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.
Nitrogen-containing analogs of chrysene, 1,10-diazachrysene (1,10-DAC) and 4,10-DAC, were tested for mutagenicity in Salmonella typhimurium TA100 in the presence of rat liver S9 and human liver microsomes to investigate the effect of nitrogen-substitution. Although these DACs could not be converted to the bay-region diol epoxide because of their nitrogen atoms in the bay-region epoxide or diol moiety, DACs were mutagenic in the Ames test with rat liver S9. Both DACs also showed mutagenicity in the Ames test using pooled human liver microsomes, although chrysene itself was not mutagenic in the presence of pooled human liver microsomes. The mutagenicity of DACs (50nmol/plate) in Ames tests using human liver microsome preparations from 10 individuals was compared with cytochrome P450 (CYP) activity in each microsome preparation to investigate the CYP isoform involved in the activation of DACs to the genotoxic forms. The numbers of induced revertants obtained by 1,10-DAC varied 6.2-folds (109-680) and those by 4,10-DAC 4.8-folds (155-751) among the 10 individuals. The number of induced revertants obtained by 1,10-DAC significantly correlated with the CYP1A2-selective catalytic activity (r=0.84, P<0.01) in each microsome preparation. On the other hand, the number of induced revertants obtained by 4,10-DAC significantly correlated with the combined activity of CYP2A6 and 1A2 (CYP2A6+0.51xCYP1A2; r=0.75, P<0.01). However, in Ames tests using microsomes from insect cells expressing various human CYP isoforms, the mutagenicity of 1,10-DAC was induced only by recombinant human CYP1A2, whereas both recombinant human CYP2A6 and 1A2 contributed to the mutagenicity of 4,10-DAC. These results suggest that 1,10-DAC shows the mutagenicity through involvement of CYP1A2 in human liver, and 4,10-DAC does so through both CYP2A6 and 1A2. In conclusion, our results suggested that the difference in the nitrogen-substituted position in the chrysene molecule might affect the mutagenic activity through influencing the ratio of participation of the metabolic activation enzyme isoforms of CYP.  相似文献   

2.
We have examined the selectivity of rat liver microsomal epoxide hydrolase (EC 3.3.2.3) toward all of the possible positional isomers of benzo-ring diol epoxides and tetrahydroepoxides of benz[a]anthracene, as well as the 1,2-diol 3,4-epoxides of triphenylene. This set includes compounds with no bay region in the vicinity of the benzo-ring, a bay-region diol group, a bay-region epoxide group, and (for the triphenylene derivatives) both a bay-region diol and a bay-region epoxide. In all cases where both the tetrahydroepoxides and the corresponding diol epoxides were examined, there is a large retarding effect of hydroxyl substitution on the rate of the enzyme-catalyzed hydration. When the tetrahydroepoxides are fair or poor substrates (epoxide group in the 1,2-, 8,9-, or 10,11-position), the additional retardation introduced by adjacent hydroxyl groups causes the enzyme-catalyzed hydrolysis of the corresponding diol epoxides to be insignificantly slow or nonexistent. In contrast, a benz[a]anthracene derivative with an epoxide group in the 3,4-position, (-)-tetrahydrobenz[a]anthracene (3R,4S)-epoxide, has been identified as the best substrate known for epoxide hydrolase, with a Vmax at 37 degrees C and pH 8.4 of 6800 nmol/min/mg of protein, and the two diastereomeric (+/-)-benz[a]anthracene 1,2-diol 3,4-epoxides, unlike all the other diol epoxides examined to date, are moderately good substrates for epoxide hydrolase. This novel observation is accounted for by the fact that the very high reactivity of the tetrahydrobenz[a]anthracene 3,4-epoxide system towards epoxide hydrolase is large enough to overcome a kinetically unfavorable effect of hydroxyl substitution. The enantioselectivity and positional selectivity of the enzyme have been determined for the tetrahydro-1,2- and -3,4-epoxides of benz[a]anthracene as well as the 1,2-diol 3,4-epoxides. When the epoxide is located in the 3,4-position, the benzylic carbon is the preferred site of attack, whereas for the enantiomers of the bay-region tetrahydro-1,2-epoxides, the chemically less reactive non-benzylic carbon is preferred. The regio- and enantioselectivity of epoxide hydrolase are discussed in terms of a possible model for the hydrophobic binding site of this enzyme.  相似文献   

