increase the mutagenic activity, and for the oxiranes that are not activated by a double bond, the relationship between mutagenic activity and the number of epoxide rings in the molecule was even inverse.The influence of bromo and hydroxyl substitution on oxirane mutagenicity is discussed. Most notably, a compound having a 4-hydroxyl group in syn position to a 1,2-epoxide ring fused to the cyclohexane ring, a structure which has been suggested to increase the electrophilic reactivity of dihydrodiol epoxides through hydrogen bonding, was almost inactive.  相似文献   

11.
Epoxide hydrolases (EHs) have been characterized and engineered as biocatalysts that convert epoxides to valuable chiral vicinal diol precursors of drugs and bioactive compounds. Nonetheless, the regioselectivity control of the epoxide ring opening by EHs remains challenging. Alp1U is an α/β-fold EH that exhibits poor regioselectivity in the epoxide hydrolysis of fluostatin C (compound 1) and produces a pair of stereoisomers. Herein, we established the absolute configuration of the two stereoisomeric products and determined the crystal structure of Alp1U. A Trp-186/Trp-187/Tyr-247 oxirane oxygen hole was identified in Alp1U that replaced the canonical Tyr/Tyr pair in α/β-EHs. Mutation of residues in the atypical oxirane oxygen hole of Alp1U improved the regioselectivity for epoxide hydrolysis on 1. The single site Y247F mutation led to highly regioselective (98%) attack at C-3 of 1, whereas the double mutation W187F/Y247F resulted in regioselective (94%) nucleophilic attack at C-2. Furthermore, single-crystal X-ray structures of the two regioselective Alp1U variants in complex with 1 were determined. These findings allowed insights into the reaction details of Alp1U and provided a new approach for engineering regioselective epoxide hydrolases.  相似文献   

12.
The interaction between unsaturated phosphatidylcholines and either exogenous or endogenous (produced by the enzyme system involving myeloperoxidase (MPO), H2 O2 ,and Cl) hypochlorite was studied in multilayer liposomes containing oleic, linoleic, and arachidonic acid residues using MALDI TOF mass spectrometry. At pH 7.4, hypochlorite reacts with the double bond of the oleic acid residue in 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine producing oleic acid chlorohydrin as the main product. Minor amounts of glycols and epoxides were also detected. The main products of the reaction of hypochlorite with 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine were mono and di chlorohydrins of linoleic acid. The signals of monoglycol, epoxide, and glycol or epoxide containing monochlorohydrin derivatives were also present in the mass spectrum. The main products of the reaction of hypochlorite with 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine were lysophosphatidylcholine (1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine) and mono-, di-, and trichlorohydrin. Monoglycol and its derivatives containing one or two chlorohydrin groups were also detected. Along with those, carbonyl compounds (aldehyde and acid) formed as a result of double bond breakage in fifth position of arachidonate were detected. Monochlorohydrin was also found when liposomes comprising 1-stearoyl-2-oleoyl-sn-glycero-3-phosphocholine were incubated in the presence of enzymatic mixture, MPO +H2 O2 +Cl,at pH 6.0. In the absence of the enzyme or either of its substrates (H2 O2 or Cl) or in the presence of the MPO inhibitor (sodium azide) or hypochlorite scavengers (taurine or methionine), monochlorohydrin formation was not observed. These data confirm the suggestion that just the hypochlorite generated in MPO catalysis provides for chlorohydrin formation. Thus, the use of MALDI TOF mass spectrometry has shown, along with chlorohydrins, glycols and epoxides as the products of hypochlorite interaction with unsaturated phosphatidylcholines at physiological pH. It was first determined that hypochlorite breaks double bonds in polyunsaturated phosphatidylcholine and also causes lysophosphatidylcholine formation.  相似文献   

13.
Optically active epoxides can be obtained by kinetic resolution of racemic mixtures using enantioselective epoxide hydrolases. To increase the productivity of the conversion of sparingly aqueous soluble epoxides, we investigated the use of a two-phase aqueous/organic system. A kinetic model which takes into account interphase mass transfer, enzymatic reaction, and enzyme inactivation was developed to describe epoxide conversion in the system by the epoxide hydrolase from Agrobacterium radiobacter. A Lewis cell was used to determine model parameters and results from resolutions carried out in the Lewis cell were compared to model predictions to validate the model. It was found that n-octane is a biocompatible immiscible solvent suitable for use as the organic phase. Good agreement between the model predictions and experimental data was found when the enzyme inactivation rate was fitted. Simulations showed that mass transfer limitations have to be avoided in order to maximize the yield of enantiomerically pure epoxide. Resolution of a 39 g/L solution of racemic styrene oxide in octane was successfully carried out in an emulsion batch reactor to obtain (S)-styrene oxide in high enantiomeric excess (>95% e.e.) with a yield of 30%.  相似文献   

