首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
R A Bednar 《Biochemistry》1990,29(15):3684-3690
The reactivity of simple alkyl thiolates with N-ethylmaleimide (NEM) follows the Br?nsted equation, log kS- = log G + beta pK, with G = 790 M-1 min-1 and beta = 0.43. The rate constant for the reaction of the thiolate of 2-mercaptoethanol with NEM is 10(7) M-1 min-1, whereas the rate constant for the reaction of the protonated thiol is less than 0.0002 M-1 min-1. The intrinsic reactivity of the protonated thiol (SH) is over (5 X 10(10]-fold less than the thiolate (S-) and makes a negligible contribution to the reactivity of thiols toward NEM. The rate of NEM modification of chalcone isomerase was conveniently measured by following the concomitant loss in enzymatic activity. The pseudo-first-order rate constants for inactivation show a linear dependence on the concentration of NEM up to 200 mM and yield no evidence for noncovalent binding of NEM to the enzyme. Evidence is presented demonstrating that the modification of chalcone isomerase by NEM is limited to a single cysteine residue over a wide range of pH. Kinetic protection against inactivation and modification by NEM is provided by competitive inhibitors and supports the assignment of this cysteine residue to be at or near the active site of chalcone isomerase. The pH dependence of inactivation of the enzyme by NEM indicates a pK of 9.2 for the cysteine residue in chalcone isomerase. At high pH, the enzymatic thiolate is only (3 X 10(-5))-fold as reactive as a low molecular weight alkyl thiolate of the same pK, suggesting a large steric inhibition of reaction on the enzyme.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The pKa of ~8.0 for the catalytic cysteine residue of thymidylate synthetase was determined from the pH dependence of inactivation by the sulfhydryl reagents methyl methanethiolsulfonate and 5,5′ dithiobis (2-nitrobenzoic acid). At low pH (5.8–6.8) a rate of reaction significantly greater than can be accounted for by the concentration of thiolate anion was observed. The observed pKa and reactivity for thymidylate synthetase are comparable to those reported for papain (Little, G. L. and Brocklehurst, K. (1972) Biochem. J. 128, 475–477) (1). On the basis of these observations we propose that the cysteines of thymidylate synthetase involved in covalent catalysis may be activated through interaction with a general base in the active site of the enzyme.  相似文献   

3.
The nucleophilic substitution reaction between glutathione and 1-chloro-2,4-dinitrobenzene has been studied at temperatures between 4 and 42°C and pH values between 6.99 and 10.80. The apparent enthalpy, entropy and free energy of ionization of the thiol group have been estimated as have the apparent enthalpy, entropy and free energy of activation of the reaction between the glutathione thiolate anion and the aromatic electrophile. The results obtained permit the calculation of values of the second order rate constant governing the reaction at a range of temperatures and pHs. These values are in accord with those reported in the literature from experimental work by others. The major glutathione S-transferase from Galleria mellonella has been studied with respect to its kinetic responses to changes in pH and temperature. There appear to be two kinetically critical ionizations governing the reaction at high pH. These ionization events are characterized by apparent pKa values of 8.61 ± 0.15 and 9.16 ± 0.22. A thermodynamic model of the kinetic behavior of the enzyme permits the prediction of its activity over a range of pH and temperature values. The apparent free energy of activation for the enzyme catalyzed reaction is only 7% lower than that for the non-catalyzed reaction between 1-chloro-2,4-dinitrobenzene and glutathione thiolate anion. This observation is compatible with the suggestion that promotion of the ionization of the glutathione thiol group is the major mechanism of catalysis.  相似文献   

4.
Sárkány Z  Skern T  Polgár L 《FEBS letters》2000,481(3):289-292
Picornains 2A are cysteine proteases of picornaviruses, a virus family containing several human and animal pathogens. The pH dependencies of the alkylations of picornain 2A of rhinovirus type 2 with iodoacetamide and iodoacetate show two reactive thiol forms, namely the free thiolate ion at high pH and an imidazole assisted thiol group at low pH. Kinetic deuterium isotope effects do not support general base catalysis by the imidazole group, but rather the existence of a catalytically competent thiolate-imidazolium ion-pair. The nature of the ion-pair differs from that of papain, the paradigm of cysteine proteases. The ion-pair is confined to the same, unusually narrow pH range in which the enzyme exhibits catalytic activity.  相似文献   

