首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Gu Y  Singh SV  Ji X 《Biochemistry》2000,39(41):12552-12557
Murine class alpha glutathione S-transferase A1-1 (mGSTA1-1), unlike mammalian class alpha GSTs, is the most efficient in the glutathione (GSH) conjugation of the ultimate carcinogenic metabolite of benzo[a]pyrene, (+)-anti-7,8-dihydroxy-9,10-oxy-7,8,9, 10-tetrahydrobenzo[a]pyrene [(+)-anti-BPDE] [Hu, X., Srivastava, S. K., Xia, H., Awasthi, Y. C., and Singh, S. V. (1996) J. Biol. Chem. 271, 32684-32688]. Here, we report the crystal structures of mGSTA1-1 in complex with GSH and with the GSH conjugate of (+)-anti-BPDE (GSBpd) at 1.9 and 2.0 A resolution, respectively. Both crystals belong to monoclinic space group C2 with one dimer in the asymmetric unit. The structures reveal that, within one subunit, the GSH moiety interacts with residues Y8, R14, K44, Q53, V54, Q66, and T67, whereas the hydrophobic moiety of GSBpd interacts with the side chains of F9, R14, M207, A215, R216, F219, and I221. In addition, the GSH moiety interacts with D100 and R130 from the other subunit across the dimer interface. The structural comparison between mGSTA1-1.GSH and mGSTA1-1.GSBpd reveals significant conformational differences. The movement of helix alpha9 brings the residues on the helix into direct interaction with the product. Most noticeable are the positional displacement and conformational change of R216, one of the residues located in helix alpha9. The side chain of R216, which points away from the H-site in the mGSTA1-1.GSH complex, probes into the active site and becomes parallel with the aromatic ring system of GSBpd. Moreover, the guanidinium group of R216 shifts approximately 8 A and forms a strong hydrogen bond with the C8 hydroxyl group of GSBpd, suggesting that the electrostatic assistance provided by the guanidinium group facilitates the ring-opening reaction of (+)-anti-BPDE. The structure of mGSTA1-1. GSBpd is also compared with those of hGSTP1-1[V104,A113].GSBpd, hGSPA1-1.S-benzylglutathione, and mGSTA4-4. 4-S-glutathionyl-5-pentyltetrahydrofuran-2-ol. The comparison provides further evidence that supports the functional roles of R216 and helix alpha9. The lack of mobility of helix alpha9 and/or the lack of electrostatic assistance from R216 may be responsible for the relatively lower activity of hGSTA1-1, mGSTA4-4, and hGSTP1-1 toward (+)-anti-BPDE.  相似文献   

2.
H Xia  Y Gu  S S Pan  X Ji  S V Singh 《Biochemistry》1999,38(31):9824-9830
We have previously identified a novel Alpha class murine glutathione (GSH) S-transferase isoenzyme (designated mGSTAl-2) which is exceptionally efficient in catalyzing the GSH conjugation of (+)-anti-7,8-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene [(+)-anti-BPDE], the ultimate carcinogen of widespread environmental pollutant benzo[a]pyrene. Furthermore, we have demonstrated that the Al-type subunit of this isoenzyme is significantly more active toward (+)-anti-BPDE than the other subunit (mGSTA2). To establish the basis for catalytic differences between mGSTAl and mGSTA2, which differ in their primary structures by 10 amino acids [distributed in three sections (I-III) as clusters of two (residues 65 and 95), three (residues 157, 162, and 169), and five (residues 207, 213, 218, 221, and 222) amino acids], three chimeric enzymes were expressed and tested for their activity toward (+)-anti-BPDE. These studies revealed that amino acid substitution(s) in section III determined the high catalytic activity of mGSTAl. Molecular modeling studies suggested that amino acid substitutions at positions 207 and/or 221, but not at positions 213, 218, and 222, may be responsible for such a difference. To test this possibility, amino acids at positions 207 and 221 of mGSTAl were mutated with the equivalent residues of mGSTA2. Kinetic analysis of the wild type and the mutant enzymes revealed that both methionine-207 and isoleucine-221 are critical for higher activity of mGSTA1-1 toward (+)-anti-BPDE compared with that of mGSTA2-2.  相似文献   

