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 390 毫秒
1.
Carcinogenic activity of many polycyclic aromatic hydrocarbons (PAHs) is mainly attributed to their respective diol epoxides, which can be classified as either bay or fjord region depending upon the location of the epoxide function. The Pi class human glutathione (GSH) transferase (hGSTP1-1), which is polymorphic in humans with respect to amino acid residues in positions 104 (isoleucine or valine) and/or 113 (alanine or valine), plays an important role in the detoxification of PAH-diol epoxides. Here, we report that the location of the epoxide function determines specificity of allelic variants of hGSTP1-1 toward racemic anti-diol epoxide isomers of benzo[c]chrysene (B[c]C). The catalytic efficiency (k(cat)/K(m)) of V104,A113 (VA) and V104,V113 (VV) variants of hGSTP1-1 was approximately 2.3- and 1.7-fold higher, respectively, than that of the I104,A113 (IA) isoform toward bay region isomer (+/-)-anti-B[c]C-1,2-diol-3,4-epoxide. On the other hand, the IA variant was approximately 1.6- and 3.5-fold more efficient than VA and VV isoforms, respectively, in catalyzing the GSH conjugation of fjord region isomer (+/-)-anti-B[c]C-9,10-diol-11,12-epoxide. The results of the present study clearly indicate that the location of the epoxide function determines specificity of the allelic variants of hGSTP1-1 in the GSH conjugation of activated diol epoxide isomers of B[c]C.  相似文献   

2.
Activities of epoxide hydratase and glutathione (GSH) S-transferase were investigated in subcellular fractions of Drosophila melanogaster, and these activities were compared with analogous enzymic activities in extracts from rat liver. Microsomes of Drosophila were active in the hydratation of styrene oxide catalyzed by epoxide hydratase. The post-microsomal supernatant of Drosophila catalyzed the conjugation of GSH with 1-chloro-2,4-dinitrobenzene. However, GSH S-transferase activity with styrene oxide as the electrophilic substrate was not measurable. The respective specific activities of epoxide hydratase (per mg microsomal protein) and GSH S-transferase (per mg cytosolic protein) were factors of 5- and 10-fold lower than the corresponding activities in rat liver. However, when expressed per gram body weight, activities of both epoxide hydratase and GSH S-transferase were 3 times higher for Drosophila enzymes. The apparent Km values for the two Drosophila enzymes were higher, whereas the apparent Km values were lower, than the values found for the rat-liver enzymes. Among 3 different Drosophila strains (a wild-type, a white eye-color carrying mutant strain and a DDT-resistant strain), preliminary experiments showed no differences as far as these two enzymic activities were concerned. It is concluded that the results obtained in genetic toxicology testing with Drosophila are probably relevant to effects to be expected in mammalian systems with compounds requiring metabolic processes involving the enzymes investigated here.  相似文献   

3.
A glutathione S-transferase (GST) with activity toward 1, 2-epoxy-2-methyl-3-butene (isoprene monoxide) and cis-1, 2-dichloroepoxyethane was purified from the isoprene-utilizing bacterium Rhodococcus sp. strain AD45. The homodimeric enzyme (two subunits of 27 kDa each) catalyzed the glutathione (GSH)-dependent ring opening of various epoxides. At 5 mM GSH, the enzyme followed Michaelis-Menten kinetics for isoprene monoxide and cis-1, 2-dichloroepoxyethane, with Vmax values of 66 and 2.4 micromol min-1 mg of protein-1 and Km values of 0.3 and 0.1 mM for isoprene monoxide and cis-1,2-dichloroepoxyethane, respectively. Activities increased linearly with the GSH concentration up to 25 mM. 1H nuclear magnetic resonance spectroscopy showed that the product of GSH conjugation to isoprene monoxide was 1-hydroxy-2-glutathionyl-2-methyl-3-butene (HGMB). Thus, nucleophilic attack of GSH occurred on the tertiary carbon atom of the epoxide ring. HGMB was further converted by an NAD+-dependent dehydrogenase, and this enzyme was also purified from isoprene-grown cells. The homodimeric enzyme (two subunits of 25 kDa each) showed a high activity for HGMB, whereas simple primary and secondary alcohols were not oxidized. The enzyme catalyzed the sequential oxidation of the alcohol function to the corresponding aldehyde and carboxylic acid and followed Michaelis-Menten kinetics with respect to NAD+ and HGMB. The results suggest that the initial steps in isoprene metabolism are a monooxygenase-catalyzed conversion to isoprene monoxide, a GST-catalyzed conjugation to HGMB, and a dehydrogenase-catalyzed two-step oxidation to 2-glutathionyl-2-methyl-3-butenoic acid.  相似文献   

