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.
When [3H]benzo[a]pyrene is incubated in vitro together with DNA, NADPH and rat lung microsomes, covalent binding of benzo[a]pyrene (BP) metabolites to DNA occurs. These metabolite-nucleoside complexes can be resolved into several distinct peaks by elution of a Sephadex LH-20 column with a water-methanol gradient. 3-Methylcholanthrene (MC) pretreatment of animals induces the total covalent binding in vitro several-fold and increases the amounts of at least five metabolite-nucleoside complexes associated with the 7,8-diol-9,10-epoxidcs, the 7,8-oxide or quinones oxygenated further, the 4,5-oxide and phenols oxygenated further. These increases correspond well with the increases in the production of both non-K-region and K-region metabolites of BP by lung microsomes, as determined by highpressure liquid chromatography (HPLC). On the other hand, when [3H]BP is metabolized in isolated perfused rat lung, only the peak representing the 7,8-diol-9,10-epoxide bound to nucleoside(s) is readily detectable and then only in lungs from MC-treated animals. The extent of binding of BP metabolites to lung DNA is very low, about 0.0004% of the total dose applied to the perfusion medium; more than 60% of this can be accounted for by the binding of the 7,8-diol-9,10-epoxides to nucleoside(s). It is suggested that the further metabolism leading to metabolites not available to covalent binding, (e.g. conjugation) of primary BP metabolites in the intact tissue is responsible for the differences in the metabolite-nucleoside patterns observed in vivo, as compared with microsomal metabolism in vitro.  相似文献   

2.
Liver nuclei from 3-methylcholanthrene-treated rats in the presence of NADPH metabolized 3- and 9-hydroxybenzo[a]pyrene and 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene to products that bound to DNA. Maximal binding was obtained with the dihydrodiol which was approximately 3-fold that with 9-hydroxybenzo[a]pyrene, and 60-fold that with 3-hydroxybenzo[a]pyrene, as substrates. Both 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene and 9,10-dihydro-9,10-dihydroxybenzo[a]pyrene were also extensively metabolized by the nuclear fraction but did not give rise to DNA-binding products.The available evidence suggests that the DNA binding species derived from 9-hydroxy-benzo[a]pyrene is 9-hydroxy-benzo[a]pyrene-4,5-oxide and from 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene, as previously observed in different systems, 7,8-dihydro-7,8-dihydroxy-benzo[a]pyrene-9,10-oxide.  相似文献   

3.
(±)-7β,8α-Dihydroxy-9β,10β-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-1) and (±)-7β,8α-dihydroxy-9α,10α-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (diol epoxide-2) are highly mutagenic diol epoxide diastereomers that are formed during metabolism of the carcinogen (±)-trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene. Remarkable stereoselectivity has been observed on metabolism of the optically pure (+)- and (?)-enantiomers of the dihydrodiol which are obtained by separation of the diastereomeric diesters with (?)-α-methoxy-α-trifluoromethylphenylacetic acid. The high stereoselectivity in the formation of diol epoxide-1 relative to diol epoxide-2 was observed with liver microsomes from 3-methylcholanthrene-treated rats and with a purified cytochrome P-448-containing monoxygenase system where the (?)-enantiomer produced a diol epoxide-2 to diol epoxide-1 ratio of 6 : 1 and the (+)-enantiomer produced a ratio of 1 : 22. Microsomes from control and phenobarbital-treated rats were less stereospecific in the metabolism of enantiomers of BP 7,8-dihydrodiol. The ratio of diol epoxide-2 to diol epoxide-1 formed from the (?)- and (+)-enantiomers with microsomes from control rats was 2 : 1 and 1 : 6, respectively. Both enantiomers of BP 7,8-dihydrodiol were also metabolized to a phenolic derivative, tentatively identified as 6,7,8-trihydroxy-7,8-dihydrobenzo[a]pyrene, which accounted for ~30% of the total metabolites formed by microsomes from control and phenobarbital-pretreated rats whereas this metabolite represents ~5% of the total metabolites with microsomes from 3-methylcholanthrene-treated rats. With benzo[a]pyrene as substrate, liver microsomes produced the 4,5-, 7,8- and 9,10-dihydrodiol with high optical purity (>85%), and diol epoxides were also formed. Most of the optical activity in the BP 7,8-dihydrodiol was due to metabolism by the monoxygenase system rather than by epoxide hydrase, since hydration of (±)-benzo[a]pyrene 7,8-oxide by liver microsomes produced dihydrodiol which was only 8% optically pure. Thus, the stereospecificity of both the monoxygenase system and, to a lesser extent, epoxide hydrase plays important roles in the metabolic activation of benzo[a]pyrene to carcinogens and mutagens.  相似文献   