5.
An important aspect of the catalytic mechanism of microsomal glutathione transferase (MGST1) is the activation of the thiol of bound glutathione (GSH). GSH binding to MGST1 as measured by thiolate anion formation, proton release, and Meisenheimer complex formation is a slow process that can be described by a rapid binding step (K(GSH)d = 47 +/- 7 mM) of the peptide followed by slow deprotonation (k2 = 0.42 +/- 0.03 s(-1). Release of the GSH thiolate anion is very slow (apparent first-order rate k(-2) = 0.0006 +/- 0.00002 s(-)(1)) and thus explains the overall tight binding of GSH. It has been known for some time that the turnover (kcat) of MGST1 does not correlate well with the chemical reactivity of the electrophilic substrate. The steady-state kinetic parameters determined for GSH and 1-chloro-2,4-dinitrobenzene (CDNB) are consistent with thiolate anion formation (k2) being largely rate-determining in enzyme turnover (kcat = 0.26 +/- 0.07 s(-1). Thus, the chemical step of thiolate addition is not rate-limiting and can be studied as a burst of product formation on reaction of halo-nitroarene electrophiles with the E.GS- complex. The saturation behavior of the concentration dependence of the product burst with CDNB indicates that the reaction occurs in a two-step process that is characterized by rapid equilibrium binding ( = 0.53 +/- 0.08 mM) to the E.GS- complex and a relatively fast chemical reaction with the thiolate (k3 = 500 +/- 40 s(-1). In a series of substrate analogues, it is observed that log k3 is linearly related (rho value 3.5 +/- 0.3) to second substrate reactivity as described by Hammett sigma- values demonstrating a strong dependence on chemical reactivity that is similar to the nonenzymatic reaction (rho = 3.4). Microsomal glutathione transferase 1 displays the unusual property of being activated by sulfhydryl reagents. When the enzyme is activated by N-ethylmaleimide, the rate of thiolate anion formation is greatly enhanced, demonstrating for the first time the specific step that is activated. This result explains earlier observations that the enzyme is activated only with more reactive substrates. Taken together, the observations show that the kinetic mechanism of MGST1 can be described by slow GSH binding/thiolate formation followed by a chemical step that depends on the reactivity of the electrophilic substrate. As the chemical reactivity of the electrophile becomes lower the rate-determining step shifts from thiolate formation to the chemical reaction.  相似文献   

6.
1. DL-alpha-Bromo-beta(5-imidazolyl)-propionic acid is a potential affinity labelling reagent for metallo-enzymes. It has been used with the alcohol dehydrogenases from liver and yeast. The liver enzyme is chemically modified and inactivated in a Michaelis-Menten-type reaction, where one molecule of the reagent is bound per subunit. The enzyme is protected from the inhibitor in a competitive manner by imidazole, 2,2'-dipyridyl, 1,10-phenanthroline and cyclohexanone, which all combine with the active-site zinc. The protection by chloride, acetate and NADH, which are considered to bind at the general anion binding site, is not strictly competitive. Inactivation has an optimum at pH 8.5. For the liver enzyme, the reagent was found to decrease the initial rate of ethanol oxidation. Prior to the irreversible alkylation of Cys-46, reversible binding is shown to occur at the active-site zinc atom. The yeast enzyme was extremely resistant to the reagent and no specific modification was found. 2. The potential affinity labelling and crosslinking reagent, symmetrical 1,3-dibromoacetone although unstable, has also been used for chemical modification. With the liver enzyme, concentrations below 5 mM gave a reaction of the Michaelis-Menten-type at pH 7.0. Several ligands known to complex with the active-site region protect the enzyme against the reagent. Dibromoacetone gave rapid inactivation of the yeast enzyme. Despite the fact that a pseudo-first-order reaction was observed with respect to enzyme as well as inhibitor, no saturating effect was found. In this work, dibromoacetone reacted like a monofunctional reagent.  相似文献   