3.
The modes of reaction of the tumorigenic bay region diol epoxide anti-BADE [+/-)-trans-3,4-diol-anti-1,2-epoxy-1,2,3,4-tetrahydrobenz[a]anthr acene) and the less potent tumor initiating diastereomer syn-BADE [+/-)-trans-3,4-diol-syn-1,2-epoxy-1,2,3,4-tetrahydrobenz[a]anthra cene) with native, double-stranded DNA were compared. The bay-region diol epoxide derived from 3-methylcholanthrene (3-MCDE, racemic trans-9,10-diol-anti-7,8-epoxy-7,8,9,10-tetrahydromethylcholanthrene+ ++) was included in this study in order to assess the effects of the methyl and methylene substituents on the reactivity with DNA. Utilizing linear dichroism and other spectroscopic methods, it is shown that all three diol epoxides forn non-covalent complexes with DNA. The diastereomers anti-BADE and syn-BADE form intercalative physical complexes, but the association constant K of the syn-diastereomer is about 6-7 times smaller than for anti-BADE; this effect is ascribed to the bulky quasi-diaxial conformation of the diol epoxide ring in the syn diastereomer. The value of K (4000 M-1) is similar for anti-BADE and 3-MCDE, although the latter is not intercalated in the classical sense since the short axis of the molecule is tilted closer to the axis of the DNA double helix. The conformations of the covalent DNA adducts are interpreted in terms of a quasi-intercalative conformation (site I), and a conformation in which the long axes of the polycyclic molecules are tilted closer to the axis of the helix (site II). Both tumorigens, anti-BADE and 3-MCDE, undergo a marked re-orientation from a non-covalent site I to a covalent site II conformation upon binding chemically with the DNA bases, although a small fraction of the covalent anti-BADE adducts remains quasi-intercalated; in contrast, the alkyl substituents in 3-MCDE not only prevent the formation of intercalative physical complexes, but also the formation of site I covalent adducts. In the case of the less tumorigenic syn-BADE, both the non-covalent complexes and the covalent adducts are of the site I-type. The bay-region diol epoxide of benz[a]anthracene and of 3-methylcholanthrene display a similar pattern of reactivities and covalent adduct conformations as the bay region diol epoxide derivatives of benz[a]pyrene, suggesting that adduct conformation might be an important factor in determining the levels of mutagenic and tumorigenic activities of this class of compounds.  相似文献   

4.
Phenanthrene, a simplest angular polycyclic aromatic hydrocarbon with a bay-region in its molecule, is reported to be non-mutagenic, although most angular (non-linear) polycyclic aromatic hydrocarbons, such as benzo[a]pyrene and chrysene, are known to show genotoxicity after metabolic transformation into a bay-region diol epoxide. On the other hand, benzo[f]quinoline (BfQ), benzo[h]quinoline (BhQ), and 1,7-phenanthroline (1,7-Phe), which are all aza-analogs of phenanthrene, are mutagenic in the Ames test using Salmonella typhimurium TA100 in the presence of a rat liver S9 fraction. In this report, we undertook to investigate the in vivo mutagenicity of BfQ, BhQ and 1,7-Phe by an in vivo mutation assay system using the lacZ transgenic mouse (Muta Mouse). BfQ and BhQ only slightly induced mutation in the liver and lung, respectively. BfQ- and BhQ-induced cII mutant spectra showed no characteristics compared with that of the control. These results suggest that the in vivo mutagenicities of BfQ and BhQ were equivocal. On the other hand, 1,7-Phe induced a potent mutation in the liver and a weak mutation in the lung. Furthermore 1,7-Phe depressed the G:C to A:T transition and increased the G:C to C:G transversion in the liver like quinoline, a hepatomutagen possessing the partial structure of 1,7-Phe, compared with the spontaneous mutation spectrum. These results suggest that the in vivo mutagenicity of 1,7-Phe might be caused by the same mechanism as that of quinoline, which induced the same mutational spectrum change (G:C to C:G transversion).  相似文献   