14.
Styrene oxide isomerase (SOI) [EC 5.3.99.7], most probably located in the cell wall, was partially purified from Coiynebacterium sp. AC-5 cells grown in a styrene gas atmospheres. The enzyme catalyzed the isomerization reaction to give phenylacetaldehyde, but did not catalyze its reverse reaction. The optimum pH of the reaction was around 7.0, and the enzyme was unstable below pH 6.0. The Km toward styrene oxide was very low (7.7 × 10?5 m), indicating its high affinity for styrene oxide. The enzyme showed strict substrate specificity, and epoxide compounds other than styrene oxide did not serve as substrates. (S)-Styrene oxide was preferentially converted by the enzyme, compared with the (R)-isomer. The possible application of SOI as a biocatalyst is also discussed.  相似文献   

15.
A variety of chlorinated and fluorinated epoxides and related compounds were synthesized and evaluated as inhibitors of epoxide hydrase. The compounds were tested using chicken liver microsomes and a radiometric assay based on [3H]styrene oxide, and using partially purified chicken liver microsomal epoxide hydrase and a continuous photometric assay based on p-nitrostyrene oxide, whose hydration could be monitored at 310 nm. For the 16 compounds studied both assays gave similar patterns of inhibitory activity. As expected from the relative Km values of the two substrates, all inhibitors were considerably more active against styrene oxide (Km =1.0 mM) than against p-nitrostyrene oxide (Km = 4.2 μM), and styrene oxide was a weak alternate-substrate inhibitor against p-nitrostyrene oxide. 1,1,1-Trichloropropene oxide, however, was a potent alternate-substrate inhibitor against p-nitrostyrene oxide. Addition of various substituents to the α-carbon of styrene oxide generated a series of compounds whose inhibitory potency toward p-nitrostyrene oxide increased in the order H ≈ CF3 < CH3 < CH2Cl < CHCl2 < CCl3 ≈ 1,1,1-trichloropropene oxide. In contrast, addition of a CH3 or CCl3 group to the β-carbon of styrene oxide resulted in only a modest increase in inhibitory potency. 2-Phenyl- and 3-phenyloxetane showed no pronounced inhibitory activity toward either styrene oxide or p-nitrostyrene oxide, but pentafluorophenyl ethylene oxide and 1,1, 1-trichlorobutane-3,4-oxide were moderately active inhibitors, although significantly less potent than 1,1,1-trichloroproene oxide. These results show that electronegativity, steric effects, and hydrophobic effects are each important in governing the interaction of epoxide hydrase substrates with the enzyme, although it is not yet possible to analyze separately the effects of each of these parameters on Km, V, and the catalytic mechanism.  相似文献   

16.
The degradation of epichlorohydrin (3-chloropropylene oxide or 1-chloro-2,3-epoxypropane) by whole-cell suspensions of Xanthobacter strain Py2 was investigated. Cell suspensions prepared from cultures grown with propylene as the carbon source readily degraded epichlorohydrin. The ability to degrade epichlorohydrin correlated with the expression of enzymes involved in alkene and epoxide metabolism, since cell suspensions prepared from cultures grown with glucose or acetone, in which the enzymes of alkene and epoxide oxidation are not expressed, did not degrade epichlorohydrin. The alkene monooxygenase-specific inhibitor propyne had no effect on the degradation of epichlorohydrin, demonstrating that alkene monooxygenase is not involved in epichlorohydrin conversion. The interaction of epichlorohydrin and epibromohydrin with the epoxidase which catalyzes aliphatic epoxide conversions was established by showing that the epihalohydrins were specific and potent inhibitors of propylene oxide-dependent O(inf2) consumption by cell suspensions. The rates of degradation of epoxides in whole-cell suspensions decreased in the series propylene oxide > epifluorohydrin > epichlorohydrin > epibromohydrin. The pathway of epichlorohydrin degradation was investigated and found to proceed with stoichiometric dechlorination of epichlorohydrin. The first detectable product of epichlorohydrin degradation was chloroacetone. Chloroacetone was further degraded by the cell suspensions, and in the process, acetone was formed as a nonstoichiometric product. Acetone was further degraded by the cell suspensions with enzymes apparently induced by the accumulation of acetone. The metabolism of allyl chloride (3-chloropropylene) by propylene-grown cells was initiated by alkene monooxygenase and proceeded through epichlorohydrin, chloroacetone, and acetone as intermediate degradation products. These studies reveal a new pathway for halogenated epoxide degradation which involves halogenated and aliphatic ketones as well as other unidentified intermediates and which is unique from previously characterized hydrolytic degradative pathways.  相似文献   