4.
Cultured human fibroblasts from healthy donors were incubated for 30 min with nine different benzo[a]pyrene (BP) derivatives in the presence or absence of liver microsomes from 3-methylcholanthrene treated rats. The induction and repair of DNA strand breaks were analysed by alkaline unwinding and separation of double and single stranded DNA (SS-DNA) by hydroxylapatite chromatography immediately after the incubation or at various times after the treatment. In the absence of microsomes DNA stand breaks were detected in fibroblasts exposed to 30 microM of each of the six BP phenols (1-, 2-, 3-, 7-, 9- or 11-OH-BP) and the three BP dihydrodiols (BP-4,5-, BP-7,8- or BP-9,10-dihydrodiol). After removal of the BP derivatives from the medium the DNA strand breaks disappeared within 24 h. alpha-Naphthoflavone (alpha-NF) caused a decrease in the induction of strand breaks by 1-, 3- and 9-OH-BP but did not affect the induction of strand breaks in cells exposed to BP-7,8-dihydrodiol. In the presence of microsomes DNA strand breaks were found after exposure to 30 microM of each of the six BP phenols (1-, 2-, 3-, 7-, 9- or 11-OH-BP), as well as BP-7,8- and 9,10-dihydrodiol. In contrast BP-4,5-dihydrodiol did not induce strand breaks under these conditions. The induction of strand breaks by BP-7,8-dihydrodiol was enhanced in the presence of cytosine-1-beta-D-arabinofuranoside (AraC). In all cases the DNA strand breaks had disappeared 24 h after removal of the BP derivatives and microsomes except after treatment with BP-7,8-dihydrodiol.  相似文献   

5.
The effect of various co- and anti-carcinogens of colon carcinogenesis on the metabolism of benzo(a)pyrene (BP) in cultured rat colon is reported. Rat colon enzymatically converted BP into metabolites which bind to cellular macromolecules i.e., DNA and protein. Activity of aryl hydrocarbon hydroxylase (AHH) activity and binding levels of BP to macromolecules were higher in the descending colon when compared to other segments. The major metabolites of BP, extractable with ethylacetate, were quinones, tetrols, 7,8-diol and a peak containing 9,10-dihydroxy-9,10-dihydrobenzo(a)pyrene and 7,8,9-trihydroxy-7,8-dihydrobenzo(a)pyrene. The binding levels of BP to DNA and protein in the explant was lowered by co-incubation with 7,8-benzoflavone (7,8-BF) (3.6 and 18.0 μM), a known inhibitor of AHH, and with disulfiram (100 μM), an anti-oxidant. The absence of vitamin A in the media also resulted in a lower level of BP binding to DNA and protein and in lower activity of AHH. Pretreatment with known inducers of AHH such as phenobarbital (PB) or benz(a)anthracene (BA), did not have any significant effect on the binding levels of BP to DNA or on the AHH activity. of the bile acids investigated only taurodeoxycholic acid significantly increased the binding level of BP to DNA.  相似文献   