5.
The bacterial mutagenic response (Ames-assay, Salmonella typhimurium strain TA98+/-S9-mix) of a series of monocyclopenta-fused polycyclic aromatic hydrocarbons (CP-PAHs) identified in combustion exhausts, viz. cyclopenta[cd]pyrene (1), acephenanthrylene (2), aceanthrylene (3) and cyclopenta[hi]chrysene (4), is re-evaluated. The mutagenic effects are compared with those exerted by the corresponding partially hydrogenated derivatives, 3,4-dihydrocyclopenta[cd]pyrene (5), 4,5-dihydroacephenanthrylene (6), 1,2-dihydroaceanthrylene (7) and 4,5-dihydrocyclopenta[hi]chrysene (8). It is shown that the olefinic bond of the externally fused five-membered ring of 1, 3 and 4 is of importance for a positive mutagenic response. In contrast, whilst CP-PAH 2 is found inactive, its dihydro analogue (6) shows a weak metabolism-dependent response. The importance of epoxide formation at the external olefinic bond in the five-membered ring is substantiated by the bacterial mutagenic response of independently synthesized cyclopenta[cd]pyrene-3,4-epoxide (9), acephenanthrylene-4,5-epoxide (10), aceanthrylene-1,2-epoxide (11) and cyclopenta[hi]chrysene-4,5-epoxide (12). Their role as ultimate, active mutagenic forms, when CP-PAHs 1, 3 and 4 exhibit a positive mutagenic response, is confirmed. Semi-empirical Austin Model 1 (AM1) calculations on the formation of the CP-arene oxides (9-12) and their conversion into the monohydroxy-carbocations (9a-12a and 9b-12b) via epoxide-ring opening support our results. For 2 and 4, which also possess a bay-region besides an annelated cyclopenta moiety, the calculations rationalize that epoxidation at the olefinic bond of the cyclopenta moiety is favoured.  相似文献   

6.
Metabolism of 3H-labeled (+)-(S,S)- and (-)-(R,R)-1,2-dihydrodiols of triphenylene by rat liver microsomes and 11 purified isozymes of cytochrome P450 in a reconstituted monooxygenase system has been examined. Although both enantiomers were metabolized at comparable rates, the distribution of metabolites between phenolic dihydrodiols and bay-region, 1,2-diol 3,4-epoxide diastereomers varied substantially with the different systems. Treatment of rats with phenobarbital (PB) or 3-methylcholanthrene (MC) caused a slight reduction or less than a twofold increase, respectively, in the rate of total metabolism (per nanomole of cytochrome P450) of the enantiomeric dihydrodiols compared to microsomes from control rats. Among the 11 isozymes of cytochrome P450 tested, only cytochromes P450c (P450IA1) and P450d (P450IA2) had significant catalytic activity. With either enantiomer of triphenylene 1,2-dihydrodiol, both purified cytochrome P450c (P450IA1) and liver microsomes from MC-treated rats formed diol epoxides and phenolic dihydrodiols in approximately equal amounts. Purifed cytochrome P450d (P450IA2), however, formed bay-region diol epoxides and phenolic dihydrodiols in an 80:20 ratio. Interestingly, liver microsomes from control or PB-treated rats produced only diol epoxides and little or no phenolic dihydrodiols. The diol epoxide diastereomers differ in that the epoxide oxygen is either cis (diol epoxide-1) or trans (diol epoxide-2) to the benzylic 1-hydroxyl group. With either purified cytochromes P450 (isozymes c or d) or liver microsomes from MC-treated rats, diol epoxide-2 is favored over diol epoxide-1 by at least 4:1 when the (-)-enantiomer is the substrate, while diol epoxide-1 is favored by at least 5:1 when the (+)- enantiomer is the substrate. In contrast, with liver microsomes from control or PB-treated rats, formation of diol epoxide-1 relative to diol epoxide-2 was favored by at least 2:1 regardless of the substrate enantiomer metabolized. This is the first instance where the ratio of diol epoxide-1/diol epoxide-2 metabolites is independent of the dihydrodiol enantiomer metabolized. Experiments with antibodies indicate that a large percentage of the metabolism by microsomes from control and PB-treated rats is catalyzed by cytochrome P450p (P450IIIA1), resulting in the altered stereoselectivity of these microsomes compared to that of the liver microsomes from MC-treated rats.  相似文献   

