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.
Naphthazarin (5,8-dihydroxy-1,4-naphthoquinone), the basic unit of several tetracyclic antitumor antibiotics, and its glutathione conjugate were reduced by the one- and two-electron transfer flavoproteins NADPH-cytochrome P450 reductase and DT-diaphorase to their semi- and hydroquinone forms, respectively. Kinetic studies performed on purified DT-diaphorase showed the following results: KNADPHm = 68 microM, KQuinonem = 0.92 microM, and Vmax 1300 nmol X min-1 X microgram enzyme-1. Similar studies performed on purified NADPH-cytochrome P450 reductase indicated a lower KNADPHm (10.5 microM) and higher KQuinonem (2.3 microM). The Vmax values were 20-fold lower (46 nmol X min-1 X micrograms enzyme-1) than those observed with DT-diaphorase. DT-diaphorase reduced the naphthazarin-glutathione conjugate with an efficiency 5-fold lower than that observed with the parent quinone. The nucleophilic addition of GSH to naphthazarin proceeded with GSH consumption at rates slower than those observed with 1,4-naphthoquinone and its monohydroxy derivative, 5-hydroxy-1,4-naphthoquinone. The initial rate of GSH consumption during these reactions did not vary whether the assay was carried out under anaerobic or aerobic conditions. Autoxidation accompanied the DT-diaphorase and NADPH-cytochrome P450 reductase catalysis of naphthazarin and its glutathionyl adduct as well as the 1,4-reductive addition of GSH to naphthazarin. Superoxide dismutase at catalytic concentrations (nM range) enhanced slightly (1.1- to 1.6-fold) the autoxidation following the enzymatic catalysis of naphthazarin. Autoxidation during the GSH reductive addition to 1,4-naphthoquinones decreased with increasing number of -OH substituents, 1,4-naphthoquinone greater than 5-hydroxy-1,4-naphthoquinone greater than 5,8-dihydroxy-1,4-naphthoquinone, thus revealing that the contribution of redox transitions other than autoxidation, e.g., cross-oxidation, to the decay of the primary product of nucleophilic addition increases with increasing number of -OH substituents. Superoxide dismutase enhanced substantially the autoxidation of glutathionyl-naphthohydroquinone adducts, thereby affecting only slightly the total GSH consumed and GSSG formed during the reaction. The present results are discussed in terms of the relative contribution of one- and two-electron transfer flavoproteins to the bioreductive activation of naphthazarin and its glutathionyl conjugate as well as the importance of autoxidation reactions in the mechanism(s) of quinone cytotoxicity.  相似文献   

2.
The nucleophilic addition of GSH to quinonoid compounds, characterized as a 1,4-reductive addition of the Michael type, was studied with p-benzoquinone- and 1,4-naphthoquinone epoxides with different degree of methyl substitution. Identification and evaluation of molecular products from the above reaction were assessed by h.p.l.c. with either reductive or oxidative electrochemical detection, based on the redox properties retained in the molecular products formed. It was found that the degree of methyl substitution of the quinone epoxide, from either the 1,4-naphthoquinone- or p-benzoquinone epoxide series, determined their rate of reaction with GSH. The reductive addition implied the rearrangement of the quinone structure with opening of the epoxide ring yielding as the primary product a hydroxy-glutathionyl substituted adduct of either p-benzohydroquinone or 1,4-naphthohydroquinone. The primary product undergoes elimination reactions and redox transitions which bring about a number of secondary molecular products. The distribution pattern of the latter depends on the degree of methyl substitution of the quinone epoxide studied and on the concentration of O2 in the solution. The occurrence of the hydroxy-substituent in position alpha, adjacent to the carbonyl group, enhances the autoxidation properties of the compound resulting in an augmented O2 consumption and H2O2 production. Therefore, it could be expected that the chemical reactivity of the products originating from the thiol-mediated nucleophilic addition to quinone epoxides would be of toxicological interest.  相似文献   