6.
NADPH-reduction of benzo[a]pyrene 4,5-oxide (BP-4,5-oxide) to BP required four components from rat liver: cytochrome P-450, NADPH cytochrome P-450 reductase, phosphatidylcholine and a soluble, heat-sensitive factor which was present in 105 000 × g supernatant and was also released from microsomes by sonication. The requirement for this factor contrasts with recently reported results from Sugiura et al. (Cancer Res., 40 (1980) 2910). Oxide-reduction was 40 times faster under anaerobic conditions, but oxygen did not affect the stimulation factor. This stimulation was highest (× 15) at low concentrations of microsomal protein (<0.1 mg/ml) and was almost absent at high concentrations of microsomal protein (>1 mg/ml). Oxide-reduction activity was proportional to microsomal protein concentration in the presence of added 105 000 × g supernatant, but for microsomes alone (>0.1 mg/ml) exhibited a parallel plot with an intercept at 0.08 mg/ml microsomal protein. Stimulation was highest at high concentrations of BP-4,5-oxide and a linear plot of V−1 vs. [BP-4,5-oxide]−1 was only obtained in the presence of 105 000 × g supernatant (Km = 3 μM, Vmax = 3.3 nmol/mg/min). Microsomal hydration of BP-4,5-oxide (inhibited in reductase assays) was unaffected by 105 000 × g supernatant, suggesting that stimulation of oxide-reduction did not derive from solubilization of BP-4,5-oxide. Stimulation was observed in the initial rate of reaction and was independent of incubation time. Inhibition of lipid peroxidation, removal of peroxides and deoxygenation were all excluded as explanations of the stimulatory effect.  相似文献   

7.
Liver nuclei from 3-methylcholanthrene-treated rats in the presence of NADPH metabolized 3- and 9-hydroxybenzo[a]pyrene and 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene to products that bound to DNA. Maximal binding was obtained with the dihydrodiol which was approximately 3-fold that with 9-hydroxybenzo[a]pyrene, and 60-fold that with 3-hydroxybenzo[a]pyrene, as substrates. Both 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene and 9,10-dihydro-9,10-dihydroxybenzo[a]pyrene were also extensively metabolized by the nuclear fraction but did not give rise to DNA-binding products.

The available evidence suggests that the DNA binding species derived from 9-hydroxy-benzo[a]pyrene is 9-hydroxy-benzo[a]pyrene-4,5-oxide and from 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene, as previously observed in different systems, 7,8-dihydro-7,8-dihydroxy-benzo[a]pyrene-9,10-oxide.  相似文献   


8.
(±)-7β,8α-Dihydroxy-9β,10β-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BP 7,8-diol-9,10-epoxide) is a suspected metabolite of benzo[a]pyrene that is highly mutagenic and toxic in several strains of Salmonellatyphimurium and in cultured Chinese hamster V79 cells. BP 7,8-diol-9,10-epoxide was approximately 5, 10 and 40 times more mutagenic than benzo[a]pyrene 4,5-oxide (BP 4,5-oxide) in strains TA 98 and TA 100 of S.typhimurium and in V79 cells, respectively. Both compounds were equally mutagenic to strain TA 1538 and non-mutagenic to strain TA 1535 of S.typhimurium. The diol epoxide was toxic to the four bacterial strains at 0.5–2.0 nmole/plate, whereas BP 4,5-oxide was nontoxic at these concentrations. In V79 cells, the diol epoxide was about 60-fold more cytotoxic than BP 4,5-oxide.  相似文献   

9.
Products that appeared to be mainly benzo[a]pyrene 7,8-oxide and benzo[a]pyrene 9,10-oxide were synthesized and their chemical and biochemical properties were investigated. The oxides were unstable and readily rearranged to phenols. They were converted by rat liver homogenates and microsomal preparations into phenols and dihydrodiols, but glutathione conjugates were not formed in appreciable amounts. The dihydrodiols formed from benzo[a]pyrene 7,8- and 9,10-oxide by rat liver microsomal preparations were identical in their chromatographic and spectrographic properties with dihydrodiols formed when benzo[a]pyrene was metabolized by rat liver homogenates. 9,10-Dihydrobenzo[a]pyrene 7,8-oxide and 7,8-dihydrobenzo[a]pyrene 9,10-oxide were also synthesized. They were converted by rat liver homogenates and microsomal preparations into the related cis- and trans-dihydroxy compounds. Glutathione conjugates were formed from the oxides by rat liver homogenates. Both 7,8- and 9,10-dihydrobenzo[a]pyrene were metabolized by rat liver homogenates to mainly the trans-isomers of the related dihydroxy compounds. In experiments with boiled homogenates, the benzo[a]pyrene oxides were converted into phenols, whereas the dihydrobenzo[a]pyrene oxides yielded small amounts of the related dihydroxy compounds.  相似文献   