7.
We have previously reported the in vivo mutagenicity of aza-polycyclic aromatic hydrocarbons (azaPAHs), such as quinoline, benzo[f]quinoline, benzo[h]quinoline, 1,7-phenanthroline and 10-azabenzo[a]pyrene. The 1,10-diazachrysene (1,10-DAC) and 4,10-DAC, nitrogen-substituted analogs of chrysene, were shown to exhibit mutagenicity in Salmonella typhimurium TA100 in the presence of rat liver S9 and human liver microsomes in our previous report, although DACs could not be converted to a bay-region diol epoxide, the ultimate active form of chrysene, because of their nitrogen atoms. In the present study, we tested in vivo mutagenicity of DACs compared with chrysene using the lacZ transgenic mouse (Mutatrade markMouse) to evaluate the effect of the nitrogen substitution. DACs- and chrysene-induced mutation in all of the six organs examined (liver, spleen, lung, kidney, bone marrow and colon). The mutant frequencies obtained with chrysene showed only small differences between the organs examined and ranged from 1.5 to 3 times the spontaneous frequency. The 4,10-DAC was more mutagenic than chrysene in all the organs tested. The highest lacZ mutation frequency was observed in the lung of 4,10-DAC-treated mice and it was 19 and 6 times the spontaneous frequency and the frequency induced by chrysene, respectively. The 1,10-DAC induced lacZ mutation in the lung with a frequency 4.3- and 1.5-fold higher than in the control and chrysene-treated mice, respectively, although the mutant frequencies in the other organs of 1,10-DAC-treated mice were almost equivalent to those of chrysene-treated mice. Not only chrysene but also DACs depressed the G:C to A:T transition and increased the G:C to T:A transversion in the liver and lung. These results suggest that the two types of nitrogen substitutions in the chrysene structure may enhance mutagenicity in the mouse lung, although they showed no difference in the target-organ specificity and the mutation spectrum.  相似文献   

8.
The polynuclear aromatic hydrocarbon chrysene undergoes a bioalkylation substitution reaction in vitro, in rat liver cytosol preparations, and in vivo, in rat dorsal subcutaneous tissue to yield 6-methylchrysene as a metabolite. In addition, both 5-methyl- and 6-methylchrysene were found to undergo a dealkylation reaction in these tissues to yield chrysene as well as both a biooxidation reaction to yield the corresponding hydroxyalkyl substituted chrysene and a bioalkylation reaction to give a dimethyl substituted chrysene. 5-Methylchrysene enzymatically cyclized to the 4,5-methylenechrysene derivative, an analog of benzo[a]pyrene in these tissues. 5,6-Dimethylchrysene was metabolized to monomethyl chrysenes, chrysene, and the hydroxyalkyl substituted chrysenes. The results suggest that chemical or biochemical substitution of a methyl group at the center of highest biochemical reactivity may be a necessary step in the metabolic activation and carcinogenicity of these compounds and their methylene bridged metabolites.  相似文献   