4.
The time-dependent absorbance change that occurs when benzo[alpha]pyrene 7,8-diol-9,10-epoxide is added to solutions of calf thymus DNA has been shown, by an unequivocal chromatographic method, to correspond to DNA-catalyzed hydrolysis of the diol-epoxide. At 25 degrees C and mu = 0.10, the kinetics of the reaction of the diol-epoxide with polyadenylic acid or DNA are consistent with preequilibrium formation of a non-covalent complex between the diol-epoxide and the polynucleotide or DNA, followed by hydrolysis of the bound epoxide by a process that is first-order in hydronium ions. Cacodylic acid also catalyzes the hydrolysis of the epoxide bound to polyadenylic acid. The rate of the DNA-catalyzed hydrolysis exhibits little or no enantiomeric selectivity for the diol-epoxide. DNA catalyzed hydrolysis of the diol-epoxide is extraordinarily sensitive to the salt concentration in the reaction medium: the rate of hydrolysis of the bound epoxide at pH 7 is retarded by a factor of approximately 45 in the presence of 0.1 M sodium chloride compared to a 1 mM buffer containing no added salt. Thus, studies of the interactions of DNA with carcinogenic diol-epoxides must take into account the ionic environment of DNA within the cell.  相似文献   

5.
1. The substrate specificity of membrane-bound and purified epoxide hydrase from rat liver microsomes has been studied. Both enzyme preparations catalyzed the hydration of a variety of alkene oxidase as well as arene oxides of several polycyclic aromatic hydrocarbons. 2. Unlike the membrane-bound enzyme, the rate of hydration for most of the substrates catalyzed by the purified epoxide hydrase was constant for only 1 or 2 min. The addition of dilauroyl phosphatidylcholine or heated microsomes to the incubation mixture extended the linearity of the reaction. 3. When rat liver microsomes were used as the source of the enzyme, the apparent Km values for many of the substrates were dependent on the amount of microsomes used. When purified epoxide hydrase was used as the enzyme source and benzo(a)pyrene 11,12-oxide as substrate, the apparent Km for benzo(a)pyrene 11,12-oxide was independent of enzyme concentration but dependent on added lipid concentration. Thus, in the absence of added dilauroyl phosphatidylcholine or in the presence of this lipid at a concentration below its critical micelle concentration, the observed Km for benzo(a)pyrene 11,12-oxide remained constant. However, when the lipid concentration was greater than the critical micelle concentration, the apparent Km value increased linearly with lipid concentration. These results are consistent with a model based on the partition of lipid-soluble substrate between the lipid micelle and the aqueous medium.  相似文献   

6.
The possible role of glutathione S-transferases (GST) in detoxification of fatty acid epoxides generated during lipid peroxidation has been evaluated. Present studies showed that cytosolic human glutathione S-transferases belonging to alpha, mu, and pi classes isolated from human liver and lung catalyzed the conjugation of glutathione and 9,10-epoxystearic acid. The product of enzymatic reaction, i.e., conjugate of GSH and epoxystearic acid, was isolated and characterized. The Michaelis constant (Km) values of the alpha, mu, and pi classes of GSTs for 9,10-epoxystearic acid were found to be 0.47, 0.32 and 0.80 mM, respectively, whereas the maximal velocity (V max) values for the alpha, mu, and pi classes of GSTs were found to be 142, 256, and 52 mol/min/mol, respectively. These results indicate that even though 9,10-epoxystearic acid is a substrate for all the three classes of GSTs, the mu class isozymes have maximum activity toward this substrate and may preferentially metabolize fatty acid epoxides more effectively as compared to the other classes of GSTs.  相似文献   