10.
The involvement of cytochrome P-450 isozymes in the activation of benzo[a]pyrene (BP) by human placental and liver microsomes was studied in vitro using monoclonal antibodies (Mab) toward the major 3-methylcholanthrene (MC)-inducible and phenobarbital-inductible rat liver P-450 isozymes (Mab 1-7-1 and Mab 2-66-3, respectively). Microsomes from human placenta and liver and rat liver were incubated with BP and DNA, and BP-diolepoxide-DNA (BPDE-DNA) adducts were measured by synchronous fluorescence spectrophotometry (SFS). The only BP metabolite giving the same fluorescence peak as chemically modified BPDE-DNA was BP-7,8-dihydrodiol. Five (smokers) out of 29 human placentas (smokers and nonsmokers), and five out of nine human livers were able to metabolically activate BP to BPDE-DNA adducts in this system. The Mab 1-7-1 totally inhibited the formation of BPDE-DNA adducts in placental microsomal incubations. Inhibition using rat or human liver microsomes was 50-60% and about 90%, respectively. The Mab 2-66-3 had no effect in any of the microsome types. Adduct formation was inhibited more strongly and at lower concentrations of Mab 1-7-1 compared with the inhibition of AHH activity. This study is a clear indication of the major role of P-450IA1 (P-450c) in human placenta and probably P-450IA2 (P-450d) in human liver in BP activation, while other isozymes also take part in the activation in rat liver. Furthermore, this clearly indicates that AHH activity and BP activation are not necessarily associated.  相似文献   

11.
The reversible DNA physical binding of a series of non-reactive metabolites and metabolite model compounds derived from benzo[a]pyrene (BP) has been examined in UV absorption and in fluorescence emission and fluorescence lifetime studies. Members of this series have steric and pi electronic properties similar to the highly carcinogenic metabolite trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and the less potent metabolite 4,5-epoxy-4,5-dihydrobenzo(a)pyrene (4,5-BPE). The molecules examined are trans-7,8-dihydroxy-7,8-dihydrobenzo[a]-pyrene (7,8-di(OH)H2BP), 7,8,9,10-tetrahydroxytetrahydrobenzo[a]pyrene (tetrol) 7,8,9,10-tetrahydrobenzo[a]pyrene (7,8,9,10-H4BP), pyrene, trans-4,5-dihydroxy-4,5-dihydrobenzo[a]pyrene (4,5-di(OH)H2BP) and 4,5-dihydrobenzo[a]pyrene (4,5-H2BP). In 15% methanol at 23 degrees C the intercalation binding constants of the molecules studied lie in the range 0.79-6.1 X 10(3) M-1. Of all the molecules examined the proximate carcinogen 7,8-di(OH)-H2BP is the best intercalating agent. The proximate carcinogen has a binding constant which in UV absorption studies is found to be 2.8-6.0 times greater than that of the other hydroxylated metabolites. Intercalation is the major mode of binding for 7,8-di(OH)H2BP and accounts for more than 95% of the total binding. Details concerning the specific role of physical bonding in BP carcinogenesis remain to be elucidated. However, the present studies demonstrate that the reversible binding constants for BP metabolites are of the same magnitude as reversible binding constants which arise from naturally occurring base-base hydrogen bonding and pi stacking interactions in DNA. Furthermore, previous autoradiographic studies indicate that in human skin fibroblasts incubated in BP, pooling of the unmetabolized hydrocarbons occurs at the nucleus. The high affinity of 7,8-di(OH)H2BP for DNA may play a role in similarly elevating in vivo nuclear concentrations of the non-reactive proximate carcinogen.  相似文献   