9.
Many anaerobic bacteria fix CO2 via the Wood pathway of acetyl-CoA synthesis. Carbon monoxide dehydrogenase (CODH), also called acetyl-CoA synthase, accepts the methyl group from the methylated corrinoid/iron-sulfur protein (C/Fe-SP), binds a carbonyl group from CO, CO2, or the carboxyl of pyruvate, and binds coenzyme A. Then CODH catalyzes the synthesis of acetyl-CoA from these enzyme-bound groups. Here, we have characterized the methyl transfer steps involved in acetyl-CoA synthesis. We have studied the reactions leading to methylation of CODH by methyl iodide and shown an absolute requirement of the C/Fe-SP in this reaction. In addition, we have discovered and partly characterized two previously unknown exchange reactions catalyzed by CODH: between the methylated C/Fe-SP and methylated CODH and between methylated CODH and the methyl moiety of acetyl-CoA. We have performed these two exchange reactions, methylation of the C/Fe-SP, and methylation of CODH at controlled potentials. The rates of all these reactions except the exchange between methylated C/Fe-SP and methylated CODH are accelerated (from 1 to 2 orders of magnitude) when run at low potentials. Our results provide strong evidence for a nucleophilic redox-active metal center on CODH as the initial acceptor of the methyl group from the methylated C/Fe-SP. This metal center also is proposed to be involved in the cleavage of acetyl-CoA in the reverse reaction.  相似文献   

10.
The solution structure of an 11-mer DNA duplex, d(CGGTCA*CGAGG) x d(CCTCGTGACCG), containing a 10R adduct at dA* that corresponds to the cis addition of the N(6)-amino group of dA(6) to (+)-(9S,10R)-9,10-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene was studied by 2D NMR methods. The NOESY cross-peak patterns indicate that the hydrocarbon is intercalated on the 5'-side of the modified base. This observation is the same as that observed for other oligonucleotides containing (10R)-dA adducts but opposite to that observed for the corresponding (10S)-dA adducts which are intercalated on the 3'-side of the modified base. The hydrocarbon is intercalated from the major groove without significant disruption of either the anti glycosidic torsion angle of the modified residue or the base pairing of the modified residue with the complementary residue on the opposite strand. The ensemble of 10 structures determined exhibits relatively small variations (6-15 degrees) in the characteristic hydrocarbon-base dihedral angles (alpha' and beta') as well as the glycosidic torsion angle chi. These angles are similar to those in a previously determined cis-opened benzo[a]pyrene diol epoxide-(10R)-dA adduct structure. Comparison of the present structure with the cis-opened diol epoxide adduct suggests that the absence of the 7- and 8-hydroxyl groups results in more efficient stacking of the aromatic moiety with the flanking base pairs and deeper insertion of the hydrocarbon into the helix. Relative to normal B-DNA, the duplex containing the present tetrahydroepoxide adduct is unwound at the lesion site, whereas the diol epoxide adduct structure is more tightly wound than normal B-DNA. Buckling of the adducted base pair as well as the C(5)-G(18) base pair that lies immediately above the hydrocarbon is much less severe in the present adducted structure than its cis-opened diol epoxide counterpart.  相似文献   

11.
12.
28 S ribosomal RNA from several vertebrate species contains some 68 to 70 methyl groups. Evidence described in this paper enables some 58 methyl groups to be located in the primary structure of 28 S ribosomal RNA from Xenopus laevis. Most of the locations are unambiguous but a few are currently tentative. In human 28 S ribosomal RNA the great majority of the same sites are methylated as in Xenopus, but there are a few differences between the respective methyl group distributions. The main features of the methyl group distribution are as follows. (1) All of the identified methyl groups are in conserved core regions of 28 S ribosomal RNA. (2) Methylation is much more heavily concentrated in the 3' region of the molecule than in the 5' region (in contrast to 18 S ribosomal RNA, in which there is a major cluster of 2'-O-methyl groups in the 5' region). (3) In addition to the heavily methylated 3' region, clusters of methyl groups occur elsewhere in 28 S ribosomal RNA in the vicinity of domain boundaries. For domains 3 to 6, the two ends of each domain are united in a helix and are linked to adjacent domains either directly or by short single-stranded regions. It therefore follows that the methyl groups near the boundaries of these domains come together into the same general region of the three-dimensional structure. Within this large-scale pattern of distribution, methyl groups occur in a variety of local environments, examples of which are discussed. The triply methylated sequence Am-Gm-Cm-A occurs in a short single-stranded region which links domain 3 to domain 4. Near the 3' end of domain 5 there is a cluster of 11 methyl groups including a 2'-O-methyl pseudouridine in a tract of 160 nucleotides whose sequence is totally conserved between Xenopus and man. These methyl groups are variously distributed between single-stranded regions and short or imperfect but conserved helices. A further cluster of methyl groups including the highly conserved Um-Gm-psi sequence occurs in a region of domain 6 which is implicated in peptidyl transfer. Possible relationships between methylation and other events in ribosome maturation are discussed.  相似文献   