3.
X Ji  J Blaszczyk  B Xiao  R O'Donnell  X Hu  C Herzog  S V Singh  P Zimniak 《Biochemistry》1999,38(32):10231-10238
Two variants of human class pi glutathione (GSH) S-transferase 1-1 with either isoleucine or valine in position 104 (hGSTP1-1[I104] and hGSTP1-1[V104]) have distinct activity toward (+)-anti-7, 8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene [(+)-anti-BPDE]. To elucidate their structure-function relationship, we determined the crystal structures of the two variants in complex with GSBpd, the GSH conjugate of (+)-anti-BPDE, at 2.1 and 2.0 A resolution, respectively. The crystal structures reveal that residue 104 in the xenobiotic substrate-binding site (H-site) dictates the binding modes of the product molecule GSBpd with the following three consequences. First, the distance between the hydroxyl group of Y7 and the sulfur atom of GSBpd is 5.9 A in the hGSTP1-1[I104].GSBpd complex versus 3.2 A in the V104 variant. Second, one of the hydroxyl groups of GSBpd forms a direct hydrogen bond with R13 in hGSTP1-1[V104].GSBpd; in contrast, this hydrogen bond is not observed in the I104 complex. Third, in the hydrophilic portion of the H-site of the I104 complex, five H-site water molecules [Ji, X., et al. (1997) Biochemistry 36, 9690-9702] are observed, whereas in the V104 complex, two of the five have been displaced by the Bpd moiety of GSBpd. Although there is no direct hydrogen bond between Y108 (OH) and the hydroxyl groups of GSBpd, indirect hydrogen bonds mediated by water molecules are observed in both complexes, supporting the previously suggested role of the hydroxyl group of Y108 as an electrophilic participant in the addition of GSH to epoxides.  相似文献   

4.
Pal A  Gu Y  Pan SS  Ji X  Singh SV 《Biochemistry》2001,40(24):7047-7053
The molecular basis for catalytic differences between structurally closely related murine class alpha glutathione (GSH) transferases mGSTA1-1 and mGSTA2-2 in the GSH conjugation of anti-diol epoxide isomers of benzo[c]phenanthrene (anti-B[c]PDE) was investigated. GSH conjugation of both (-)- and (+)-enantiomers of anti-B[c]PDE was observed in the presence of mGSTA1-1 (60 and 40% GSH conjugation, respectively), whereas mGSTA2-2 exhibited a preference for the (-)-anti-isomer (>97%). In addition, the specific activity of mGSTA2-2 toward the (-)-anti-B[c]PDE isomer was relatively higher than that of mGSTA1-1. The amino acid sequences of mGSTA1-1 and mGSTA2-2 differ at 10 positions that are distributed in three sections. Section I contains amino acid residues in positions 65 and 95; section II contains residues in positions 157, 162, and 169, and section III contains residues in positions 207, 213, 218, 221, and 222. Enzyme activity measurements with chimeras of mGSTA1-1 and mGSTA2-2 revealed that amino acid substitutions in section III account for their differential enantioselectivity and catalytic activity toward anti-B[c]PDE. Site-directed mutagenesis of amino acid residues in section III of mGSTA2-2 with corresponding residues of mGSTA1-1 followed by activity measurements of the wild type and mutated enzymes indicates that leucine 207 and phenylalanine 221 may be critical for the high catalytic activity of mGSTA2-2 toward (-)-anti-B[c]PDE. Molecular modeling studies demonstrated that the active site of mGSTA1-1 accommodates both enantiomers of anti-B[c]PDE, whereas the (-)-anti-isomer interacts more favorably with active site residues in mGSTA2-2. The results of this study clearly indicate that amino acid substitutions in the C-terminal region contribute to catalytic differences between mGSTA1-1 and mGSTA2-2 with respect to anti-B[c]PDE.  相似文献   

