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.
The force driving the conversion from the acyl intermediate to the tetrahedral intermediate in the deacylation reaction of serine proteases remains unclear. The crystal structure of 6‐guanidinohexanoyl trypsin was determined at pH 7.0, near the optimum reaction pH, at 1.94 Å resolution. In this structure, three water molecules are observed around the catalytic site. One acts as a nucleophile to attack the acyl carbonyl carbon while the other two waters fix the position of the catalytic water through a hydrogen bond. When the acyl carbonyl oxygen oscillates thermally, the water assumes an appropriate angle to catalyze the deacylation. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

2.
L D Byers  H S She  A Alayoff 《Biochemistry》1979,18(12):2471-2480
The glycolytic enzyme glyceraldehyde-3-phosphate dehydrogenase catalyzes the oxidative phosphorylation of D-glyceraldehyde 3-phosphate. A variety of phosphonates have been shown to substitute for phosphate in this reaction [Gardner, J. H., & Byers, L. D., (1977) J. Biol. Chem. 252, 5925--5927]. The dependence of the logarithm of the equilibrium constant for the reaction on the pKa2 value of the phosphonate is characterized by a Br?nsted coefficient, betaeq, of approximately 1. This represents the sensitivity of the transfer of the phosphoglyceroyl group between the active-site sulfhydryl residue (in the acyl-enzyme intermediate) and the acyl acceptor on the basicity of the acyl acceptor. Molybdate (MoO42-) can also serve as an acyl acceptor in the glyceraldehyde-3-phosphate dehydrogenase catalyzed reaction. The second-order rate constant for the reaction with molybdate is only approximately 12 times lower than the reaction with phosphate even though the pKa2 of molybdate is 3.1 units lower than the pKa2 of phosphate. The immediate product of the molybdate reaction is the acyl molybdate, 1-molybdo-3-phosphoglycerate. The acyl molybdate, like the acyl arsenate (the immediate product of the reaction when arsenate is the acyl acceptor), is kinetically unstable. At pH 7.3 (25 degrees C), the half-life for hydrolysis of the acyl molybdate, or the acyl arsenate, is less than 2.5 s. Thus, hydrolysis of 1-molybdo- and 1-arseno-3-phosphoglycerate is at least 2000 times faster than hydrolysis of 1,3-diphosphoglycerate under the same conditions. Glyceraldehyde-3-phosphate dehydrogenase has a fairly broad specificity for acyl acceptors. Most tetrahedral oxy anions tested are substrates for the enzyme (except SO4(2-) and SeO4(2-)). Tetrahedral monoanions such as ReO4- and GeO(OH)3- are not substrates but do bind to the enzyme. These results suggest the requirement of at least one anionic site on the acyl acceptor required for binding and another anionic group on the acyl receptor required for nucleophilic attack on the acyl enzyme.  相似文献   

3.
The spontaneous hydrolyses of lactic acid O-phenyl phosphate (I) and, to a lesser extent, 3-hydroxybutyric acid O-phenyl phosphate (II) have been investigated and compared with similar intramolecular and bimolecular reactions. Compared to bimolecular nucleophilic reactions, the reactivity of II is similar to other systems involving the formation of a six-membered ring intermediate, which suggests that the electrostatic barrier to attack of an anionic nucleophile on a phosphate diester anion is fully present in II. The reactivity of I, as compared to that of II, would suggest that at least a partial overcoming of the electrostatic barrier takes place upon closer approimation of the two reacting centers. The Mn-2+-catalyzed hydrolysis of I exhibits saturation kinetics, consistent with the enhanced reactivity of the metal ion-substrate complex. The binding constant for this complex, determined from kinetics, is in good agreement with that obtained by electron spin resonance (ESR) titration. It is argued that the complex of Mn-2+ with II, as observed by pulsed Fourier transform nuclear magnetic resonance (NMR) techniques, is a precursor to the complex of catalytic significance. The hydrolysis of I as catalyzed by a variety of divalent metal ions suggests an optimal metal ion size. The spontaneous and metal ion catalyzed hydrolyses of salicyclic acid O-aryl phosphates (IIIa-d) proceed through cyclic acyl phosphate intermediates after expulsion of phenol. Product studies on the parent compound have failed to detect phenyl phosphate as a product in either the spontaneous or metal ion catalyzed process. The dependence of the second-order rate constant for the metal-catalyzed hydrolysis on leaving group pKa, beta-1-g, decreases significantly relative to beta-1-g for the spontaneous hydrolysis. From the collective data a specific interation of the metal ion with a pentacovalent intermediate is inferred in the rate-determining step for esters I and III. The probable consequences of these mechanistic postulates for phosphoryl transfer reactions in biological systems are discussed.  相似文献   

