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.
The steady-state kinetic parameters for pig liver carboxylesterase (PLE)-catalyzed hydrolysis of the prochiral substrate dimethyl phenylmalonate (DMPM) (product enantioselectivity) and the separate enantiomers of three chiral 2-phenylpropionic acid esters (substrate enantioselectivity) were measured at seven temperatures between 288 K and 312 K. Arrhenius plots of turnover numbers against the reciprocal of experimental temperatures yielded enthalpies and entropies of activation at enzyme saturation. (+)-(S)-methyl-2-phenylpropionate, (+)-(S)-4-nitrophenyl 2-phenylpropionate, and both enantiomers of phenyl 2-phenylpropionate showed very similar activation enthalpies and entropies (approximately 50 kJ mol?1 and ?50 J mol?1 K?1, respectively), but differences were observed for (?)-(R)-methyl 2-phenylpropionate and (?)-(R)-4-nitrophenyl 2-phenylpropionate. Whereas the entropies of activation of all 2-phenylpropionates were negative, positive entropies of activation were measured in the formation of monomethyl phenylmalonate enantiomers from DMPM. Enthalpy–entropy compensation analysis of the data indicates a common mechanism of PLE substrate and product enantiospecificity in the reactions studied here. © 1994 Wiley-Liss, Inc.  相似文献   

2.
The kinetics of AMP-aminohydrolase, which under steady state conditions shows a typical sigmoid dependence of initial velocities versus substrate concentration, have been examined by rapid mixing methods. Using this technique it was observed that when substrate or substrate plus activator (K(+)) were mixed with enzyme, the rate of appearance of product markedly increased during the first few tenths of a second. The time course of this change in rate was taken to reflect the progress of activation by substrate or by K(+). On the other hand, addition of activator to enzyme prior to mixing with substrate gave process curves for the formation of product consistent with normal Michaelis-Menten behaviour.Under the conditions where the reaction was examined, the enzyme at time zero had less than 10% of the activity of the fully active enzyme. The time course for activation with K(+) followed a first order process with a rate constant of 10.6 sec(-1) at 20 degrees C. A simple mechanism consistent with the data and capable of explaining the sigmoid dependence of initial velocities versus substrate concentrations observed in steady state kinetics was proposed.  相似文献   

3.
The role of the activation peptide in determining the substrate specificity of intrinsic pathway factor X (fX) activation was studied by using a novel derivative of fX in which 49 residues were removed enzymatically from the NH2 terminus of the 52-residue activation peptide by an enzyme from the venom of the snake Agkistrodon rhodostoma. The modified protein, designated fXdes-143-191, is inactive but is activated to alpha-fXa by either the intrinsic fX activation complex (intrinsic fXase) composed of factor IXa beta, thrombin-activated factor VIII (fVIIIaIIa), and phospholipid vesicles or by the fX coagulant protein from Russell's viper venom (RVV-XCP). Both the Km and kcat for the activation of fX by RVV-XCP were greater than for fXdes-143-191, resulting in less than a 2-fold difference in the catalytic efficiency (kcat/Km) suggestive of nonproductive binding of fXdes-143-191 to RVV-XCP. The activation of each substrate by intrinsic fXase revealed that the kcat was 100-fold greater for fX than fXdes-143-191 (16 and 0.16 s-1, respectively), although there was no detectable difference in Km (60 and 80 nM, respectively). Activations by fIXa beta/phospholipid in the absence of fVIIIaIIa also revealed a difference in kcat but not Km, but the difference in kcat was smaller (kcat of 0.007 and 0.002 s-1 and Km of 220 and 170 nM for fX and fXdes-143-191, respectively). Analysis of product versus time curves demonstrated that fVIIIaIIa promotes formation of the actyl-enzyme intermediate during fX activation. We conclude that the activation peptide plays a critical role during acyl-enzyme formation that is most pronounced in the presence of fVIIIaIIa. The absence of Km differences suggests that residues NH2-terminal to P3 do not contribute to the initial formation of the enzyme-substrate complex.  相似文献   