5.
The ultimate diol epoxide carcinogens derived from polycyclic aromatic hydrocarbons, such as benzo[a]pyrene (BP), are metabolized primarily by glutathione (GSH) conjugation reaction catalyzed by GSH transferases (GSTs). In human liver and probably lung, the alpha class GSTs are likely to be responsible for the majority of this reaction because of their high abundance. The catalytic efficiency for GSH conjugation of the carcinogenic (+)-anti-benzo[a]pyrene-7,8-diol-9,10-epoxide [(+)-anti-BPDE] is more than 5-fold higher for hGSTA1-1 than for hGSTA2-2. Here, we demonstrate that mutation of isoleucine-11 of hGSTA2-2, a residue located in the hydrophobic substrate-binding site (H-site) of the enzyme, to alanine (which is present in the same position in hGSTA1-1) results in about a 7-fold increase in catalytic efficiency for (+)-anti-BPDE-GSH conjugation. Thus, a single amino acid substitution is sufficient to convert hGSTA2-2 to a protein that matches hGSTA1-1 in its catalytic efficiency. The increased catalytic efficiency of hGSTA2/I11A is accompanied by greater enantioselectivity for the carcinogenic (+)-anti-BPDE over (-)-anti-BPDE. Further remodeling of the H-site of hGSTA2-2 to resemble that of hGSTA1-1 (S9F, I11A, F110V, and S215A mutations, SIFS mutant) results in an enzyme whose catalytic efficiency is approximately 13.5-fold higher than that of the wild-type hGSTA2-2, and about 2.5-fold higher than that of the wild-type hGSTA1-1. The increased activity upon mutations can be rationalized by the interactions of the amino acid side chains with the substrate and the orientation of the substrate in the active site, as visualized by molecular modeling. Interestingly, the catalytic efficiency of hGSTA2-2 toward (-)-anti-BPDE was increased to a level close to that of hGSTA1-1 upon F110V, not I11A, mutation. Similar to (+)-anti-BPDE, however, the SIFS mutant was the most efficient enzyme for GSH conjugation of (-)-anti-BPDE.  相似文献   

6.
Benzo[a]pyrene-(7R,8S)-diol (9S,10R)-epoxide [(+)-anti-BPDE] is believed to be the activated form of the widely spread environmental pollutant benzo[a]pyrene. Glutathione (GSH) S-transferase (GST)-catalyzed conjugation of (+)-anti-BPDE with GSH is an important mechanism in its cellular detoxification. Here, we report that the GSH conjugate of (+)-anti-BPDE [(-)-anti-BPD-SG] is a potent inhibitor (K(i) 15 microM) of class Mu human GST isoenzyme, which, among human liver GSTs, is a highly efficient detoxifier of (+)-anti-BPDE. Thus, the inhibition of GST activity by (-)-anti-BPD-SG may hinder GSH conjugation of (+)-anti-BPDE, unless the conjugate is metabolized and/or eliminated. The results of the present study show that gamma-glutamyltranspeptidase (gamma-GT) can metabolize (-)-anti-BPD-SG at a rate of about 0.29 nmol/min/mg protein. Our studies also show that (-)-anti-BPD-SG is transported across the human canalicular liver plasma membrane (cLPM) in an ATP-dependent manner at a rate of about 0.33 nmol/min/mg protein. The ATP-dependent transport of (-)-anti-[(3)H]BPD-SG across human cLPM follows Michaelis-Menten kinetics (K(m) 84 microM; V(max) 0.33 nmol/min/mg). In conclusion, the results of the present study suggest that both gamma-GT-mediated metabolism and ATP-dependent canalicular transport may be important steps in overall detoxification of (+)-anti-BPDE in the human liver.  相似文献   