17.
Goniothalamin oxide (1) is a styryl lactone which was isolated from bark and leaves of several Goniothalamus species. This natural product has some interesting biological properties such as larvicidal and tripanocidal activities. However, no studies on the antiproliferative profile of goniothalamin oxide (1) and its stereoisomers have been reported yet. Here, goniothalamin epoxide (1), isogoniothalamin epoxide (2) and their enantiomers were prepared via epoxidation of (R)-and (S)-goniothalamin (4). A 3:2 molar ratio in favor of goniothalamin oxide (1) and ent-1 was observed from (R)- and (S)-4, respectively, when 3-chloroperbenzoic acid (mCPBA) was employed while an increase to 6:1 molar ratio was achieved with (S,S)-Jacobsen’s catalyst. Antiproliferative activity of these epoxides revealed that ent-isogoniothalamin oxide (ent-2) was the most active against the eight cancer cell lines studied. These results indicate that 6S, 7R and 8R absolute configurations are beneficial for the activity of these epoxides.  相似文献   

18.
Epoxide hydrolases catalyze the conversion of epoxides to diols. The known functions of such enzymes include detoxification of xenobiotics, drug metabolism, synthesis of signaling compounds, and intermediary metabolism. In plants, epoxide hydrolases are thought to participate in general defense systems. In the present study, we report the first structure of a plant epoxide hydrolase, one of the four homologous enzymes found in potato. The structure was solved by molecular replacement and refined to a resolution of 1.95 A. Analysis of the structure allows a better understanding of the observed substrate specificities and activity. Further, comparisons with mammalian and fungal epoxide hydrolase structures reported earlier show the basis of differing substrate specificities in the various epoxide hydrolase subfamilies. Most plant enzymes, like the potato epoxide hydrolase, are expected to be monomers with a preference for substrates with long lipid-like substituents of the epoxide ring. The significance of these results in the context of biological roles and industrial applications is discussed.  相似文献   

19.
MANY chemical carcinogens are mutagenic1 and some non-mutagenic carcinogens are metabolized to mutagenic derivatives2,3. Recent work4–6 has confirmed that epoxides are intermediates in the metabolism of the aromatic double bonds of carcinogenic polycyclic hydrocarbons to hydroxylated derivatives, as Boyland suggested7. In addition to chemical reactions with nucleic acids and histone8, epoxides derived from polycyclic hydrocarbons bind more extensively to the nucleic acids of cells in culture than the parent hydrocarbons9. Hydrocarbon epoxides are also more active in inducing malignant transformation in vitro of hamster embryo and mouse prostate cells10 although, in whole animals, they were less potent carcinogens than the hydrocarbons themselves11–13. As potential mutagens, polycyclic hydrocarbon epoxides are therefore of particular interest, mainly because of the support positive results would give to the somatic mutation theory of carcinogenesis. In the work described here we have tested K-region epoxides of hydrocarbons for their ability to cause host range mutations of T2h+ bacteriophage, specifically because there is no possibility, in this test system, of the epoxides being further metabolized. The epoxides tested were phenanthrene 9,10-oxide (Ph-E), benz(a)anthracene 5,6-oxide (BA-E), dibenz(a,h)anthracene 5,6-oxide (DBA-E), 7-methylbenz(a)anthracene 5,6-oxide (7-MeBA-E), 3-methylcholanthrene 11,12-oxide (MCA-E) and chrysene 5,6-oxide (Ch-E). Ethylene oxide and propylene oxide were used as examples of aliphatic epoxides which do not increase the frequency of host range mutants of T2 bacteriophage and ethyl methanesulphonate (EMS) was used as a known mutagen14.  相似文献   

20.
Abstract

The reaction mechanisms of two isomeric bay-region diol epoxides of 5-methylchrysene (trans-1,2-dihydroxy-anti/-3,4-epoxy-1,2,3,4-tetrahydro-5-methylchrysene (DE-I) and trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydro-5-methylchrysene (DE-II) with double-stranded DNA in aqueous solutions were studied utilizing kinetic flow dichroism and fluorescence techniques. As in the case of the previously studied benzo(a)pyrene-7,8-diol-9,10-oxide isomers (BaPDE), both DE-I and DE-II rapidly form intercalation-type complexes (association constants K = 2700 and 1500 M?1 respectively in a neutral 5mM phosphate solution). The physically bound diol epoxide molecules react on time scales of minutes to form predominantly tetraols; a greater fraction (6±1%) of DE-I than of DE-II (2–3%) molecules react with the DNA to form covalent products. The DE-II isomer is characterized by a greater reactivity than DE-I, and the rates of reaction are markedly accelerated in the presence of DNA in both cases. The linear dichroism spectra of the covalent adducts reveal that the conformations of both types of adducts are similar, with the long axes of the phenanthrenyl chromophores tilted, on the average, at angles of 38-52° with respect to the average orientations of the transition moments (at 260 nm) of the DNA bases. The conformations of the covalently bound DE-I and DE-II molecules resemble those observed in the case of the highly tumorigenic (+) enantiomer of anti-BaPDE. The differences in the biological properties of these and other polycyclic aromatic diol epoxides are discussed in terms of their reactivities with DNA and the conformations of the adducts formed.  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