7.
Kinetic data for the inactivation of horse liver alcohol dehydrogenase with S-2-chloro-3-(imidazol-5-yl)propionate at pH8.2 were correlated with the three-dimensional structure of the enzyme. The R-2-chloro-3-(imidazol-5-yl)propionate enantiomer did not inactivate the enzyme, and the reaction is thus enantioselective. Inactivation follows an affinity-labelling mechanism where a reversible complex is formed before the irreversible alkylation and inactivation of the enzyme. A reversible complex is also formed with the non-inactivating enantiomer, and this shows that the selectivity occurs at the irreversible step. By using a computer-controlled display system, models of the two enantiomers of 2-chloro- and 2-bromo-3-(imidazol-5-yl)propionate were built into a model of the enzyme so that the imidazole moiety was liganded to the active-site metal, while the carboxylate group interacted with the general anion-binding site. The conformation of the imidazole derivatives and their orientation in the active site were adjusted to minimize unfavourable steric interactions. It was clear that alkylation of cysteine-46 could proceed with the S-enantiomer bound in this way, but not with the R-enantiomer. Model building thus agrees with the inactivation kinetics and indicates the structural origin of the enantioselectivity.  相似文献   

8.
The serine and cysteine proteinases represent two important classes of enzymes that use a catalytic triad to hydrolyze peptides and esters. The active site of the serine proteinases consists of three key residues, Asp...His...Ser. The hydroxyl group of serine functions as a nucleophile and the imidazole ring of histidine functions as a general acid/general base during catalysis. Similarly, the active site of the cysteine proteinases also involves three key residues: Asn, His, and Cys. The active site of the cysteine proteinases is generally believed to exist as a zwitterion (Asn...His+...Cys-) with the thiolate anion of the cysteine functioning as a nucleophile during the initial stages of catalysis. Curiously, the mutant serine proteinases, thiol subtilisin and thiol trypsin, which have the hybrid Asp...His...Cys triad, are almost catalytically inert. In this study, ab initio Hartree-Fock calculations have been performed on the active sites of papain and the mutant serine proteinase S195C rat trypsin. These calculations predict that the active site of papain exists predominately as a zwitterion (Cys-...His+...Asn). However, similar calculations on S195C rat trypsin demonstrate that the thiol mutant is unable to form a reactive thiolate anion prior to catalysis. Furthermore, structural comparisons between native papain and S195C rat trypsin have demonstrated that the spatial juxtapositions of the triad residues have been inverted in the serine and cysteine proteinases and, on this basis, I argue that it is impossible to convert a serine proteinase to a cysteine proteinase by site-directed mutagenesis.  相似文献   

9.
In phosphate buffer at pH 7.0, 5,5'-dithio-bis(2-nitrobenzoic acid), N-ethylmaleimide or iodoacetamide do not alter the activity of beef liver glutamate dehydrogenase. Iodoacetate, however, inactivities the enzyme irreversibility by alkylation. Combined addition of the coenzyme NADH and the substrate 2-oxoglutarate or the effector GTP protects against this inactivation. The alkylation reaction is independent of pH between pH 6-9 indicating that amino, imidazole or phenolic groups are probably not involved in this reaction. Titration of the thiol groups, after denaturation of the enzyme, revealed the loss of approximately one group per polypeptide chain. However, this is not due to the exclusive alkylation of a cysteine residue, since alkylation with iodo-[2-14C]acetic acid also labels a methionine residue. 50% of the label is incorporated into methionine-169 and only 7% into cysteine-115, the remaining radioactivity is distributed in minor quantities (4%) in several unidentified residues. A probable cause of the erroneous thiol groups titration is discussed.  相似文献   

10.
Factors have been investigated which govern the electrophilic reactivity of alkyl halides with thiolate anions in aqueous solution. In the series of alkyl halides studied, some are potential metal-directed affinity labels, while others are frequently used in protein modification. Previous data on the kinetics of this type of alkylation are compared with the present results. The influence of electronic, polar, and steric factors on alkyl halide reactivity is seen. The following order of reactivity for alkyl halides bearing different α substituents was observed: RCH2CH(X)COOCH3 > RCH2CH(X)CONH2 > RCH2CH(X)COOH > RCH2CH2X > RCH2CH(X)CH2OH. The metal-directed affinity labels are imidazole derivatives, some of which have substituents in their imidazole ring. The effect of the imidazole ring and of ring substitution on reactivity is seen. The nucleophilic reactivity of thiols is highly pH dependent since the thiolate anion (RS?) is the reactive species, but only minor differences emerged between different free thiolates.  相似文献   