7.
(7S,8S)--Dihydroxy--7,8--dihydrobenzo[a]pyrene ((+)-BP-7,8-diol) is epoxidized to (7S,8R)-dihydroxy-(9S,10R)-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene ((+)-syn-BPDE) by cytochrome P-450 isoenzymes and to (7S,8R)-dihydroxy-(9R,10S)-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene ((-)-anti-BPDE) by peroxyl free radicals. 32P postlabeling analysis of the diastereomeric BPDE-DNA adducts was used to investigate the pathways of (+)-BP-7,8-diol oxidation in mouse skin in vivo. The pattern of deoxynucleoside 3',5'-bisphosphate adducts in epidermal scrapings from female CD-1 mice indicated that cytochrome P-450 was the major oxidant. Similar results were obtained when the tumor-promoting phorbol ester tetradecanoylphorbolacetate (TPA) was coadministered with (+)-BP-7,8-diol. However, when animals were pretreated with TPA 24 h before coadministration of TPA and (+)-BP-7,8-diol, the pattern of BPDE-DNA adducts indicated that peroxyl radicals made a major contribution to (+)-BP-7,8-diol epoxidation. Peroxyl radical-dependent epoxidation was maximal when the time between the two TPA administrations was 24-72 h. No increase in (-)-anti-BPDE-DNA was observed when the non-tumor-promoting phorbol ester 4-O-methyl-TPA was substituted for TPA. The calcium ionophore A23187 stimulated peroxyl radical generation when substituted for the first, but not the second, TPA treatment. The antiinflammatory steroid fluocinolone acetonide inhibited (-)-anti-BPDE-DNA adduct formation when coadministered with the first but not the second TPA treatment. These findings demonstrate the existence of two independent pathways of metabolic activation of (+)-BP-7,8-diol in mouse epidermis, one dependent on cytochrome P-450 and the other dependent on peroxyl free radicals. The results also suggest that repetitive topical administration of tumor-promoting phorbol esters remodels epidermal metabolism leading to a significant increase in free radical generation.  相似文献   

8.
Levels of reduced glutathione (GSH) in C3H/10T1/2 cells were selectively altered to determine what quantitative role GSH transferase-catalyzed conjugation plays in regulating the cytotoxic effects of benzo(a)pyrene anti-7,8-dihydrodiol 9,10-epoxide (r-7,t-8-dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo(a)pyrene, anti-diol epoxide). A 65% decrease in 10T1/2 cell GSH content from 0.16 mM (control cell GSH concentration) to 0.06 mM was accompanied by a 46% decrease in the anti-diol epoxide LD80; a 98% increase in GSH content resulted in a 44% increase in anti-diol epoxide LD80. This nonlinear relationship between changes in cellular GSH concentration and anti-diol epoxide LD80 was directly relatable to the nonlinear change in the rate of anti-diol epoxide conjugation which was catalyzed by 10T1/2 cell GSH transferases. Purified 10T1/2 cell cytosol catalyzed the GSH conjugation of anti-diol epoxide to yield a GSH conjugation product with a distinct UV absorbance spectrum; the apparent GSH Km for this cell cytosol-catalyzed reaction was 0.20 mM. Variations in the cellular GSH concentration around the GSH Km resulted in a nonlinear change in the amount of anti-diol epoxide-GSH conjugate formed, and a reciprocal change in the amount of free anti-diol epoxide available for cytotoxic alkylation events. These results clarify in quantitative, biochemical terms how GSH transferase-catalyzed conjugation can regulate the level of an electrophilic carcinogen metabolite in a biological system.  相似文献   