4.
K Wimalasena  K R Alliston 《Biochemistry》1999,38(45):14916-14926
Previous studies have shown that the dopamine beta-monooxygenase (DbetaM; E.C. 1.14.17.1)/1-(2-aminoethyl)-1,4-cyclohexadiene (CHDEA) reaction partitions between side chain and ring H-abstraction to produce the side-chain-hydroxylated product, 2-amino-1-(1, 4-cyclohexadienyl)ethanol, and the aromatized product, phenylethylamine, and that the two pathways do not crossover. [Wimalasena, K., and May, S. W. (1989) J. Am. Chem. Soc. 111, 2729-2731; Wimalasena, K., and Alliston, K. R. (1995) J. Am. Chem. Soc. 117, 1220-1224]. We now report that the ring H-abstraction pathway of the reaction further partitions to produce the ring hydroxylated product, CHDEA-6OH, and the aromatized product, PEA, at the carbon-oxygen bond formation step. The ring hydroxylation is shown to be stereospecific, exclusively producing the (S) product. The absolute stereospecificity of the ring and side-chain hydroxylations of the DbetaM/CHDEA reaction suggests that the side-chain pro-R hydrogen of the enzyme-bound substrate is close to perpendicular to the aromatic ring of the phenylethylamine substrate or cyclohexadiene ring of CHDEA. The relative activation energy parameters suggest that the partitioning of the ring H abstraction pathway between aromatized and ring hydroxylated products is due to the partitioning of the high-energy intermediates, the cyclohexadienyl radical and the Cu(II)-O(*) species, between carbon-oxygen bond formation and direct electron transfer. The relatively high activation enthalpic favorability and entropic unfavorability for the carbon-oxygen bond formation strongly suggest that the critical balancing of these two opposing forces is mandatory for the desired product formation.  相似文献   

5.
Hydroxylations of octane and lauric acid by Cytochrome P450-BM3 (CYP102A1) wild-type and three active site mutants--F87A, L188Q/A74G, and F87V/L188Q/A74G--were rationalized using a combination of substrate orientation from docking, substrate binding statistics from molecular dynamics simulations, and barrier energies for hydrogen atom abstraction from quantum mechanical calculations. Wild-type BM3 typically hydroxylates medium- to long-chain fatty acids on subterminal (omega-1, omega-2, omega-3) but not the terminal (omega) positions. The known carboxylic anchoring site Y51/R47 for lauric acid, and hydrophobic interactions and steric exclusion, mainly by F87, for octane as well as lauric acid, play a role in the binding modes of the substrates. Electrostatic interactions between the protein and the substrate strongly modulate the substrate's regiodependent activation barriers. A combination of the binding statistics and the activation barriers of hydrogen-atom abstraction in the substrates is proposed to determine the product formation. Trends observed in experimental product formation for octane and lauric acid by wild-type BM3 and the three active site mutants were qualitatively explained. It is concluded that the combination of substrate binding statistics and hydrogen-atom abstraction barrier energies is a valuable tool to rationalize substrate binding and product formation and constitutes an important step toward prediction of product ratios.  相似文献   

6.
With and without p-chlorophenol as an activator, the rates of hydrolysis of p-nitrophenyl acetate catalyzed by alpha-chymotrypsin were measured at pressures up to 2 kbar at 25 degrees C. From the pressure dependence of the rate constant (kcat)A and (kcat)0 of the product formation with and without an activator, the activation volumes (delta V not equal to cat)A and (delta not equal to cat)0 were +2 and -6 +/- 1 cm3.mol-1. From the pressure dependence of the equilibrium constant (KA) of incorporation of p-chlorophenol into the enzyme, the volume change (delta VA) was -10 +/- 1 cm3.mol-1. The mechanisms of the substrate activation are discussed in terms of the activation and reaction volumes.  相似文献   