3.
The reaction between H2O2 and p-benzoquinone proceeds with consumption of both reactants with second order rate constants of 1.66- and 0.77 M-1S-1, respectively. The process is mainly supported by oxygen addition reactions to the quinone resulting in the formation of both 2,3-epoxy-p-benzoquinone and 2-OH-p-benzoquinone. The former product accumulates in the assay mixture without participating in further reactions. The formation of the latter product implies free radical intermediates such as 2-OH-p-benzosemiquinone anion, which supports the generation of electronically excited states upon its oxidation by H2O2, presumably as part of an organic Fenton reaction. The relaxation of the excited state is accompanied by photoemission at 485-530 nm. Glutathione was found to counteract the oxidative aspects of the reaction between p-benzoquinone and H2O2 by a series of processes involving (a) a rapid reductive addition to the quinone with formation of a substituted p-benzohydroquinone; (b) an effective quenching of photoemission, which might be attributed to the deactivation of the excited state by the p-benzohydroquinone-glutathione adduct, and (c) the decomposition of the formed 2,3-epoxy-p-benzoquinone, also by reductive cleavage of the epoxide ring.  相似文献   

4.
The kinetic parameters of the redox transitions subsequent to the two-electron transfer implied in the glutathione (GSH) reductive addition to 2- and 6-hydroxymethyl-1,4-naphthoquinone bioalkylating agents were examined in terms of autoxidation, GSH consumption in the arylation reaction, oxidation of the thiol to glutathione disulfide (GSSG), and free radical formation detected by the spin-trapping electron spin resonance method. The position of the hydroxymethyl substituent in either the benzenoid or the quinonoid ring differentially influenced the initial rates of hydroquinone autoxidation as well as thiol oxidation. Thus, GSSG- and hydrogen peroxide formation during the GSH reductive addition to 6-hydroxymethyl-1,4-naphthoquinone proceeded at rates substantially higher than those observed with the 2-hydroxymethyl derivative. The distribution and concentration of molecular end products, however, was the same for both quinones, regardless of the position of the hydroxymethyl substituent. The [O2]consumed/[GSSG]formed ratio was above unity in both cases, thus indicating the occurrence of autoxidation reactions other than those involved during GSSG formation. EPR studies using the spin probe 5,5'-dimethyl-1-pyrroline-N-oxide (DMPO) suggested that the oxidation of GSH coupled to the above redox transitions involved the formation of radicals of differing structure, such as hydroxyl and thiyl radicals. These were identified as the corresponding DMPO adducts. The detection of either DMPO adduct depended on the concentration of GSH in the reaction mixture: the hydroxyl radical adduct of DMPO prevailed at low GSH concentrations, whereas the thiyl radical adduct of DMPO prevailed at high GSH concentrations. The production of the former adduct was sensitive to catalase, whereas that of the latter was sensitive to superoxide dismutase as well as to catalase. The relevance of free radical formation coupled to thiol oxidation is discussed in terms of the thermodynamic and kinetic properties of the reactions involved as well as in terms of potential implications in quinone cytotoxicity.  相似文献   

5.
The reaction between p-benzoquinone and H2O2 in slightly alkaline solutions yields three major quinoid products that accumulate in the reaction mixture: (a) 2,3-epoxy-p-benzoquinone, (b) 2-hydroxy-p-benzoquinone and (c) p-benzohydroquinone. The reaction is accompanied by photoemission, probably originating from excited triplet 2-hydroxy-p-benzoquinone. These products originate from hydrogen peroxide and hydroxide nucleophilic addition to the C2?C3 double bond, as well as secondary redox interactions. The hydroxy substituent and the epoxide ring exert a substantial influence on the electronic distribution in the p-benzoquinone molecule leading to a decrease in the half-wave potential, as compared to the parent p-benzoquinone. The generation of electronically excited states is the result of reactions secondary to the nucleophilic additions involving 2-hydroxy-p-benzosemiquinone, H2O2 and hydroxyl radical. The process involves the primary oxidation of 2-hydroxy-p-benzosemiquinone by hydrogen peroxide, followed by oxidation of the semiquinone by hydroxyl radical leading to the formation of the electronically excited quinone. The decay of the excited triplet to the ground state is accompanied by photoemission with maximal intensity at 485–530 nm. Thermodynamic calculations along with an observed increase of photoemission intensity in anaerobiosis point to the triplet (n, π*) multiplicity of the excited state. The efficiency of chemiluminescence could be calculated as 10?8 photons/2-hydroxy-p-benzoquinone molecule formed. Photoemission arising from the p-benzoquinone/H2O2 reaction was inhibited efficiently by addition of GSH to the reaction mixture. This may be due to deactivation of the triplet quinone by a 2-glutathionyl-p-benzohydroquinone adduct, involving thioether α-hydrogen atom-transfer to the triplet ketone.  相似文献   