9.
Incubation of 15-hydroperoxy-5,8,11,13-eicosatetraenoic acid with ram seminal vesicle microsomes (RSVM) triggers the oxygenation of trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene (BP-7,8-diol). The principal oxidation products are 7,8,9,10-tetrahydroxy-7,8,9,10-tetrahydrobenzo[a]pyrenes which are non-enzymatic hydrolysis products of r-7,t-8-dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene. At short incubation times, an additional product is isolated which is identified as r-7,t-8,t-9-trihydroxy-c-10-methoxy-7,8,9,10-tetrahydrobenzo[a]pyrene. This product appears to arise by solvolysis of the extracted diolepoxide during high performance liquid chromatography using methanol-water solvent systems. The incubation of 18O-labeled 15-hydroperoxy-5,8,11,13-eicosatetraenoic acid with BP-7,8-diol and RSVM leads to very little incorporation of 18O into the stable solvolysis products (analyzed by gc-ms of their peracetates). Parallel incubations conducted with 16O-labeled hydroperoxide under an 18O2 atmosphere indicate that the principle source of the epoxide oxygen is molecular oxygen.  相似文献   

10.
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   

11.
Bay-region diol epoxides are considered the putative ultimate carcinogens of polynuclear aromatic hydrocarbons. However, the results of studies on tumorigenesis and DNA binding of benzo[a]pyrene (BP) and its bay-region diol epoxide, (+)-trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyren e [(+)-anti-BPDE] suggest that, in addition to anti-BPDE, other reactive metabolite(s) of BP may also be involved in BP-induced carcinogenesis. Recent studies have demonstrated that 3-hydroxy-trans-7,8-dihydroxy-anti-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a ]pyrene (anti-BPTE) is another highly reactive metabolite of BP. In order to identify syn- and anti-BPTE-derived DNA adducts and their base selectivity, we synthesized both compounds by two different methods and reacted in vitro with calf thymus DNA and individual nucleotides. The resultant adducts were analyzed by nuclease P1-enhanced 32P-postlabeling. Anti-BPTE produced three major and several minor adducts with DNA; dAp and dGp were the preferred substrates, while dCp and dTp were the least reactive. In contrast, syn-BPTE produced two major adducts each with DNA and dGp; dAp generated only one adduct. Co-chromatography of anti-BPTE-derived DNA adducts with those of mononucleotide adducts revealed that the major adducts in DNA were guanine derived. Further, co-chromatographic results revealed that the anti-BPTE-DNA adducts were distinctly different from that of anti-BPDE-DNA adducts. These observations indicate that both syn- and anti-BPTE can react with DNA bases and these DNA adducts may also contribute to BP-induced carcinogenesis.  相似文献   

12.
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   

13.
(±)-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   

14.
The syn- and anti-diastereoisomers of trans-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) were further metabolized by rat liver microsomes obtained from 3-methylcholanthrene(MC)-pretreated rats and NADPH to reactive intermediates, presumably 1,7,8- and 3,7,8-trihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrenes (triol-epoxides), that bound to macromolecules or decomposed to products consistent with pentahydroxy derivatives of benzo[a]pyrene (BP-pentols). Three major metabolites of syn-BPDE and four major metabolites of anti-BPDE were isolated by high performance liquid chromatography and characterized by spectroscopic techniques. When fluorescence spectroscopy was employed all metabolites exhibited very similar spectral properties and showed substantial shifts in excitation and emission maxima to longer wavelengths when measured under alkaline conditions, consistent with the presence of a phenolic hydroxyl group. Furthermore, the spectral properties of the metabolites from syn- and anti-BPDE were similar to those of 1-hydroxypyrene. Previous data from this laboratory together with the data presented in this study thus strongly suggest that further metabolism of BPDE involves hydroxylation at the 1- and 3-positions to yield the corresponding triol-epoxides and various BP-pentols. The pentols could also be formed by incubating tetrols derived from syn- and anti-BPDE with microsomes and NADPH. However, the rate of formation of pentols from the BP-tetrols was much slower than the rate of further metabolism of BPDE. Accordingly, the major route of BP-pentol formation is likely to be via the intermediate formation of triol-epoxides. Isolated liver nuclei from MC-pretreated rats were also found to catalyze the activation of anti-BPDE in presence of NADPH to reactive intermediates. This resulted in a substantial increase in binding to histone and non-histone proteins, with a concomitant decrease in binding to DNA. No qualitative change in the distribution of DNA-bound products of anti-BPDE could be demonstrated as a result of the further metabolism of anti-BPDE.  相似文献   