12.
The metabolism of the polycyclic aromatic hydrocarbon (PAH) carcinogen benzo[a]pyrene (BaP) was studied using microsomes prepared from the skin of the mouse and rat. Topical application of the polychlorinated biphenyl (PCB) Aroclor 1254 or the PAH 3-methylcholanthrene (3-MC) to the skin of the C57BL/6N and DBA/2N mouse and the Sprague-Dawley rat caused statistically significant enhancement of cutaneous microsomal aryl hydrocarbon hydroxylase (AHH) activity in each animal. PCB was a more potent inducer of the enzyme than was 3-MC. BaP metabolism by skin microsomes from the same animals was assessed using high performance liquid chromatography (HPLC). The skin of untreated animals metabolized BaP into 9,10-, 7,8- and 4,5-dihydrodiols, phenols and quinones. Skin application of PCB caused greater than 16–18-fold enhancement of BaP metabolism in the C57BL/6N mouse and the rat and 2–5-fold enhancement in the DBA/2N mouse. Skin application of 3-MC enhanced BaP metabolism 2–8-fold in the C57BL/6N mouse and 5–10-fold in the rat and had no effect in the DBA/2N mouse. The formation of procarcinogenic metabolite BaP-7, 8-diol was greatly enhanced (4–12-fold) by treatment with the PCB and 3-MC in the tumor susceptible C57BL/6N mouse and in the tumor-resistant neonatal Sprague-Dawley rat. In contrast, the formation of BaP-7,8-diol was either slightly enhanced (2-fold) or unaffected by treatment with the PCB or 3-MC in the tumor-resistant DBA/2N mouse. Our data indicate that neither the patterns of metabolism nor the amount of BaP-7,8-diol formation in the skin are reliable predictors of tumor susceptibility to the PAH in rodent skin.  相似文献   

13.
The metabolic activation of benzo[a]pyrene (BP) was examined in six samples of human skin after topical application of the hydrocarbon to the skin in short-term organ culture. The results show that all of the samples were capable of metabolizing BP to water-soluble products and to ether-soluble products that included the 4,5-, 7,8- and 9,10-dihydrodiols and a product which had chromatographic properties identical with those of authentic trans-11,12-dihydro-11,12-dihydroxybenzo[a]pyrene (BP-11,12-diol). The major BP-deoxyribonucleoside adduct detected in each skin sample appeared to be formed from the reaction of r-7,t-8-dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (anti-BP-7,8-diol 9,10-oxide) with deoxyguanosine residues in DNA.  相似文献   

14.
A dual-label HPLC assay to measure femtomole quantities of ethyl acetate-extractable [3H]benzo[a]pyrene metabolites was developed. 14C-labeled metabolites of benzo[a]pyrene formed by rat liver 9000g supernatant were used as both internal standards and chromatographic markers. The percentage deviation between assays was determined to be between 11 and 13% for 9,10-dihydro-9,10-dihydroxybenzo[a]pyrene, 7,8-dihydro-7,8-dihydroxybenzo[a]pyrene, benzo[a]pyrene-3,6-quinone, benzo[a]pyrene-1,6-quinone, and 9-hydroxybenzo[a]pyrene, 22% for 4,5-dihydro-4,5-dihydroxybenzo[a]pyrene, and less than 5% for 3-hydroxybenzo[a]pyrene. The detection limit of this assay was between 3 and 10 fmol per metabolite. The application of this technique to the metabolism of [3H]benzo[a]pyrene by microsomes of hamster and human oral cavity tissue is described.  相似文献   

15.
Fluoro substitution of benzo[a]pyrene (BP) has been very useful in determining the mechanism of cytochrome P450-catalyzed oxygen transfer in the formation of 6-hydroxyBP (6-OHBP) and its resulting BP 1,6-, 3,6-, and 6,12-diones. We report here the metabolism of 1-FBP and 3-FBP, and PM3 calculations of charge densities and bond orders in the neutral molecules and radical cations of BP, 1-FBP, 3-FBP, and 6-FBP, to determine the mechanism of oxygen transfer for the formation of BP metabolites. 1-FBP and 3-FBP were metabolized by rat liver microsomes. The products were analyzed by HPLC and identified by NMR. Formation of BP 1,6-dione and BP 3,6-dione from 1-FBP and 3-FBP, respectively, can only occur by removal of the fluoro ion from C-1 and C-3, respectively, via one-electron oxidation of the substrate. The combined metabolic and theoretical studies reveal the mechanism of oxygen transfer in the P450-catalyzed formation of BP metabolites. Initial abstraction of a pi electron from BP by the [Fe(4+)=O](+)(*) of cytochrome P450 affords BP(+)(*). This is followed by oxygen transfer to the most electropositive carbon atoms, C-6, C-1, and C-3, with formation of 6-OHBP (and its quinones), 1-OHBP, and 3-OHBP, respectively, or the most electropositive 4,5-, 7,8-, and 9,10- double bonds, with formation of BP 4,5-, 7,8-, or 9,10-oxide.  相似文献   