7.
The present investigation deals with role of Ca++ ions in increasing the yield of citric acid in a repeated-batch cultivation system (working volume 9-1) and its kinetic basis. Five different hyper-producing strains of Aspergillus niger were evaluated for citric acid production using clarified cane-molasses as basal substrate. Among the cultures, NGGCB101 (developed by u.v./chemical mutation in our labs) gave maximum production of citric acid i.e., 87.98 g/1, 6 days after mycelial inoculation. The addition of CaCl2 to the culture medium promoted the formation of small rounded fluffy pellets (1.55 mm, diameter), which were desirable for citric acid productivity. CaCl2 at a level of 2.0 M, added during inoculation time, was optimized for commercial exploitation of molasses. During repeated-batch culturing, a yield of citric acid monohydrate of 128.68 g/1 was obtained when the sampling vs. substrate feeding was maintained at 4-1 (44.50% working volume). The incubation period was reduced from 6 to only 2 days. The values of kinetic parameters such as substrate consumption and product formation rates revealed the hyperproducibility of citric acid by the selected Aspergillus niger NGGCB101 (LSD = 0.456a, HS). Case studies are highly economical because of higher yield of product, lower energy consumption and the use of raw substrate without any additional supplementation.  相似文献   

8.
A conserved threonine found in the majority of cytochromes P450 (P450s) has been implicated in the activation of dioxygen during the catalytic cycle. P450cin (CYP176A) has been found to be an exception to this paradigm, where the conserved threonine has been replaced with an asparagine. Prior studies with a P450cin N242A mutant established that the Asn-242 was not a functional replacement for the conserved threonine but was essential for the regio- and stereocontrol of the oxidation of cineole. To explore further how P450cin controls the activation of the dioxygen in the absence of the conserved threonine, two concurrent lines of investigation were followed. Modification of P450cin indicated that the Thr-243 was not involved in controlling the protonation of the hydroperoxy species. In addition, the N242T mutant did not enhance the rate and/or efficiency of catalytic turnover of cineole by P450cin. In parallel experiments, the substrate cineole was modified by removing the ethereal oxygen to produce camphane or 2,2-dimethylbicyclo[2.2.2]octane (cinane). An analogous experiment with P450EryF showed that a hydroxyl group on the substrate was vital, and in its absence catalytic turnover was effectively abolished. Catalytic turnover of P450cin with either of these alternative substrates (camphane or cinane) revealed that in the absence of the ethereal oxygen there was still a significant amount of coupling of the NADPH-reducing equivalents to the formation of oxidised product. Again the substrate itself was not found to be important in controlling oxygen activation, in contrast to P450EryF, but was shown to be essential for regio- and stereoselective substrate oxidation. Thus, it still remains unclear how dioxygen activation in the catalytic turnover of cineole by P450cin is controlled.  相似文献   

9.
Substrate and energy costs of the production of exocellular enzymes from glucose and citrate by B. Iicheniformis S1684 as well as molar growth yields corrected for these costs of product formation were calculated using data from chemostat experiments. The calculations showed that 1.46-1.73 mol glucose and 2.31-2.77 mol citrate are needed for formation and excretion of 1 mol protein. Consequently, the values of the maximal product yield from substrate (Y(psm') g/mol) are 80 < Y(psm) < 95 when product is formed from glucose and 50 < Y(psm) < 60 when product is formed from citrate. The higher substrate costs for product formation from citrate are due to a higher level of CO(2) production during protein formation and a higher substrate requirement for the energy supply of product formation and excretion than when product is formed from glucose. The theoretical ATP requirement for protein synthesis could be determined reasonably well, but the energy costs of protein excretion could not be determined exactly. The energy costs of protein formation are higher than those of biomass formation or protein excretion. Molar growth yields corrected for the substrate costs of product formation were high, indicating a high efficiency of growth.Growth and production parameters were determined as well from experimental data of recycling fermentor experiments using a parameter optimization procedure based on a mathematical model describing biomass growth as a linear function of the substrate consumption rate and the rate of product formation as a linear function of biomass growth rate. The fitting procedure yielded two growth and production domains during glucose limitation. In the first domain the values for the maximal growth yield and maintenance coefficient were in agreement with those found in chemostat experiments at corresponding values of Y(spm). Domain 2 could be described best with linear growth and product formation. In domain 2 the rate of product formation decreased and more substrate became available for biomass formation. As a consequence the specific growth rate increased in the shift from domain 1 to 2. Domain 2 behavior most probably is caused by the rel-status of B. Iicheniformis S1684.  相似文献   