6.
The effect of superoxide dismutase on the autoxidation of hydro- and semi-1,4-naphthoquinones with different substitution pattern and covering a one-electron reduction potential range from -95 to -415 mV was examined. The naphthoquinone derivatives were reduced via one or two electrons by purified NADPH-cytochrome P-450 reductase or DT-diaphorase, respectively. Superoxide dismutase did not alter or slightly enhance the initial rates of enzymic reduction, whereas it affected in a different manner the following autoxidation of the semi- and hydroquinones formed. Autoxidation was assessed as NADPH oxidation in excess to the amounts required to reduce the quinone present, H2O2 formation, and the redox state of the quinones. Superoxide dismutase enhanced 2--8-fold the autoxidation of 1,4-naphthosemiquinones, following the reduction of the oxidized counterpart by NADPH-cytochrome P-450 reductase, except for the glutathionyl-substituted naphthosemiquinones, whose autoxidation was not affected by superoxide dismutase. Superoxide dismutase exerted two distinct effects on the autoxidation of naphthohydroquinones formed during DT-diaphorase catalysis: on the one hand, it enhanced slightly the autoxidation of 1,4-naphthohydroquinones with a hydroxyl substituent in the benzene ring: 5-hydroxy-1,4-naphthoquinone and the corresponding derivatives with methyl- and/or glutathionyl substituents at C2 and C3, respectively. On the other hand, superoxide dismutase inhibited the autoxidation of naphthohydroquinones that were either unsubstituted or with glutathionyl-, methyl-, methoxyl-, hydroxyl substituents (the latter in the quinoid ring). The inhibition of hydroquinone autoxidation was reflected as a decrease of NADPH oxidation, suppression of H2O2 production, and accumulation of the reduced form of the quinone. The enhancement of autoxidation of 1,4-naphthosemiquinones by superoxide dismutase has been previously rationalized in terms of the rapid removal of O2-. by the enzyme from the equilibrium of the autoxidation reaction (Q2-. + O2----Q + O2-.), thus displacing it towards the right. The superoxide dismutase-dependent inhibition of H2O2 formation as well as NADPH oxidation during the autoxidation of naphthohydroquinones--except those with a hydroxyl substituent in the benzene ring--seems to apply to those organic substrates which can break down with simultaneous formation of a semiquinone and O2-.. Inhibition of hydroquinone autoxidation by superoxide dismutase can be interpreted in terms of suppression by the enzyme of O2-.- dependent chain reactions or a direct catalytic interaction with the enzyme that might involve reduction of the semiquinone at expense of O2(-.).(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
M Sono 《Biochemistry》1986,25(20):6089-6097
The dioxygen adduct of the heme protein indoleamine 2,3-dioxygenase has been generated at -30 degrees C in mixed solvents, and spectroscopic and equilibrium studies of its L-tryptophan (substrate) binding properties have been carried out for the first time. Comparative studies have also been performed with the NO and CO adducts of the ferrous enzyme. Under the conditions employed (-30 degrees C), both autoxidation and turnover (L-tryptophan + O2----formylkynurenine) of the ternary complex are effectively suppressed. Structural identification of the ternary complex is based on the 1:1 molar stoichiometry for the substrate-oxygenated enzyme adduct formation (Kd approximately 10(-4) M), the time-dependent linear product formation (turnover) at -20 degrees C, and the quantitative conversion of the complex to the ferrous CO derivative by bubbling with CO. Binding of L-tryptophan to the oxygenated enzyme leads to decreases in the intensities of its major absorption bands (lambda max 415, 541, 576 nm) and to a blue shift of its Soret peak. Interestingly, among the ferrous enzyme derivatives examined, only the substrate-bound oxygenated enzyme exhibits solvent-dependent Soret absorption peak positions, e.g., lambda max 411.5 and 413.5 nm in 65% (v/v) aqueous glycerol and ethylene glycol, respectively. In addition, indole binds to the oxygenated enzyme, causing a red shift of its Soret peak in these solvents only in the presence of substrate (411.5----414 nm and 413.5----414.5 nm, respectively), while similar effects of indole are independent of tryptophan for the other ferrous enzyme derivatives.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
4-Hydroxy-2,3-trans-nonenal, a lipid peroxidation product, inhibits glutathione peroxidase in a concentration-dependent manner. The concentration providing 50% inhibition is 0.12 mM. This inhibition can be almost completely (89%) prevented by 1 mM glutathione added to the incubation mixture 30 min before 4-hydroxy-2,3-trans-nonenal or 2,3-trans-nonenal, but not by other thiol-containing antioxidants such as 0.5 mM dithiothreitol or beta-mercaptoethanol. Again the addition of 1 mM glutathione, and not of 0.5 mM dithiothreitol or beta-mercaptoethanol, to the enzyme 30 min after incubation with 4-hydroxy-2,3-trans-nonenal restores activity to the same extent as does the preincubation with GSH. In view of the known reactivity of 4-hydroxy-2,3-trans-nonenal with lysine residues and the reversibility of the inhibition, the involvement of a lysine residue in GSH binding to glutathione peroxidase is proposed. The potential relevance of the inhibition of glutathione peroxidase by 4-hydroxy-nonenal to oxidative tissue damage is discussed with particular emphasis on neurological disorders.  相似文献   