11.
Kinetic analysis of inactivation of isocitrate lyase from Pseudomonas indigofera by 3-bromopyruvate established that enzyme binds this compound prior to alkylation and that substrate, Ds-isocitrate, competes for the same site on the enzyme. The rate of inactivation was increased by EDTA which is a promoter of catalysis in the presence of activated (reduced) enzyme and substrate. The combination of products, glyoxylate plus succinate, also protected against inactivation. Glyoxylate plus itaconate, phosphoenolpyruvate, or maleate also protected. However, each of the latter three compounds or glyoxylate or succinate alone provided little or no protection. Pyruvate, a competitive inhibitor with respect to glyoxylate in the condensation reaction, also failed to protect. However, two dicarboxylates, meso-tartrate and oxalate, that are also competitive inhibitors with respect to glyoxylate provide some protection against inactivation by BrP perhaps by bridging across cationic sites that facilitate glyoxylate and succinate binding. These and other results imply that alkylation by 3-bromopyruvate occurs at the succinate part of the active site. A mechanism which includes a catalytic role for the cysteine residue at the active site is presented and discussed.  相似文献   

12.
A K Knap  R F Pratt 《Proteins》1989,6(3):316-323
The RTEM-1 thiol beta-lactamase (Sigal, I.S., Harwood, B.G., Arentzen, R., Proc. Natl. Acad. Sci. U.S.A. 79:7157-7160, 1982) is inactivated by thiol-selective reagents such as iodoacetamide, methyl methanethiosulfonate, and 4,4'-dipyridyldisulfide, which modify the active site thiol group. The pH-rate profiles of these inactivation reactions show that there are two nucleophilic forms of the enzyme, EH2 and EH, both of which, by analogy with the situation with cysteine proteinases, probably contain the active site nucleophile in the thiolate form. The pKa of the active site thiol is therefore shown by the data to be below 4.0. This low pKa is thought to reflect the presence of adjacent functionality which stabilizes the thiolate anion. The low nucleophilicity of the thiolate in both EH2 and EH, with respect to that of cysteine proteinases and model compounds, suggests that the thiolate of the thiol beta-lactamase is stabilized by two hydrogen-bond donors. One of these, of pKa greater than 9.0, is suggested to be the conserved and essential Lys-73 ammonium group, while the identity of the other group, of pKa around 6.7, is less clear, but may be the conserved Glu-166 carboxylic acid. beta-Lactamase activity is associated with the EH2 form, and thus the beta-lactamase active site is proposed to contain one basic or nucleophilic group (the thiolate in the thiol beta-lactamase) and two acidic (hydrogen-bond donor) groups (one of which is likely to be the above-mentioned lysine ammonium group).  相似文献   

13.
Phosphoenolpyruvate carboxylase [EC 4.1.1.31] from Escherichia coli W was alkylated by incubation with bromopyruvate, substrate analog, leading to irreversible inactivation. The reaction followed pseudo-first-order kinetics. Mg2+, an essential cofactor for catalysis, enhanced the inactivation, and the enhancing effect increased as the pH increased. The inactivation rate showed a tendency to saturate with increasing concentrations of bromopyruvate, indicating that an enzyme-bromopyruvate complex was formed prior to the alkylation. DL-Phospholactate, a potent competitive inhibitor with respect to phosphoenolpyruvate, protected the enzyme from inactivation in a competitive manner. Examination of the acid hydrolysate of the enzyme modified with [14C]bromopyruvate by paper chromatography showed that radioactivity was solely incorporated into carboxyhydroxyethyl cysteine. In addition, determination of sulfhydryl groups of the native and modified enzymes with 5,5'-dithiobis(2-nitrobenzoate) showed that inactivation occurred concomitant with the modification of one cysteinyl residue per subunit. The results indicate that bromopyruvate reacted with the enzyme as an active-site-directed reagent.  相似文献   