16.
The metabolism of benzo(a)pyrene (BP) by rat liver microsomes has been examined in the presence of competitive (styrene oxide), uncompetitive (3,3,3-trichloropropene oxide, TCPO), and noncompetitive (cyclohexene oxide) inhibitors of arene oxide (AO) hydrase. Formation of BP-dihydrodiols was inhibited selectively, with 9,10-dihydrodiol at the lowest inhibitor concentration, and then 7,8- and 4,5-dihydrodiols were decreased at higher inhibitor concentrations. Increased levels of 9-phenol, 7-phenol, and 4,5-oxide appeared selectively in the same order. Appearance of these alternate products did not quantitatively compensate for the loss of dihydrodiols so that there was a net loss of oxidation products. A 1000-fold increase in the concentration of TCPO did not further inhibit BP oxidation. Formation of quinones and 3-phenol was completely unaffected by the inhibitors. The limiting decrease in BP oxidation products was the same for each inhibitor and was greater for 3-methylcholanthrene-induced microsomes (25–30%) than for phenobarbital-induced microsomes (15–20%), which produced a smaller proportion of dihydrodiols. Several mechanisms for this specific loss of oxide-derived reaction products have been considered. BP-oxidation products, particularly 9-phenol, significantly inhibit BP oxidation; however, this inhibition is nonspecific in that 3-phenol, quinones, and oxide-derived products are all decreased. 9-Phenol was far more effective as an inhibitor than as a substrate. Glutathione conjugation of oxides due to cytosolic contamination was excluded by virtue of the near absence of water-soluble products. Reduction of 4,5-oxide occurred, in the absence of oxygen, at a rate which was about half the rate of BP monooxygenation, but this rate decreased 75-fold in the presence of air. Enhanced reduction of BP-oxides in the presence of hydrase inhibitors can explain the action of these inhibitors on BP oxidation if the reduction of microsomally generated 4,5-oxide is several times faster than reduction of added 4,5-oxide. The selective effect of hydrase inhibitors on different dihydrodiols can be attributed to differences in the relative stabilities of the intermediate oxides. The formation of 4,5-dihydrodiol from BP is relatively insensitive to hydrase inhibitors in comparison to the hydration of added 4,5-oxide; this results from the rate-determining monooxygenation step.  相似文献   

17.
The 7-methylbenzo[a]pyrene (7-MBaP) was incubated with liver microsomes of rats pretreated with polychlorinated biphenyls (Aroclor 1254) (PCBs). Metabolites of 7-MBaP were isolated by both reversed-phase and normal-phase high performance liquid chromatography (HPLC) and were characterized by nuclear magnetic resonance, UV-visible and mass spectral analyses. The predominant metabolite of 7-MBaP was found to be 3-hydroxy-7-methylbenzo[a]pyrene (3-hydroxy-7-MBaP). Other identified metabolites include 7-MBaP 4,5-, 7,8-, and 9,10-trans-dihydrodiols, 7-hydroxymethyl-BaP, 7-hydroxymethyl-BaP trans-9,10-dihydrodiol, 9-hydroxy-7-MBaP, 3-hydroxy-7-hydroxymethyl-BaP, 7-MBaP 1,6- and 3,6- quinones, and a hydroquinone which is also formed by further metabolism of the 3-hydroxy-7-MBaP. Comparative metabolic studies of 7-MBaP and BaP indicated that, relative to that of BaP, the methyl substituent of 7-MBaP slightly increases the formation of 3-hydroxy-7-MBaP and decreases the metabolism at other regions of the 7-MBaP molecule. The finding that a 7,8-dihydrodiol is a metabolite indicates that, like BaP, 7-MBaP may also be activated to the potentially reactive 7,8-dihydrodiol 9,10-epoxides although their formations are significantly reduced.  相似文献   