7.
COS cells transiently expressing glutathione S-transferase (GST) pi, Ya, or Yb1 (human Pi, rat Alpha or Mu, cytosolic classes) were purified by flow cytometry and used in colony-forming assays to show that GST confers cellular resistance to the carcinogen benzo[a]pyrene (+/-)-anti-diol epoxide (anti-BPDE). We developed a sorting technique to viably separate recombinant GST+ cells (20%) from the nonexpressing electroporated population (80%) on the basis of a GST-catalyzed intracellular conjugation of glutathione to the fluorescent labeling reagent monochlorobimane (mClB). The concentration of mClB, length of time cells are exposed to mClB, and activity of the expressed GST isozyme determined the degree to which recombinant GST+ cells fluoresced more intensely than controls. On-line reagent addition ensured that all cells were exposed to 25 microM mClB for 30-35 s during transit before being analyzed for fluorescence intensity and sorted. The apparent Km for mClB of the endogenous COS cell GST-catalyzed intracellular reaction was 88 microM. Stained GST Ya+ or Yb1+ cells catalyzed the conjugation 2 or 5 times more effectively than GST pi+ cells. Enzyme activity in cytosolic fractions prepared from sorted recombinant GST+ cells was 1.8 +/- 0.3-fold greater than that of the control (80 +/- 4 nmol/min/mg protein). Upon a 5-fold purification of GST pi+ cells in the electroporated population, resistance to anti-BPDE in colony-forming assays increased 5 times, from 1.1-fold (unsorted) to 1.5-fold (sorted) (P less than 0.001).  相似文献   

8.
Optically active 7-hydroxy-7,8-dihydrobenzo[a]pyrene and 8-hydroxy-7,8-dihydrobenzo[a]pyrene were identified as two of the major metabolites formed by incubation of 7,8-dihydrobenzo[a]pyrene with rat liver microsomes. Optically active 9-hydroxy-9,10-dihydrobenzo[a]pyrene and 10-hydroxy-9,10-dihydrobenzo[a]pyrene were similarly identified as two of the minor metabolites of 9,10-dihydrobenzo[a]pyrene. The formation of these metabolites was abolished either by prior treatment of liver microsomes with carbon monoxide or the absence of NADPH, but was not inhibited by an epoxide hydrolase inhibitor. The results indicate that the aliphatic carbons of dihydro polycyclic aromatic hydrocarbons may undergo stereoselective hydroxylation reactions catalyzed by the cytochrome P-450 system of rat liver microsomes.  相似文献   

9.
Zhang K  Wong KP  Chow P 《Life sciences》2003,72(23):2629-2640
Chlorambucil (CMB) combines with glutathione (GSH) spontaneously in vitro to form monochloromonoglutathionyl CMB (MG-CMB). This was identified and quantified by an HPLC-UV method. Glutathione S-transferase (GST) purified from human colon adenocarcinoma cells increased the formation of the conjugate significantly. The GST-mediated conjugation, represented by the difference between total and spontaneous conjugation showed Michaelis-Menten kinetics with apparent Km and Vmax values of 0.2 mM and 75.8 nmol/min/mg for CMB and 5.2 mM and 127.0 nmol/min/mg for GSH respectively. Unexpectedly, we found in our study that both the spontaneous and the enzymatic conjugation of chlorambucil with GSH were affected markedly by a change in pH from 6.0 to 8.0. The optimum for the enzymatic conjugation was about 7.0, above which the spontaneous conjugation increased rapidly, while the enzymatic conjugation became lower. The plant polyphenols namely tannic acid, butein, quercetin, morin, 2-hydroxychalcone and 2'-hydroxychalcone at 40 microM inhibited the GST-mediated conjugation of CMB with GSH by 38 to 62%. Their action in this respect may contribute to sensitisation of tumour cells to anticancer drugs.  相似文献   