14.
Alcohol dehydrogenase from horse liver is shown to catalyze ester hydrolysis. Nicotinamide coenzymes do not affect the rate of esterolysis. A kinetic approach to study esterase reaction at low substrate to enzyme ratio is described. Kinetic effects of ester structure, temperature, pH, solvent polarity, and ionic strength were investigated. The liver enzyme enhances the rate of esterolysis by lowering activation energy of reaction according to the Uni-Bi kinetic sequence. Two ionizable groups, cysteine and lysine, are tentatively assigned at the esterolytic site of liver alcohol dehydrogenase from pH-rate profiles and chemical modification studies. A plausible mechanism for the esterase reaction proceeds via the acid-assisted nucleophilic catalysis involving the ammonium ion of lysine and the thiolate of cysteine in the acyl-oxygen cleavage.  相似文献   

15.
The fate of the iodide liberated during carboxymethylation of Cys-46 in horse liver alcohol dehydrogenase has been determined with 125I-labeled iodoacetate. The [125I]iodoacetic acid was prepared from mesyloxyacetic acid and sodium [125I]iodide. When carboxymethylation of the enzyme is carried out in solution or in the crystalline state, no iodide is bound to the protein. The rate of iodide during the reaction of iodoacetate, determined with an iodide-specific electrode, has been found to be biphasic: the fast phase corresponds to the carboxymethylation and the slow phase to iodide liberation due to the presence of protein. With 3-iodopropionate (2.5 mM), no inactivation was detected, but in the presence of the enzyme, 10 equivalents of iodide were liberated per subunit in 1 hr. NADH does not inhibit this reaction. The electron density attributed to an iodide bound to the zinc atom of the crystalline enzyme is reinterpreted in view of these results as due to an imidazole bound to the active-site zinc. In the carboxymethylation, the reactivity of bromoacetate is higher than that of iodoacetate.  相似文献   

16.
The enzyme N(1)-(5'-phosphoribosyl) adenosine-5'-monophosphate cyclohydrolase (PR-AMP cyclohydrolase) is a Zn(2+) metalloprotein encoded by the hisI gene. It catalyzes the third step of histidine biosynthesis, an uncommon ring-opening of a purine heterocycle for use in primary metabolism. A three-dimensional structure of the enzyme from Methanobacterium thermoautotrophicum has revealed that three conserved cysteine residues occur at the dimer interface and likely form the catalytic site. To investigate the functions of these cysteines in the enzyme from Methanococcus vannielii, a series of biochemical studies were pursued to test the basic hypothesis regarding their roles in catalysis. Inactivation of the enzyme activity by methyl methane thiosulfonate (MMTS) or 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB) also compromised the Zn(2+) binding properties of the protein inducing loss of up to 90% of the metal. Overall reaction stoichiometry and the potassium cyanide (KCN) induced cleavage of the protein suggested that all three cysteines were modified in the process. The enzyme was protected from DTNB-induced inactivation by inclusion of the substrate N(1)-(5'-phosphoribosyl)adenosine 5'-monophosphate; (PR-AMP), while Mg(2+), a metal required for catalytic activity, enhanced the rate of inactivation. Site-directed mutations of the conserved C93, C109, C116 and the double mutant C109/C116 were prepared and analyzed for catalytic activity, Zn(2+) content, and reactivity with DTNB. Substitution of alanine for each of the conserved cysteines showed no measurable catalytic activity, and only the C116A was still capable of binding Zn(2+). Reactions of DTNB with the C109A/C116A double mutant showed that C93 is completely modified within 0.5 s. A model consistent with these data involves a DTNB-induced mixed disulfide linkage between C93 and C109 or C116, followed by ejection of the active site Zn(2+) and provides further evidence that the Zn(2+) coordination site involves the three conserved cysteine residues. The C93 reactivity is modulated by the presence of the Zn(2+) and Mg(2+) and substantiates the role of this residue as a metal ligand. In addition, Mg(2+) ligand binding site(s) indicated by the structural analysis were probed by site-directed mutagenesis of three key aspartate residues flanking the conserved C93 which were shown to have a functional impact on catalysis, cysteine activation, and metal (zinc) binding capacity. The unique amino acid sequence, the dynamic properties of the cysteine ligands involved in Zn(2+) coordination, and the requirement for a second metal (Mg(2+)) are discussed in the context of their roles in catalysis. The results are consistent with a Zn(2+)-mediated activation of H(2)O mechanism involving histidine as a general base that has features similar to but distinct from those of previously characterized purine and pyrimidine deaminases.  相似文献   