18.
The metabolism of benzo[a]pyrene (BP) by microsomal fractions of the skin, lungs and liver of the mouse, and the effects on this process of pretreatment with the xenobiotics phenobarbital (PB) and 3-methylcholanthrene (3-MC) were examined. Differences between the untreated tissues were found both in terms of the total amounts of diol recovered and in the relative proportions of the individual diols extracted following incubation. Induction with PB or 3-MC significantly altered the profiles of metabolic diols obtained with epidermal and hepatic microsomes compared with their respective controls. Pulmonary microsomes showed similar trends to those obtained with liver microsomes but these were not statistically significant. The optical purity of the BP-7,8-diol that was formed by each microsomal type was examined by direct resolution of the enantiomers on HPLC using a chiral stationary phase. In each case the (-)-7R,8R-enantiomer predominated. Pretreatment with 3-MC significantly decreased the optical purity of BP-7,8-diol recovered from incubations with skin microsomes, but significantly increased the optical purity of the diol extracted from incubations with lung and liver microsomes. In addition to the diols, an unidentified BP metabolite was found that eluted between BP-9,10- and 4,5-diol on a reverse-phase high-performance liquid chromatography (HPLC) system and which represented a major product in extracts of incubations of BP with both induced and uninduced skin and lung microsomal fractions.  相似文献   

19.
Administration of 3-methylcholanthrene (3MC) to rats greatly enhanced the aryl hydrocarbon hydroxylase (AHH) activity of liver nuclei. However, the binding in vitro [3H]benzo[alpha]pyrene (BP) to DNA within the nuclei which occurred at the same time as hydroxylation of BP was much less enhanced. Thin layer chromatography of the metabolites of BP produced by these nuclei revealed the same metabolites in similar relative amounts as were produced by rat liver microsomes prepared from rats which had received 3MC. The binding to DNA was further analysed by hydrolysis of the DNA and fractionation on a Sephadex column. This analysis revealed that the binding to DAN in nuclei was very similar in nature to that which occurred when calf-thymus DNA was added to microsomes metabolising BP.  相似文献   

20.
Conversion of benzo[a]pyrene (BP) to BP 7,8-dihydrodiol 9,10-oxides (DE) (measured as 7,10/8,9-tetrols) by untreated (UT) rat liver microsomes is over 10 times slower than following 3-methylcholanthrene (MC) induction. Time courses have been subjected to a kinetic analysis analogous to that previously reported for metabolism by MC-induced microsomes (J. Biol. Chem., 259 (1984) 13770–13776). Competition between BP and 7,8-dihydrodiol for P-450 is the major determinant of the rate of DE formation. Glucuronidation of quinones and phenols only increases the isolated BP metabolites including DE by 40%. This indicates far less inhibition by these products than for metabolism in MC-microsomes (4–6-fold). Thus stimulation may result from a decreased quinone-mediated oxidation of metabolites. In the presence of DNA, UT-microsomes metabolize BP to approximately equal amounts of 9-phenol-4,5-oxide (9-PO) and DE/DNA adducts. Addition of uridine diphosphoglucuronic acid (UDPGA) fails to enhance modification of DNA by DE, but formation of the 9-PO adduct is reduced as a result of lower free 9-phenol levels. The kinetic characteristics of BP metabolism by UT-microsomes are highly sensitive to the presence of very small but variable amounts (2–25 pmol/mg) of the very active cytochrome P-450c, which is the predominant form in MC-microsomes. The major effect of elevated levels of P-450c is an 8-fold increase in DE formation at low concentrations of BP due to a lowering of Km (7.9–2.6 μM) and an increase in the regioselectivity for DE formation from 7,8-dihydrodiol (5–15% of total BP metabolites). The formation of DE was directly correlated with the content of P-450c (r = 0.94). The presence of increased levels of P-450c in UT-microsomes is probably due to previous exposure of the animals to environmental inducers and is minimized by controlled housing and feeding.  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