10.
Soluble epoxide hydrolase purified from soybean catalyzes trans-addition of water across the oxirane ring of cis-9,10-epoxystearic acid with inversion of configuration at the attacked carbon, yielding threo-9,10-dihydroxystearic acid. Kinetic analyses of the progress curves, obtained at low substrate concentrations (i.e. [S] much less than Km), and determination of the enantiomeric excess of the residual substrate by chiral-phase high-performance liquid chromatography at different reaction times, indicate that the epoxide hydrolase hydrates preferentially cis-9R, 10S-epoxystearic acid (V/Km ratio, approximately 20). Interestingly, this enantiomer is obtained by epoxidation of oleic acid catalyzed by peroxygenase, a hydroperoxide-dependent oxidase, we have previously described in soybean (Blée, E., and Schuber, F. (1990) J.Biol. Chem. 265, 12887-12894). For the epoxide hydrolase to show high enantioselectivity there must be a free carboxylic acid functionality on the substrate which probably influences its positioning within the active site. This selectivity, which in principle can be used for kinetic resolution of the cis-9,10-epoxystearic acid enantiomers, is much reduced with methyl cis-9,10-epoxystearate. 18O-Labeling experiments indicate that water attacks both cis-9,10-epoxystearic acid enantiomers on the oxirane carbon which has the S-chirality. Results show that soybean epoxide hydrolase produces exclusively threo-9R,10R-dihydroxystearic acid, i.e. a naturally occurring metabolite in higher plants. cis-9,10-Epoxy-18-hydroxystearic acid, a cutin monomer, was a poorer substrate of the epoxide hydrolase than 9,10-epoxystearic acid (V/Km ratio for the preferred enantiomers, approximately 19). From a physiological point of view, peroxygenase and this newly described epoxide hydrolase could be responsible, in vivo, for the biosynthesis of a class of oxygenated fatty acid compounds known to be involved in cutin monomers production and in plant defense mechanisms.  相似文献   

11.
The covalent binding of the anti-diol epoxide of benzo[a]pyrene to cellular DNA of mouse skin in organ culture is affected by the presence of ellagic acid in the culture medium. At 10(-4) M, BaPDE /DNA formation is 40% less than that observed when no ellagic acid is present. Caffeic acid, a similar plant phenolic compound, demonstrates no inhibitory effect on BaPDE /formation. The plant phenolic acids do not drastically interfere with the metabolism of benzo[a]pyrene as shown by the BaP-metabolite profiles of the skin or of the culture medium.  相似文献   

12.
Cell lysates of mouse peritoneal macrophages, in the presence of reduced glutathione, converted leukotriene LTA4 to LTC4, and neither LTD4 nor LTE4 was detected. Therefore, like cultured rat basophilic leukemia cells (RBL cells), the peritoneal macrophage contains LTC4 synthetase and appears to contain little, if any, gamma-glutamyl transpeptidase. When LTA4 was added to subcellular fractions of mouse macrophage lysate, the highest specific activity of LTC4 synthetase (nmol LTC4/mg protein per 10 min) was associated with the particulate or membrane fractions (i.e., 10(4) and 10(5) X g pellets). The 10(5) X g supernatant contains approx. 1% of the specific activity and 6% of the total LTC4 synthetase activity compared with that of the 10(5) X g pellet. Conversely, the 10(5) X g supernatant had four-times more specific activity and 19-times more total GSH S-transferase activity than did the 10(5) X g pellet when evaluated using 1-chloro-2,4-dinitrobenzene (DNCB) as the substrate. LTA4 was converted to LTC4 by the membrane enzyme LTC4 synthetase in a dose-dependent manner at low LTA4 concentrations (3-50 microM) and reached a plateau of approx. 30 microM LTA4 using the macrophage 10(5) X g pellet as an enzyme source. The apparent Km value of LTC4 synthetase for LTA4 was estimated to be 5 microM based on Lineweaver-Burk plots. Enzyme in the 10(5) X g supernatant produced negligible quantities of LTC4 (1% or less of the particulate fractions) over a wide range of LTA4 concentrations. However, an enzyme in the 10(5) X g supernatant fraction presumed to be GSH S-transferase effectively catalyzes the conjugation of glutathione (GSH) with the aromatic compound DNCB. The apparent Km value of GSH S-transferase for DNCB was estimated to be 1.0-1.5 mM. On the other hand, enzyme from the membrane fraction (i.e., 10(5) X g pellet) catalyzed this reaction at a negligible rate over a wide range of DNCB concentrations. The apparent Km value of LTC4 synthetase for GSH was estimated to be 0.36 mM and the corresponding Km value estimated for the glutathione S-transferase was 0.25-0.76 mM. These values indicate similar kinetics for GSH utilization by both enzymes. These Km values are also significantly lower than the intracellular GSH levels of 2 to 5 mM. Therefore, it is suggested that the substrate limiting LTC4 synthetase activity is LTA4 and not GSH. Our results indicate that LTC4 synthetase from mouse peritoneal macrophages is a particulate or membrane-bound enzyme, as was reported by Bach et al.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