15.
Quantitation of protein-benzo[a]pyrene adducts represent a more sensitive analysis method than quantitation of benzo[a]pyrene-DNA adducts. By accurate analysis of benzo[a]pyrene-protein adducts several different molecular adduct forms can be studied. Male Wistar rats were injected i.p. with benzo[a]pyrene, and serum albumin was isolated and subjected to acid hydrolysis at 90 degrees C for 3 h. The hydrolysate was analyzed by HPLC with fluorescence detection. The HPLC profiles obtained after albumin hydrolysis from benzo[a]pyrene exposed animals were compared to similar HPLC profiles from in vitro adducted bovine serum albumin (BSA) and direct hydrolysis of both r-10,t-9-dihydrodiol-c-7,8-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (syn-BPDE-III) and r-10,t-9-t-dihydrodiol-t-7,8-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (anti-BPDE-III). After acid hydrolysis of albumin from benzo[a]pyrene exposed rats, 6 fluorescent peaks were separated. Four of the peaks were isomers of benzo[a]pyrene-tetrahydrotetrols, (+/-)-benzo[a]pyrene-r-7,t-8,9,10-tetrahydrotetrol, (+/-)-benzo[a]pyrene-r-7,t-8,9,c-10-tetrahydrotetrol, (+/-)-benzo[a]pyrene-r-7,t-8,c-9,t-10-tetrahydrotetrol and (+/-)-benzo[a]pyrene-r-7,t-8,c-9,10-tetrahydrotetrol. In addition we found two fluorescent peaks, named X1 and X2 with retention times similar to the benzo[a]pyrene-tetrols. The unknown fluorescent peaks reacted similar to the four known tetrols in both dose response experiments and time course experiments. Fluorescent material with retention times equal to X1 and X2 were found after acid hydrolysis of syn-BPDE-III and anti-BPDE-III in acid and in hydrolysates from BSA treated in vitro with syn-BPDE-III and anti-BPDE-III. The ratio X1/X2 was relatively constant indicating epimerization equilibrium between these to species. Synchronous fluorescence analysis of fractions containing X1 or X2 from both in vivo and in vitro experiments showed fluorescence spectra characteristic of benzo[a]pyrene tetrols using a wavelength difference of 34 nm.  相似文献   

16.
The frequencies of the induction of sister-chromatid exchanges and the levels of deoxyribonucleoside-hydrocarbon adducts formed in Chinese hamster ovary cells that had been treated with either dihydrodiols or a diol-epoxide derived from polycyclic aromatic hydrocarbons were determined. Up to 6-fold increases in the incidence of these exchanges were observed when the cells were treated either with the dihydrodiols, trans-3,4-dihydro-3,4-dihydroxy-7-methylbenz[a]anthracene,trans-7,8-dihydro-7,8-dihydroxybenzo[a]pyrene or the diol-epoxide, (±)-r-7, t-8dihydroxy-t-9,10-oxy-7,8,9,10-tetrahydrobenzo[a] pyrene but when the cells were transferred to media free of these compounds, there were rapid reductions in the frequency of these exchanges. When the exchanges were induced by the diol-epoxide, the decreases in frequency were paralleled by decreases in the levels of deoxyribonucleoside-diol-epoxide adducts that were present in hydrolysates of DNA isolated from the cells. There thus appears to be a close relationship between the frequency of sister-chromatid exchanges and the levels of deoxyribonucleoside-diol-epoxide adduct formation.  相似文献   