10.
Joseph E  Wei W  Tittmann K  Jordan F 《Biochemistry》2006,45(45):13517-13527
The X-ray crystal structure of pyruvamide-activated yeast pyruvate decarboxylase (YPDC) revealed a flexible loop spanning residues 290 to 304 on the beta-domain of the enzyme, not seen in the absence of pyruvamide, a substrate activator surrogate. Site-directed mutagenesis studies revealed that residues on the loop affect the activity, with some residues reducing k(cat)/K(m) by at least 1000-fold. In the pyruvamide-activated form, the loop located on the beta domain can transfer information to the active center thiamin diphosphate (ThDP) located at the interface of the alpha and gamma domains. The sigmoidal v(0)-[S] curve with wild-type YPDC attributed to substrate activation is modulated for most variants, but is not abolished. Pre-steady-state stopped-flow studies for product formation on these loop variants provided evidence for three enzyme conformations connected by two transitions, as already noted for the wild-type YPDC at pH 5.0 [Sergienko, E. A., and Jordan, F. (2002) Biochemistry 41, 3952-3967]. (1)H NMR analysis of the intermediate distribution resulting from acid quench [Tittmann et al. (2003) Biochemistry 42, 7885-7891] with all YPDC variants indicated that product release is rate limiting in the steady state. Apparently, the loop is not solely responsible for the substrate activation behavior, rather it may affect the behavior of residue C221 identified as the trigger for substrate activation. The most important function of the loop is to control the conformational equilibrium between the "open" and "closed" conformations of the enzyme identified in the pyruvamide-activated structure [Lu et al. (2000) Eur. J. Biochem. 267, 861-868].  相似文献   

11.
The aqueous substrate and product toxicity thresholds in the microbial biotransformation of (-)-trans-carveol to the fragrance/flavor compound (R)-(-)-carvone by Rhodococcus erythropolis were determined. Above aqueous phase concentrations of approx. 500 mg carveol/l and 200-600 mg carvone/l, the biotransformation activity of the biocatalyst was inhibited. This biotransformation was undertaken in a single aqueous phase 3 l [corrected] reactor in which a total of 5 ml carveol (mixture of isomers) was added before the biotransformation rate decreased significantly. The carvone volumetric productivity was 31 mg/lh. Although the growth of the organism post-exposure was not affected, dramatic morphological changes in response to the accumulation of the inhibitory substrate and product were observed.  相似文献   

12.
An immunosensor based on surface plasmon resonance (SPR) onto a protein G layer by self-assembly technique was developed for detection ofLegionella pneumophila. The protein G layer by self-assembly technique was fabricated on a gold (Au) surface by adsorbing the 11-mercaptoundecanoic acid (MUA) and an activation process for the chemical binding of the free amine (-NH2) of protein G and 11-(MUA) using 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide hydrochloride (EDAC) in series. The formation of the protein G layer by self-assembly technique on the Au substrate and the binding of the antibody and antigen in series were confirmed by SPR spectroscopy. The surface topographies of the fabricated thin films on an Au substrate were also analyzed by using an atomic force microscope (AFM). Consequently, an immunosensor for the detection ofL. pneumophila using SPR was developed with a detection limit of up to 102 CFU per mL.  相似文献   