13.
The effects of amino acids on glutathione (GSH) production by Saccharomyces cerevisiae T65 were investigated in this paper. Cysteine was the most important amino acids, which increased intracellular GSH content greatly but inhibited cell growth at the same time. The suitable amino acids addition strategy was two-step addition: in the first step, cysteine was added after two hours culture to 2 mM and then, the three amino acids (glutamic acid, glycine, and serine) were added after seven hours culture. The optimum concentration of those three key amino acids (10 mM glutamic acid, 10 mM glycine, and 10 mM serine) was obtained by orthogonal matrix method. With the optimum amino acids addition strategy a 1.63% intracellular GSH content was obtained in shake flask culture. Intracellular GSH content was 55.2% higher than the experiments without three amino acids addition. The cell biomass and GSH yield were 9.4 g/L and 153.2 mg/L, respectively. Using this amino acids addition strategy in the fed-batch culture of S. cerevisiae T65, GSH content, the biomass, and GSH yield reached 1.41%, 133 g/L, and 1875 mg/L, respectively, after 44 hours fermentation. GSH yield was about 2.67 times as that of amino acids free.  相似文献   

14.
Gu Y  Xiao B  Wargo HL  Bucher MH  Singh SV  Ji X 《Biochemistry》2003,42(4):917-921
Murine class alpha glutathione S-transferase subunit types A2 (mGSTA2-2) and A1 (mGSTA1-1) have high catalytic efficiency for 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]. Only 10 residues differ between the sequences of mGSTA1-1 and 2-2. However, the catalytic efficiency of mGSTA1-1 for GSH conjugation of (+)-anti-BPDE is >3-fold higher as compared with mGSTA2-2. The crystal structure of mGSTA1-1 in complex with the GSH conjugate of (+)-anti-7,8-dihydroxy-9,10-oxy-7,8,9,10-tetrahydrobenzo[a]pyrene (GSBpd) reveals that R216 and I221 in the last helix play important roles in catalysis [Gu, Y., Singh, S. V., and Ji, X. (2000) Biochemistry 39, 12552-12557]. The crystal structure of mGSTA2-2 in complex with GSBpd has been determined, which reveals a different binding mode of GSBpd. Comparison of the two structures suggests that residues 207 and 221 are responsible for the different binding mode of GSBpd and therefore contribute to the distinct catalytic efficiency of the two isozymes.  相似文献   

15.
Enzymic activities catalyzing allylic epoxide, leukotriene A4, to leukotriene C4 by conjugation with glutathione were present mainly in microsomal fractions of spleens and lungs of guinea pigs and rats. Leukotriene C4 (LTC4) synthase was solubilized from the microsomes of guinea-pig lung by the new procedures of a combination of 3-[3-cholamidopropyl)dimethylammonio)-1-propanesulfonate (CHAPS), digitonin and KCl. The enzyme was partially purified by two steps of column chromatography which resulted in a complete resolution of the enzyme from glutathione S-transferases (EC 2.5.1.18). The partially purified LTC4 synthase showed a Vmax value of 40 nmol/min per mg, and the apparent Km values for LTA4 and glutathione were 36 microM and 1.6 mM, respectively. The enzyme was unstable, and half of the activity was lost by incubation at 37 degrees C for 3 min. Glutathione at 10 mM completely protected the enzyme against this inactivation, while other sulfhydryl-group-reducing reagents were ineffective. The partially purified enzyme revealed a high specificity towards 5,6-epoxide leukotrienes (LTA4 and its methyl ester), while rat cytosolic glutathione S-transferases catalyzed conjugation of glutathione to various positional isomers of epoxide leukotrienes.  相似文献   