13.
The properties of a-chymotrypsin methylated at histidine-57 were examined to explain the mechanism of this enzyme which is about 105 times less active than chymotrypsin. Studies on the protein showed (i) an alteration in the acyl and leaving group specificity, (ii) decreased binding of some protein protease inhibitors by methyl chymotrypsin, (iii) lack of dimerization of methyl chymotrypsin at low pH, (iv) decreased stability of methyl chymotrypsin in urea, (v) a larger solvent deuterium isotope effect with methyl chymotrypsin, and (vi) decreased binding of a tetrahedral intermediate analog to methyl chymotrypsin. These properties suggest that while only subtle alterations occur in the active site upon methylation of His-57, the transition state and the tetrahedral intermediate are destabilized but not to the same extent. General base catalysis remains an integral feature of the hydrolytic mechanism of the modified chymotrypsin, and the base appears to be the methylated nitrogen of the imidazole moiety of His-57.  相似文献   

14.
The steric course of the methyl group transfer to polygalacturonic acid to form the methyl ester group in pectin was studied using S-adenosylmethionine (AdoMet) carrying a methyl group made chiral by labeling with 1H, 2H, 3H, in an asymmetric arrangement. The incubation of the two diastereomers of this substrate with a particulate enzyme preparation obtained from Phaseolus aureus (mung bean) shoots gave the corresponding pectins. These were degraded in a series of stereochemically unambiguous reactions that converted the methoxy group into the methyl group of acetate, which was then analyzed for its configuration. The results indicate that the transfer of the methyl group from the sulfur of AdoMet to the oxygen of the carboxyl group proceeds with inversion of configuration of the methyl group.  相似文献   

15.
The 18 S ribosomal RNA from a variety of vertebrate species contains some 40 to 47 methyl groups. The majority of these are 2'-O-ribose substituents; the remaining few are on bases. Several lines of evidence have permitted the identification of the precise locations of the methyl groups in the primary structure of 18 S ribosomal RNA of Xenopus laevis and man. Digestion of RNA with T1 ribonuclease, followed by analysis of the methylated oligonucleotides yielded data on sequences immediately surrounding the methyl groups. Preparative hybridization of X. laevis 18 S ribosomal RNA restriction fragments of ribosomal DNA, followed by fingerprinting analysis on RNA recovered from the hybrids, allowed each methylated oligonucleotide to be mapped to a specific region within 18 S ribosomal RNA. The data on RNA oligonucleotides were correlated with Xenopus ribosomal DNA sequence data in the regions defined by the mapping experiments to identify the precise locations of most of the methyl groups in the X. laevis 18 S RNA sequence. The remaining uncertainties in Xenopus were solved with the aid of data from ribonuclease A fingerprints and, in a few instances, relevant oligonucleotide or sequence data from other laboratories. The locations of most of the methyl groups in human 18 S ribosomal RNA were deduced from the high degree of correspondence between methylated oligonucleotides from human and X. laevis 18 S RNA, together with knowledge of the human 18 S ribosomal DNA sequence. The remaining methylation sites in human 18 S RNA were located with assistance from relevant published comparative data. In the aligned sequences, human and other mammalian 18 S RNA are methylated at all the same positions as in X. laevis, and there are seven additional 2'-O-methylation sites in mammalian 18 S RNA. Further features of the methyl group distribution are briefly reviewed.  相似文献   