13.
Summary Bacillus licheniformis S 1684 is able to produce an alkaline serine protease exocellularly. In glucose-limited chemostat cultures the specific rate of protease production was maximal at a -value of 0.22. Above this growth rate protease production was repressed. Dependent on 10–20% of the glucose input was used for exocellular product formation. The degree of reduction of exocellular products was 4.1.Maximum molar growth yields were high and indicate a high efficiency of growth. The values of Y glu max and YO 2 max were 83.8 and 53.3, respectively. When Y glu max was corrected for the amount of glucose used for product formation a value of 100.3 was obtained. These high maximum molar growth yields are most probably caused by a high Y ATP max . Anaerobic batch experiments showed a Y ATP of 14.6.Sometimes the used strain was instable in cell morphology and protease production. Non-protease producing cells most probably develop from producing cells by mutation in the rel-gene. Producing cells most probably are relaxed (rel -) and non-producing cells stringent (rel +).Glossary specific growth rate (h-1) - Y sub growth yield permol substrate (g biomass/mol) - Y max maximum molar growth yield, corrected for maintenance requirements (g biomass/mol) - Y max(corr) Y max corrected for product formation (g biomass/mol) - m sub maintenance requirements (mol/g biomass·h) - m sub(corr) maintenance requirements corrected for product formation (mol/g biomass·h) - Y c fraction of organic substrate converted in biomass - z fraction of organic substrate converted in exocellular products - d fraction of organic substrate converted in CO2 (g mol/g atom C) - Crec% carbon recovery % - average degree of reduction of exocellular products - P/O amount of ATP produced during electron-transport of 2 electrons to oxygen  相似文献   

14.
Studying the oxygenation kinetics of (19R/S,5Z,8Z,11Z,14Z)-19-hydroxyeicosa-5,8,11,14-tetraenoic acid (19-OH-AA) by rabbit 15-lipoxygenase-1 we observed a pronounced oxygen dependence of the reaction rate, which was not apparent with arachidonic acid as substrate. Moreover, we found that peroxide-dependent activation of the lipoxygenase depended strongly on the oxygen concentration. These data can be described with a kinetic model that extends previous schemes of the lipoxygenase reaction in three essential aspects: (a) the product of 19-OH-AA oxygenation is a less effective lipoxygenase activator than (13S,9Z,11E)-13-hydroperoxyoctadeca-9,11-dienoic acid; (b) molecular dioxygen serves not only as a lipoxygenase substrate, but also impacts peroxide-dependent enzyme activation; (c) there is a leakage of radical intermediates from the catalytic cycle, which leads to the formation of inactive ferrous lipoxygenase. This enzyme inactivation can be reversed by another round of peroxide-dependent activation. Taken together our data indicate that both peroxide activation and the oxygen affinity of lipoxygenases depend strongly on the chemistry of the lipid substrate. These findings are of biological relevance as variations of the reaction conditions may turn the lipoxygenase reaction into an efficient source of free radicals.  相似文献   

15.
A Candida rugosa lipase immobilized on polypropylene powder was employed as the biocatalyst for the enantioselective hydrolysis of (R, S)-suprofen 2,2,2-trifluorothioester in cyclohexane, in which trioctylamine was added as the catalyst to perform in situ racemization of the remaining (R)-thioester. A hollow-fiber membrane was also integrated with the dynamic kinetic resolution process in order to continuously extract the desired (S)-suprofen into an aqueous solution containing NaOH. A kinetic model for the whole process (operating in batch and feed-batch modes) was developed, in which enzymatic hydrolysis and deactivation, lipase activation, racemization and non-enantioselective hydrolysis of the substrate by trioctylamine, and reactive extraction of (R)- and (S)-suprofen into the aqueous phase in the membrane were considered. Theoretical predictions from the model for the time-course variations of substrate and product concentrations in each phase were compared with experimental data.  相似文献   

16.
A number of vicinal diols were found to react with propanediol dehydratase, typically resulting in the conversion of enzyme-bound adenosylcobalamin to cob(II)alamin and formation of aldehyde or ketone derives from substrate. Moreover, all are capable of effecting the irreversible inactivation of the enzyme. The kinetics and mechanism of product formation and inactivation were investigated. Glycerol, found to be a very good substrate for diol dehydratase as well as a potent inactivator, atypically, did not induce cob(II)alamin formation to any detectable extent. With glycerol, the inactivation process was accompanied by conversion of enzyme-bound adenosylcobalamin to an alkyl or thiol cobalamin, probably by substitution of an amino acid chain near the active site for the 5'-deoxy-5'-adenosyl ligand on the cobalamin. The inactivation reaction with glycerol as the inactivator exhibits a deuterium isotope effect of 14, strongly implicating hydrogen transfer as an important step in the mechanism of inactivation. The isotope effect on the rate of product formation was found to be 8.0. Experiments with isotopically substituted glycerols indicate that diol dehydrase distinguishes between "R" and "S" binding conformations, the enzyme-(R)-glycerol complex being predominately responsible for the product-forming reaction, while the enzyme-(S)-glycerol complex results primarily in the activation reaction. Mechanistic implications are discussed. A method for removing enzyme-bound hydroxycobalamin that is nondestructive to the enzyme and a technique for measuring the binding constants of (R)- and (S)-1,2-propanediols are presented.  相似文献   