9.
The oxidation of GSH coupled to the redox transitions of 1, Cnaphthoquinone derivatives during DT-diaphorase catalysis was examined. The quinones studied included 1,4-naphthoquinone and its dimethoxy-and hydroxy derivatives and were selected according to their different ability to undergo nucleophilic addition with GSH and the dual effect of superoxide dismutase on hydroquinone autoxidation

GSH was oxidized to GSSG during the redox transitions of the above quinones, regardless of their substitution pattern. This effect was accompanied by an increase of total O2 consumption, indicating the ability of GSH to support quinone redox cycling. The values for the relationship [O2]consumed[GSSG]formde were, with every quinone examined, above unity. thus pointing to the occurrence of autoxidation reactions other than those involved during GSSG formation

These results are discussed in terms of the functional group chemistry of the quinones and the ther-modynamic properties of the reactions involved in the reduction of the semi- to the hydro-quinone by GSH  相似文献   

10.
Trapping of 3-methylcholanthrene (MC) radical cation by nucleophilic compounds occurs specifically at the 1-carbon atom. With the purpose of providing more evidence for the hypothesis that the critical mechanism of activation of MC is one-electron oxidation, the carcinogenicity of MC was compared to that of 1-hydroxy-3-methylcholanthrene (MC-1-OH), 3-methylcholanthrene-1-one (MC-1-one), 2-hydroxy-3-methylcholanthrene (MC-2-OH), 3-methylcholanthrene-2-one (MC-2-one) and 3-methylcholanthrylene (MCL) by repeated application on mouse skin. Seven-week-old female Swiss mice in 6 groups of 30 were treated on the back with 0.2 mumol of compound in acetone twice weekly for 20 weeks. In addition, the metabolism of MC and its derivatives was studied using mouse skin homogenates. The compounds tested were classified according to carcinogenicity in 4 groups: MC and MC-2-OH, the strongest carcinogens; MC-2-one and MCL, weaker than MC and MC-2-OH; MC-1-OH, the weakest carcinogen; and MC-1-one, non-carcinogenic. These results support the hypothesis that one-electron oxidation for MC, MC-2-OH and MC-1-one might be the critical mechanism of carcinogenic activation, with C-1 the binding site to cellular nucleophiles. The carcinogenic effect of MC-1-OH is speculated to be the formation of an ester bearing a good leaving group, which might be the ultimate alkylating compound in the in vivo reaction. The lack of carcinogenic activity for MC-1-one may be attributed to absence of nucleophilic trapping at C-1 via the radical cation pathway as well as the inability of mouse skin to reduce MC-1-one to the carcinogenic MC-1-OH.  相似文献   