16.
The polycyclic aromatic hydrocarbon (PAH) benzo[ghi]perylene (BghiP) lacks a "classic" bay-region and is therefore unable to form vicinal dihydrodiol epoxides thought to be responsible for the genotoxicity of carcinogenic PAHs like benzo[a]pyrene. The bacterial mutagenicity of BghiP increases considerably after inhibition of the microsomal epoxide hydrolase (mEH) indicating arene oxides as genotoxic metabolites. Two K-region epoxides of BghiP, 3,4-epoxy-3,4-dihydro-BghiP (3,4-oxide) and 3,4,11,12-bisepoxy-3,4,11,12-tetrahydro-BghiP (3,4,11,12-bisoxide) identified in microsomal incubations of BghiP are weak bacterial mutagens in strain TA98 of Salmonella typhimurium with 5.5 and 1.5 his+-revertant colonies/nmol, respectively. After microsomal activation of BghiP in the presence of calf thymus DNA three DNA adducts were detected using 32P-postlabeling. The total DNA binding of 2.1 fmol/microg DNA, representing 7 adducts in 10(7) nucleotides, was raised 3.6-fold when mEH was inhibited indicating arene oxides as DNA binding metabolites. Co-chromatography revealed the identity between the main adduct of metabolically activated BghiP and the main adduct of the 3,4-oxide. DNA adducts of BghiP originating from the 3,4,11,12-bisoxide were not found. Therefore, a K-region epoxide is proposed to be responsible for the genotoxicity of BghiP and possibly of other PAHs without a "classic" bay-region.  相似文献   

17.
Hydration and recognition of methylated CpG steps in DNA.   总被引:2,自引:0,他引:2       下载免费PDF全文
C Mayer-Jung  D Moras    Y Timsit 《The EMBO journal》1998,17(9):2709-2718
The analysis of the hydration pattern around methylated CpG steps in three high resolution (1.7, 2.15 and 2.2 A) crystal structures of A-DNA decamers reveals that the methyl groups of cytosine residues are well hydrated. In comparing the native structure with two structurally distinct forms of the decamer d(CCGCCGGCGG) fully methylated at its CpG steps, this study shows also that in certain structural and sequence contexts, the methylated cytosine base can be more hydrated that the unmodified one. These water molecules seem to be stabilized in front of the methyl group through the formation C-H...O interactions. In addition, these structures provide the first observation of magnesium cations bound to the major groove of A-DNA and reveal two distinct modes of metal binding in methylated and native duplexes. These findings suggest that methylated cytosine bases could be recognized by protein or DNA polar residues through their tightly bound water molecules.  相似文献   

18.
Polycyclic aromatic hydrocarbons (PAHs) are ubiquitous environmental pollutants, and some are potent carcinogens in rodents. Carcinogenic PAHs are activated in the cells to metabolites that react with DNA to form covalent adducts. For most PAHs the reactive, electrophilic species which bind to DNA, are bay-region diol-epoxides. Application of 32P-postlabeling to PAH-DNA adducts analysis revealed that for some PAHs the adduct profiles generated in model systems are more complex and include adducts which are more polar than those formed by classic bay-region diol-epoxides. This minireview summaries the information gained on typical representatives of polar PAH-DNA adducts. Formation of triol-epoxide-DNA adducts was proposed for chrysene and a non-alterant PAH, benzo[b]fluoranthene (B[b]F). 5-OH-B[b]F, the precursor of B[b]F triol-epoxide, was found to be a potent tumor initiator in mouse skin. For planar PAHs such as dibenzanthracenes the possibility of bis-diol epoxide-DNA adducts formation was suggested. The most comprehensive data were obtained for dibenz[a,j]anthracene (DB[a,j]A). This hydrocarbon when applied to SENCAR mouse skin forms up to 23 species of adducts, most of which are polar. Among these polar adducts seven were identified as derived from DB[aj]A-3,4-10,11-bis-diol. Analysis of tumor-initiating activity showed, however, that this proximate metabolite was inactive in this respect. In contrast, an excellent correlation was observed between levels of less polar DNA adducts (i.e. those derived from bay-region diolepoxides) and skin tumor initiating activity of DB[a,j]A. Thus, while triol-epoxides seems to be involved in tumor initiating activity of the parent compound, non alterant B[b]F, the significance of bis-diol epoxide-DNA adducts, at least those derived from DB[aj]A, is minor.  相似文献   