17.
The enzyme UDP-N-acetylglucosamine (UDP-NAG) enolpyruvyltransferase (MurA) catalyzes the formation of enolpyruvyl-UDP-NAG, a precursor in peptidoglycan biosynthesis. The residue at position 115 in MurA has been proposed to act as a general acid in the enzymatic reaction. This is also the primary site of action of the antibiotic fosfomycin. In this paper, the pK(a) of Cys-115 has been determined to be 8.3, by titration of Enterobacter cloacae MurA with the alkylating agent iodoacetamide as a function of pH. Use of site-directed mutagenesis has established that only C115 is essential for catalysis, and the three other cysteine residues (C251, C354, and C381) are nonessential. Mass spectrometric analysis demonstrated that C115 is not alkylated at pH <7, but is alkylated significantly at pH >7. Measurement of the enzymatic inhibition by iodoacetamide as a function of pH showed maximum inhibition at pH >9, with a second-order rate constant of inhibition of 44 M(-)(1) s(-)(1) at pH 10. The presence of either one of the substrates did not influence the inactivation behavior, while the presence of both substrates resulted in a 5-fold reduction in the extent of alkylation. The covalent species that results from PEP bound to C115 of MurA exhibited 50-100-fold increased resistance against alkylation by iodoacetamide. These results imply that C115 is appreciably protonated at physiological pH and, therefore, is capable of acting as a proton donor in the enzyme-catalyzed reaction. However, it also implies that C115 is appreciably deprotonated at physiological pH also, whereupon the resultant thiolate nucleophile may play an important role in the formation of the covalent O-phosphothioketal species, whose role in catalysis is yet to be established.  相似文献   

18.
Treatment of aconitase with phenacyl bromide prior to activation with Fe(II) and reductant results in complete, irreversible enzyme inactivation. Inactivation is due to the alkylation of a cysteine residue at the active site of the enzyme, the inactivation being inhibited by the competitive inhibitor, tricarballylate. Active enzyme is similarly inactivated, citrate affording greater protection than tricarballylate.  相似文献   

19.
M R Eftink  R L Biltonen 《Biochemistry》1983,22(22):5134-5140
Various kinetic aspects of the nonenzymatic hydrolysis of cytidine cyclic 2',3'-phosphate and uridine cyclic 2',3'-phosphate have been studied in order to provide a basis for comparison with the ribonuclease A catalyzed hydrolysis reaction. Studies of the pH dependence of the nonenzymatic reaction reveal mechanisms that are first order in hydroxide concentration and second order in hydrogen ion concentration, in addition to a "water" reaction. The rate constant for the water reaction was found to be very small, approximately equal to 2.5 X 10(-6) min-1. General base catalyzed hydrolysis reactions were also studied with imidazole as the catalyst. At pH values in which both the protonated and neutral forms of imidazole are present, a kinetic mechanism was observed that appears to be second order in total imidazole concentration, thus suggesting that bifunctional catalysis occurs. The activation enthalpy for the hydroxide, hydrogen ion, water, and imidazole catalyzed reactions was determined.  相似文献   

20.
The experimental data presented in this paper comprise kinetic deuterium isotope effects on acylation of papain with various substrates when conducted in H2O and 2H2O. With alkyl esters of N-acylamino acids there is no or very little isotope effect, whereas with N-acylamino acid amides the ratio kappa H2O/kappa 2H2O is less than 1, i.e. there is an inverse isotope effect. Similarly, alkylation of papain with methyl bromoacetate exhibits no kinetic isotope effect, whereas for the analogous alkylation with bromoacetamide an inverse isotope effect is observed. It is concluded that (a) general base catalysis does not occur in the acylation of papain and (b) kinetic deuterium isotope effects can be affected substantially by interaction between the substrate leaving group and the enzyme, which has not been considered in previous mechanistic investigations.  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