11.
The rates of autoxidation of a number of pure naphthohydroquinones have been determined, and the effects of pH, superoxide dismutase (SOD) and of the parent naphthoquinone on the oxidation rates have been investigated. Most compounds were slowly oxidised in acid solution with the rates increasing with increasing pH, although 2-hydroxy-, 2-hydroxy-3-methyl- and 2-amino-1,4-naphthohydroquinone were rapidly oxidised at pH 5 and the rates of oxidation of these substances were comparatively unresponsive to changes in pH. At pH 7.4, autoxidation rates decreased in the order 2,3-dichloro-1,4-naphthohydroquinone > 5-hydroxy > 2-bromo > 2-hydroxy-3-methyl > 2-amino > 2-hydroxy > 2-methoxy > 2,3-dimethoxy > 2,3-dimethyl > 2-methyl > unsubstituted hydroquinone. The autoxidation rates of the alkyl, alkoxy, hydroxy and amino derivatives were decreased in the presence of SOD, but this enzyme had no effect on the rate of autoxidation of the 2,3-dichloro and 2-bromo derivatives while that of the 5-hydroxy derivative was increased. The rates of autoxidation of all compounds except the halogen derivatives and 5-hydroxy-1,4-naphthohydroquinone were increased by addition of the parent naphthoquinone, and quinone addition partially or completely overcame the inhibitory effect of SOD. There is evidence that the reduction of quinones to hydroquinones in vivo may lead either to detoxification or to activation. This may be due to differences in the rate or mechanism of autoxidation of the hydroquinones that are formed, and the data gained in this study will provide a framework for testing this possibility.  相似文献   

12.
The effect of glutathione (GSH) on the ultraviolet (UV) induction of lambda prophage was investigated in lysogenic Escherichia coli. The data showed that extracellular GSH could inhibit the UV induction of lambda prophage. The inhibitory rates were concentration dependent, and the maximal rate obtained was 94% with 3.0 M GSH. The effect was also measured in three different lambda lysogens: a wild-type strain (wt), an isogenic GSH-deficient strain, and an isogenic strain producing increased amounts of GSH. The result showed that when subjected to UV irradiation (254 nm, 60 J m−2), GSH-deficient strain was approximately fivefold more sensitive to be lysed than wt, whereas the strain with higher intracellular GSH levels was only 28% susceptible to be lysed. With electron spin resonance and spin trapping techniques, we observed that free radical signals occurred in the suspensions of UV irradiated lysogenic cells and the intensity of signals was influenced by GSH levels. These results indicate that GSH can significantly inhibit the UV induction of lambda prophage, and that this effect is correlated to its capacity to scavenge free radicals generated after UV irradiation.  相似文献   

13.
Several 6(7)-alkyl-1,4-naphthoquinones (NQ) have been prepared by cycloaddition reactions between the monoterpene alpha-myrcene and p-benzoquinones and halogen and nitrogen-containing functional groups have been introduced at the C-2 position of the naphthoquinone ring via nucleophilic addition or substitution reactions. These substituents at positions 2/3 of the NQ clearly influence the cytotoxic potency of this type of compound. Of particular interest is substitution by arylamino, specifically p-oxyarylamino, groups, which considerably enhance their bioactivity and selectivity.  相似文献   

14.
An active site His107 residue distinguishes human glutathione S-transferase hGSTM1-1 from other mammalian Mu-class GSTs. The crystal structure of hGSTM1a-1a with bound glutathione (GSH) was solved to 1.9 A resolution, and site-directed mutagenesis supports the conclusion that a proton transfer occurs in which bound water at the catalytic site acts as a primary proton acceptor from the GSH thiol group to transfer the proton to His107. The structure of the second substrate-binding site (H-site) was determined from hGSTM1a-1a complexed with 1-glutathionyl-2,4-dinitrobenzene (GS-DNB) formed by a reaction in the crystal between GSH and 1-chloro-2,4-dinitrobenzene (CDNB). In that structure, the GSH-binding site (G-site) is occupied by the GSH moiety of the product in the same configuration as that of the enzyme-GSH complex, and the dinitrobenzene ring is anchored between the side chains of Tyr6, Leu12, His107, Met108, and Tyr115. This orientation suggested a distinct transition state that was substantiated from the structure of hGSTM1a-1a complexed with transition state analogue 1-S-(glutathionyl)-2,4,6-trinitrocyclohexadienate (Meisenheimer complex). Kinetic data for GSTM1a-1a indicate that kcat(CDNB) for the reaction is more than 3 times greater than kcat(FDNB), even though the nonenzymatic second-order rate constant is more than 50-fold greater for 1-fluoro-2,4-dinitrobenzene (FDNB), and the product is the same for both substrates. In addition, Km(FDNB) is about 20 times less than Km(CDNB). The results are consistent with a mechanism in which the formation of the transition state is rate-limiting in the nucleophilic aromatic substitution reactions. Data obtained with active-site mutants support transition states in which Tyr115, Tyr6, and His107 side chains are involved in the stabilization of the Meisenheimer complex via interactions with the ortho nitro group of CDNB or FDNB and provide insight into the means by which GSTs adapt to accommodate different substrates.  相似文献   