16.
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   

17.
L Recio  A W Hsie 《Mutation research》1987,178(2):257-269
Biologically reactive metabolites of benzo[a]pyrene (BP) and benzo[a]-pyrene 7,8-diol (BP-diol), formed by the mixed-function oxidase (MFO) system, are substrates for conjugation and detoxication by glutathione (GSH) when catalyzed by glutathione S-transferases (GSHT). We have investigated the detoxication of BP- and BP-diol-induced cytotoxicity and mutagenicity with GSH by supplementing the S9 mix used in the Chinese hamster ovary cells/hypoxanthine-guanine phosphoribosyltransferase (CHO/HGPRT) assay with GSH (6.5 mM) or GSH plus GSHT. The addition of GSH to the S9 mix resulted in a reduction of BP- and BP-diol induced cytotoxicity. GSH plus GSHT eliminated BP-induced cytotoxicity and reduced the mutagenicity of BP. GSH inhibited the mutagenicity at low (essentially non-lethal) concentrations of BP-diol, but did not do so at toxic concentrations. GSH plus GSHT inhibited the cytotoxicity and mutagenicity of BP-diol at concentrations not affected by GSH alone. These studies indicate that biochemical mechanisms of detoxication can affect the biological activity of a carcinogen, such as BP or BP-diol as profoundly as bioactivation by the MFO system.  相似文献   

18.
(±)-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   

19.
The conjugation of benzoyl-CoA with the aliphatic and acidic amino acids by glycine N-acyltransferase, as well as the amides of the latter group, was investigated. Bovine and human liver benzoyl-amino acid conjugation were investigated using electrospray ionization tandem mass spectrometry (ESI-MS-MS). Bovine glycine N-acyltransferase catalyzed conjugation of benzoyl-CoA with Gly (Km(Gly) = 6.2 mM), Asn (Km(Asn) = 129 mM), Gln (Km(Gln) = 353 mM), Ala (Km(Ala) = 1573 mM), Glu (Km(Glu) = 1148 mM) as well as Ser in a sequential mechanism. In the case of the human form, conjugation with Gly (Km(Gly) = 6.4 mM), Ala (Km(Ala) = 997 mM), and Glu was detected. The presence of these alternative conjugates did not inhibit bovine glycine N-acyltransferase activity significantly. Considering the relatively low levels at which these conjugates are formed, it is unlikely that they will have a significant contribution to acyl-amino acid conjugation under normal conditions in vivo. However, their cumulative contribution to acyl-amino acid conjugation under metabolic disease states may prove to have a useful contribution to detoxification of elevated acyl-CoAs.  相似文献   

20.
M L Kuo  J K Lin 《Mutation research》1989,212(2):231-239
The induction of DNA single-strand breaks in C3H10T1/2 mouse fibroblasts and Chinese hamster ovary (CHO) cells by N-nitroso-N-2-fluorenylacetamide (N-NO-2-FAA) was demonstrated by the alkaline elution technique. Without metabolic activating system (i.e., rat liver S9 fraction), N-NO-2-FAA exhibits more direct and strong damaging effects on DNA than its parent compound, 2-FAA, at equal concentration in both cell lines. To compare the DNA-damaging potency of N-NO-2-FAA with other well-known carcinogens, such as benzo[a]pyrene, 2-nitrofluorene, and N-methyl-N'-nitrosoguanidine (MNNG), the order of potency is as follows: MNNG (5 microM) greater than N-NO-2-FAA (150 microM) greater than benzo[a]pyrene (20 microM) at equitoxic concentrations, LD37, in the same cell system. Another parallel experiment indicated that N-NO-2-FAA could disrupt the superhelicity of circular plasmid DNA (pBR 322) at a dose range of 0.1-50 mM; however, a complete conversion to form III linear DNA was found at the highest concentration (50 mM). After treatment with various concentrations of N-NO-2-FAA, ouabain resistance (ouar) was induced in C3H10T1/2 cells, while both ouar and 6-thioguanine resistance (6-TGr) were induced in CHO cells. The mutation frequency in the Na+/K+-ATPase locus in CHO cells (1.5 X 10(-6) mutants/microM) is higher than that in C3H10T1/2 cells (1.0 X 10(-6) mutants/microM). The maximal mutation frequency at the Na+/K+-ATPase gene locus was attained with 30 min of exposure in C3H10T1/2 cells, whereas the mutation frequency in CHO cells continued to increase up to 80 min of treatment. Similarly, the maximal mutation frequency at the HPRT locus also continued to increase up to 80 min of treatment. Finally, a linear plot of alkali-labile lesions versus 6-TGr mutations was obtained; but the same relationship was not observed in the case of ouar mutation.  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