17.
18.
Hemoglobin plus H2O2 oxidizes trans-stilbene to trans-stilbene oxide, cis-stilbene to cis- and trans-stilbene oxide, and trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene to anti-trans-7,8,9,10-tetrahydroxy-7,8,9,10-tetrahydrobenzo[a]pyrene. Oxidation of cis- and trans-stilbene to the corresponding cis- and trans-epoxides proceeds exclusively with incorporation of oxygen from the peroxide. Oxidation of cis-stilbene to the trans-epoxide, however, proceeds without detectable incorporation of oxygen from the peroxide and partial incorporation of oxygen from O2. The epoxidations in which stereochemistry is conserved thus appear to involve ferryl oxygen transfer, whereas the epoxidations in which stereochemistry is inverted are proposed to involve protein-mediated cooxidation [Ortiz de Montellano, P.R., & Catalano, C.E. (1985) J. Biol. Chem. 260, 9265-9271] and possibly electron abstraction-water addition. The epoxidation of trans-7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene incorporates oxygen from H2O2 and H2O but not O2. The oxidation of this substrate is thus consistent with ferryl oxygen transfer and electron abstraction but not protein-mediated cooxidation.  相似文献   

19.
The repair of human DNA after damage by known and potential metabolites of benzo(a)pyrene has been examined utilizing the bromodeoxyuridine photolysis assay. Repair was characterized as either ultraviolet (“long”) or ionizing radiation type (“short”) repair utilizing normal cells and cells deficient in ultraviolet-type repair endonuclease from a patient with xeroderma pigmentosum (XP). We have found that only (±)-7β,8-dihydroxy-9β,-10β-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (BP diol epoxide 1) and its disastereomer, (±)-7β,8,-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene (BP diol epoxide 2) elicit damage to DNA which is recognizable by the ultraviolet excision repair system in normal human cells. Benzo(a)pyrene 4,5-, 9,10-, 11,12-oxides do not elicit damage which is repairable by this repair system. The 1,2-diol-3,4-epoxides from naphthalene have no measurable activity in our assay. These results indicate that both the benzo(a)pyrene ring structure and the diol epoxide groups are important in causing the damage to DNA which is repairable by the ultraviolet excision repair system. These results parallel the reported high mutagenic activity of these compounds and support the concept that benzo(a)pyrene 7,8-diol-9,10-epoxides may be the ultimate, metabolically activated forms of benzo(a)pyrene.  相似文献   

20.
Mutations induced by polycyclic aromatic hydrocarbons (PAH) are expected to be produced when error-prone DNA replication occurs across unrepaired DNA lesions formed by reactive PAH metabolites such as diol epoxides. The mutagenicity of the two PAH-diol epoxides (+)-anti-7,8-dihydroxy-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene (BPDE) and (+/-)-anti-11,12-dihydroxy-13,14-epoxy-11,12,13,14-tetrahydrodibenzo[a,l]pyrene (DBPDE) was compared in nucleotide excision repair (NER) proficient and deficient hamster cell lines. We applied the (32)P-postlabelling assay to analyze adduct levels and the hprt gene mutation assay for monitoring mutations. It was found that the mutagenicity per target dose was 4 times higher for DBPDE compared to BPDE in NER proficient cells while in NER deficient cells, the mutagenicity per target dose was 1.4 times higher for BPDE. In order to investigate to what extent the mutagenicity of the different adducts in NER proficient cells was influenced by repair or replication bypass, we measured the overall NER incision rate, the rate of adduct removal, the rate of replication bypass and the frequency of induced recombination in the hprt gene. The results suggest that NER of BPDE lesions are 5 times more efficient than for DBPDE lesions, in NER proficient cells. However, DBPDE adducts block replication more efficiently and also induce 6 times more recombination events in the hprt gene than adducts of BPDE, suggesting that DBPDE adducts are, to a larger extent, bypassed by homologous recombination. The results obtained here indicate that the mutagenicity of PAH is influenced not only by NER, but also by replication bypass fidelity. This has been postulated earlier based on results using in vitro enzyme assays, but is now also being recognized in terms of forward mutations in intact mammalian cells.  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