15.
2,3-Dihydroxybenzohydroxamoyl adenylate (I) was prepared as a potential product analog inhibitor of EntE (EC# 2.7.7.58), a 2,3-dihydroxybenzoate AMP ligase from Escherichia coli that is required for the biosynthesis of enterobactin. This compound, obtained by the aqueous reaction of imidazole-activated adenosine 5'-phosphate and 2,3-dihydroxybenzohydroxamic acid, is a competitive inhibitor with a Ki value of 4.5 x 10(-9)M. Deletion of the catecholic 3-OH group of (I), in compound (II), reduced inhibitory activity by a factor of 3.5, whereas, removal of both the 3-OH and 2-OH groups, in (III), reduced inhibitory activity by a factor of approximately 2000. Acetohydroxamoyl adenylate (IV), in which the entire catechol moiety of (I) is replaced by a hydrogen atom, gave 相似文献   

16.
Ye M  English AM 《Biochemistry》2006,45(42):12723-12732
In addition to its superoxide dismutase (SOD) activity, Cu,Zn-superoxide dismutase (CuZnSOD) catalyzes the reductive decomposition of S-nitroso-L-glutathione (GSNO) in the presence of thiols such as L-glutathione (GSH). The GSNO-reductase activity but not the superoxide dismutase (SOD) activity of CuZnSOD is inhibited by the commonly used polyaminocarboxylate metal ion chelators, EDTA and DTPA. The basis for this selective inhibition is systematically investigated here. Incubation with EDTA or DTPA caused a time-dependent decrease in the 680 nm d-d absorption of Cu(II)ZnSOD but no loss in SOD activity or in the level of metal loading of the enzyme as determined by ICP-MS. The chelators also protected the SOD activity against inhibition by the arginine-specific reagent, phenylglyoxal. Measurements of both the time course of SNO absorption decay at 333 nm and oxymyoglobin scavenging of the NO that is released confirmed that the chelators inhibit CuZnSOD catalysis of GSNO reductive decomposition by GSH. The decreased GSNO-reductase activity is correlated with decreased rates of Cu(II)ZnSOD reduction by GSH in the presence of the chelators as monitored spectrophotometrically at 680 nm. The aggregate data suggest binding of the chelators to CuZnSOD, which was detected by isothermal titration calorimetry (ITC). Dissociation constants of 0.08 +/- 0.02 and 8.3 +/- 0.2 microM were calculated from the ITC thermograms for the binding of a single EDTA and DTPA, respectively, to the CuZnSOD homodimer. No association was detected under the same conditions with the metal-free enzyme (EESOD). Thus, EDTA and DTPA must bind to the solvent-exposed active-site copper of one subunit without removing the metal. This induces a conformational change at the second active site that inhibits the GSNO-reductase but not the SOD activity of the enzyme.  相似文献   

17.
The extent of substrate enantioselectivity and regioselectivity of a series of aliphatic 2,3-dialkyl- and trialkylsubstituted oxiranes in their in vitro epoxide-hydrolase-catalyzed hydrolysis depends on the size of the alkyl residues and on the substitution pattern of the oxirane ring. The enzyme-catalyzed hydrolysis of cis-oxiranes, containing at least one methyl substituent, shows complete or nearly complete substrate enantioselectivity and regioselectivity with nucleophilic attack by water occurring with inversion of configuration at the methylsubstituted ring carbon atom of (S)-configuration. In the hydrolysis of the isomeric trans-oxiranes, both enantiomers are metabolized with a higher rate for the (2S;3S)-enantiomer. The conversion of trimethyloxirane occurs with high substrate enantioselectivity in favor of the (S)-enantiomer and with complete regioselectivity at the monomethylsubstituted ring carbon atom. The differentiation of the enantiotopic ring carbon atoms (product enantioselectivity) in the smallest aliphatic meso-oxirane, cis-2,3-dimethyloxirane, leads to (2R;3R)-butane-2,3-diol with ee = 86%. cis-2-Ethyl-3-propyloxirane, possessing alkyl residues larger than methyl, represents an extremely poor substrate in the epoxide-hydrolase-catalyzed hydrolysis process.  相似文献   