17.
Whole cells and cell-free extracts derived from Rhodococcus rhodochrous NCIMB 11216 were shown to hydrolyse both aliphatic and aromatic nitriles, when the organism had been grown on either propionitrile or benzonitrile as the source of carbon and nitrogen. Whole cell suspensions and cell-free extracts derived from bacteria grown on either substrate were able to biotransform R-(-),S-(+)-2-methylbutyronitrile. The S-(+) enantiomer was biotransformed more rapidly than the the R-(-) enantiomer. For whole cell biotransformations at 30°C, the maximum enantiomeric excess (ee) of the remaining R-(-)-2-methylbutyronitrile was 93% when 70% of the R-(-) enantiomer had been converted to the product, 2-methylbutyric acid. For the corresponding biotransformation at 4°C, there was an ee of 93% for the residual R-(-) enantiomer of the substrate when only 60% of it had been converted to product. For biotransformations by cell-free extracts at 30°C the 2-methylbutyric acid product had an ee of 17% for the S-(+) enantiomer at the time of optimal ee for the remaining R-(-) enantiomer of the substrate. In contrast, when the reaction was carried out by whole cells, the ee for the product acid was 0.36%. This was probably due to further, non-selective metabolism of the acid, which was especially significant at the beginning of the reaction. At both temperatures, the ee for the S-(+) enantiomer of 2-methylbutyric acid was at a maximum in the early stage of the biotransformation; for example, at 4°C the maximum detectable ee was 100% when the yield was 11%.Abbreviations EDTA Ethylenediaminetetraacetic acid - ee enantiomeric excess - FID flame ionisation detector - GC gas chromatography - 1HNMR H nuclear magnetic resonance - K m Michaelis constant - NCIMB National Collection of Industrial and Marine Bacteria - td doubling time - V max Maximum velocity  相似文献   

18.
19.
Yeast AMP deaminase is allosterically activated by ATP and MgATP and inhibited by GTP and PO4. The tetrameric enzyme binds 2 mol each of ATP, GTP, and PO4/subunit with Kd values of 8.4 +/- 4.0, 4.1 +/- 0.6, and 169 +/- 12 microM, respectively. At 0.7 M KCl, ATP binds to the enzyme, but no longer activates. Titration with coformycin 5'-monophosphate, a slow, tight-binding inhibitor, indicates a single catalytic site/subunit. ATP and GTP bind at regulatory sites distinct from the catalytic site and their binding is mutually exclusive. Inorganic phosphate competes poorly with ATP for the ATP sites (Kd = 20.1 +/- 4.1 mM). However, near-saturating ATP reduces the moles of phosphate bound per subunit to 1 PO4, which binds with a Kd = 275 +/- 22 microM. In the presence of ATP, PO4 cannot effectively compete with ATP for the nucleotide triphosphate sites. The PO4 which binds in the presence of ATP is competitive with AMP at the catalytic site since the Kd equals the kinetic inhibition constant for PO4. Initial reaction rate curves are a cooperative function of AMP concentration and activation by ATP is also cooperative. However, no cooperativity is observed in the binding of any of the regulator ligands and ATP binding and kinetic activation by ATP is independent of substrate analog concentration. Cooperativity in initial rate curves results, therefore, from altered rate constants for product formation from each (enzyme.substrate)n species and not from cooperative substrate binding. The traditional cooperative binding models of allosteric regulation do not apply to yeast AMP deaminase, which regulates catalytic activity by kinetic control of product formation. The data are used to estimate the rates of AMP hydrolysis under reported metabolite concentrations in yeast.  相似文献   

20.
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   

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