19.
A cytosolic protein carboxyl methyltransferase (S-adenosyl-L-methionine:protein O-methyltransferase, E.C. 2.1.1.24) purified from human erythrocytes catalyzes the methylation of erythrocyte membrane proteins in vitro using S-adenosyl-L-[methyl-3H]methionine as the methyl group donor. The principal methyl-accepting proteins have been identified by sodium dodecyl sulfate-gel electrophoresis at pH 2.4 and fluorography as the anion transport protein (band 3), ankyrin (band 2.1), and integral membrane proteins with molecular weights of 45,000, 28,000, and 21,000. Many of the methylation sites associated with intrinsic membrane proteins may reside in their extracellular portions, since these same proteins are methylated when intact cells are used as the substrate. The maximal number of methyl groups transferred in these experiments is approximately 30 pmol/mg of membrane protein, a value which represents less than one methyl group/50 polypeptide chains of any methyl-accepting species. The number of methylation sites associated with the membranes is increased, but not to stoichiometric levels, by prior demethylation of the membranes. The additional sites are associated primarily with bands 2.1 and 4.1, the principal methyl acceptors in vivo, suggesting that most methylation sites are fully modified in vivo. Extracellular methylation sites are not increased by demethylation of membranes. The aspartic acid beta-methyl ester which can be isolated from carboxypeptidase Y digests of [3H]methylated membranes is in the unusual D-stereoconfiguration. Similar results have been obtained with [3H]methylated membranes isolated from intact cells (McFadden, P.N., and Clarke, S. (1982) Proc. Natl. Acad. Sci. U.S.A. 79, 2460-2464). It is proposed that the methyltransferase recognizes D-aspartyl residues in proteins and is involved with the metabolism of damaged proteins in vivo.  相似文献   

20.
Hepatic cholesterol-epoxide hydrolase is a microsomal enzyme which appears to be catalytically distinct from the epoxide hydrolase responsible for the catabolism of a wide variety of aromatic and aliphatic epoxides. The diastereomeric forms of cholesterol epoxide, cholesterol 5 alpha,6 alpha-, and cholesterol 5 beta,6 beta-epoxides are converted to cholestane-3 beta,5 alpha,6 beta-triol with equal facility. Kinetic analysis of cholesterol-epoxide hydrolase demonstrated that both diastereomers bind to a common catalytic site. Apparent Km values of 3.69 and 4.42 microM were derived for cholesterol 5 alpha,6 alpha- and cholesterol 5 beta,6 beta-epoxide, respectively. In addition, enzyme activity with both diastereomers was product-inhibited by cholestanetriol through a competitive mechanism with the apparent Ki for cholestanetriol being 10.8 and 6.8 microM against cholesterol alpha- and beta-epoxides, respectively. This inhibitory effect of cholestanetriol may account for the difference observed in the hydration rates for the cholesterol epoxide isomers when they are incubated together in the presence of liver microsomes. Inhibitors of epoxide hydrolase were studied, and three oxidation products were found to be particularly effective against cholesterol-epoxide hydrolase while producing no significant inhibition of styrene-epoxide hydrolase. These inhibitors were 7-ketocholesterol, 6-ketocholestanol, and 7-ketocholestanol, the latter displaying an apparent Ki lower than the Km for either cholesterol epoxide isomer. None of the xenobiotic epoxide hydrolase inhibitors or activators studied affected cholesterol-epoxide hydrolase activity.  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