18.
Fecl2, in Na phosphate buffer autoxidizes forming active oxygen species which damage deoxyribose. Di-and triphosphate adenine-nucleotides inhibit both Fe2+ autoxidation and deoxyribose damage in Na phosphate buffer pH 7.4. The inhibition is related to the number of charges of the adenine-nucleotide molecule: ATP at pH 7.4 is a better inhibitor than ADP; at a pH (6.5) close to the pK's of the third and fourth charge of ADP and ATP, ADP inhibition is greatly decreased whereas ATP inhibition is slightly affected. The extent of ATP inhibition of Fe2+ autoxidation depends both on ATP/Mg2+ and ATP/Fe2+ ratios in the reaction mixture. Formation of a Fe2+ -nucleotide complex appears to be the mechanism through which ATP and ADP inhibit autoxidation and thus the generation of active oxygen species. These findings are discussed in relation to physiological and pathological fluctuations of nucleotide concentrations.  相似文献   

19.
A primary pathway for metabolism of electrophilic compounds in Schistosoma japonicum involves glutathione S-transferase (SjGST)-catalyzed formation of glutathione (GSH) conjugates. As part of a program aimed at gaining a better understanding of the defense system of parasites, a series of aromatic halides (1-8), aliphatic halides (9, 10), epoxides (11-20), alpha,beta-unsaturated esters (21, 22), and alpha,beta-unsaturated amides (23, 24) were prepared, and their participation in glutathione conjugate formation was evaluated. Products from enzymatic and nonenzymatic reactions of these substances with glutathione were characterized and quantified by using reverse-phase high-performance liquid chromatography (HPLC), NMR, and fast atom bombardment mass spectrometry (FAB-MS) analysis. Mechanisms for formation of specific mono(glutathionyl) or bis(glutathionyl) conjugates are proposed. Although the results of this effort indicate that SjGST does not catalyze addition or substitution reactions of 1, 3, 4, 7-9, 11-13, 15-17, 19-21, and 24, they demonstrate that 2, 5, 6, 14, 18, and 23 undergo efficient enzyme-catalyzed conjugation reactions. The kcat values for SjGST with 23 and 18 are about 886-fold and 14-fold, respectively, larger than that for 5. This observation suggests that 23 is a good substrate in comparison to other electrophiles. Furthermore, the initially formed conjugation product, 23a, is also a substrate for SjGST in a process that forms the bis(glutathionyl) conjugate 23b. Products arising by enzymatic and nonenzymatic pathways are generated under the conditions of SjGST-activated GSH conjugation. Interestingly, production of nonenzymatic GSH conjugates with electrophilic substrates often overwhelms the activity of the enzyme. The nonenzymatic GSH conjugates, 9a-11a, 16a, 21a, and 22a, are inhibitors of SjGST with respective IC50 values of 1.95, 75.5, 0.96, 19.0, 152, and 0.36 microM, and they display moderate inhibitory activities against human GSTA2. Direct evidence has been gained for substrate inhibition by 10 toward SjGST and GSTA2 that is more potent than that of its GSH conjugate 10a. The significance of this work is found in the development of a convenient NMR-based technique that can be used to characterize glutathione conjugates derived from small molecule libraries as part of efforts aimed at uncovering specific potent SjGST and GSTA2 inhibitors. This method has potential in applications to the identification of novel inhibitors of other GST targets that are of chemotherapeutic interest.  相似文献   

20.
The biochemical effects of aryl substituents on the reductive dechlorination of 3-chlorobenzoate analogs were quantified with (i) a stable 3-chlorobenzoate-grown methanogenic sludge enrichment, (ii) Desulfomonile tiedjei DCB-1, isolated from this enrichment and able to catalyze the reductive dechlorination of 3-chlorobenzoate, and (iii) a defined 3-chlorobenzoate-degrading methanogenic consortium with D. tiedjei as the key dechlorinating organism. The addition of hydrogen stimulated the dechlorination rate in the consortium. The extent of this stimulation depended on the substituent. The data were evaluated with various sets of substituent constants compiled for the Hammett equation. None of the sets yielded a satisfactory correlation between experimental values and theoretical constants. This suggests that the microbially catalyzed reductive dechlorination of 3-chlorobenzoate cannot be described simply as either a nucleophilic or an electrophilic substitution reaction. Nevertheless, observations that the presence of a para-amino or -hydroxy group inhibited the rate of dechlorination suggest that the rate-limiting step in the reductive dechlorination of 3-chlorobenzoate is a nucleophilic attack on the negatively charged π electron cloud around the benzene nucleus.  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