4.
The selectivity of preparations of alpha-chymotrypsin immobilized on Celite or polyamide and carrying out syntheses of di- and tripeptides in acetonitrile medium were studied. The study concerns the effect of mass-transfer limitations on three different kinds of selectivity: acyl donor, stereo- and nucleophile selectivities, defined respectively as the ratio of initial rates with different acyl donors; the enantioselectivity factor (E); and the ratio of initial rates of peptide synthesis and hydrolysis of the acyl donor. Strong mass-transfer limitations caused by increased enzyme loading had a very strong effect on acyl donor selectivity, with reductions of up to 79%, and on stereoselectivity, with reductions of up to 77% in relation to optimum values, both on Celite. Nucleophile selectivity was not affected as strongly by mass-transfer limitations. Using a small molecule (AlaNH(2)) as nucleophile, the onset of these limitations caused only minor reductions in selectivity, while when using a larger nucleophilic species (AlaPheNH(2)) it was reduced by up to 60% when increasing enzyme loading on Celite from 2 to 100 mg/g. The different way these kinds of selectivity are affected by the onset of mass-transfer limitations can be explained by a combination of different aspects: the kinetic behavior of the enzyme toward nucleophile and acyl donor concentrations, the relative concentrations of reagents used in the reaction media, and their relative diffusion coefficients. In short, higher concentrations of nucleophile than acyl donor are generally used, and the nucleophile most often used in the experiments hereby described (AlaNH(2)) diffuses faster than the acyl donors employed. These factors combined are expected to give rise to concentration gradients inside porous biocatalyst particles higher for acyl donor than for nucleophile under conditions of mass-transfer limitations. This explains why acyl donor selectivity and stereoselectivity are much more influenced by mass transfer limitations than nucleophile selectivity.  相似文献   

5.
The dissimilatory nitrite reductase (cytochrome c,d1) from Pseudomonas aeruginosa was observed at pH 7.5 to catalyze nitrosyl transfer (nitrosation) between [15N]nitrite and several N-nucleophiles or H2 18O, with rate enhancement of the order of 10(8) relative to analogous chemical reactions. The reducing system (ascorbate, N,N,N',N'-tetramethylphenylenediamine) could reduce nitrite (but not NO) enzymatically and had essentially no direct chemical reactivity toward nitrite or NO. The N-nitrosations showed saturation kinetics with respect to the nucleophile and, while exhibiting Vmax values which varied by about 40-fold, nevertheless showed little or no dependence of Vmax on nucleophile pKa. The N-nitrosations and NO-2/H2O-18O exchange required the reducing system, whereas NO/H2O-18O exchange was inhibited by the reducing system. NO was not detected to serve as a nitrosyl donor to N-nucleophiles. These and other kinetic observations suggest that the enzymatic nitrosyl donor is an enzyme-bound species derived from reduced enzyme and one molecule of nitrite, possibly a heme-nitrosyl compound (E-FeII X NO+) for which there is precedence. Nitrosyl transfer to N-nucleophiles may occur within a ternary complex of enzyme, nitrite, and nucleophile. Catalysis of nitrosyl transfer by nitrite reductase represents a new class of enzymatic reactions and may present another example of electrophilic catalysis by a metal center. The nitrosyl donor trapped by these reactions is believed to represent an intermediate in the reduction of nitrite by cytochrome c,d1.  相似文献   

6.
The enzyme cis-3-chloroacrylic acid dehalogenase (cis-CaaD) is found in a bacterial pathway that degrades a synthetic nematocide, cis-1,3-dichloropropene, introduced in the 20th century. The previously determined crystal structure of cis-CaaD and its promiscuous phenylpyruvate tautomerase (PPT) activity link this dehalogenase to the tautomerase superfamily, a group of homologous proteins that are characterized by a catalytic amino-terminal proline and a β-α-β structural fold. The low-level PPT activity of cis-CaaD, which may be a vestige of the function of its progenitor, prompted us to search the databases for a homologue of cis-CaaD that was annotated as a putative tautomerase and test both its PPT and cis-CaaD activity. We identified a mycobacterial cis-CaaD homologue (designated MsCCH2) that shares key sequence and active site features with cis-CaaD. Kinetic and 1H NMR spectroscopic studies show that MsCCH2 functions as an efficient PPT and exhibits low-level promiscuous dehalogenase activity, processing both cis- and trans-3-chloroacrylic acid. To further probe the active site of MsCCH2, the enzyme was incubated with 2-oxo-3-pentynoate (2-OP). At pH 8.5, MsCCH2 is inactivated by 2-OP due to the covalent modification of Pro-1, suggesting that Pro-1 functions as a nucleophile at pH 8.5 and attacks 2-OP in a Michael-type reaction. At pH 6.5, however, MsCCH2 exhibits hydratase activity and converts 2-OP to acetopyruvate, which implies that Pro-1 is cationic at pH 6.5 and not functioning as a nucleophile. At pH 7.5, the hydratase and inactivation reactions occur simultaneously. From these results, it can be inferred that Pro-1 of MsCCH2 has a pKa value that lies in between that of a typical tautomerase (pKa of Pro-1~6) and that of cis-CaaD (pKa of Pro-1~9). The shared activities and structural features, coupled with the intermediate pKa of Pro-1, suggest that MsCCH2 could be characteristic of an evolutionary intermediate along the past route for the divergence of cis-CaaD from an unknown superfamily tautomerase. This makes MsCCH2 an ideal candidate for laboratory evolution of its promiscuous dehalogenase activity, which could identify additional features necessary for a fully active cis-CaaD. Such results will provide insight into pathways that could lead to the rapid divergent evolution of an efficient cis-CaaD enzyme.  相似文献   

7.
8.
Tertiary structure in globular RNA folds can create local environments that lead to pKa perturbation of specific nucleotide functional groups. To assess the prevalence of functionally relevant adenosine-specific pKa perturbation in RNA structure, we have altered the nucleotide analog interference mapping (NAIM) approach to include a series of a phosphorothioate-tagged adenosine analogs with shifted N1 pKa values. We have used these analogs to analyze the hairpin ribozyme, a small self-cleaving/ligating RNA catalyst that is proposed to employ a general acid-base reaction mechanism. A single adenosine (A10) within the ribozyme active site displayed an interference pattern consistent with a functionally significant base ionization. The exocyclic amino group of a second adenosine (A38) contributes substantially to hairpin catalysis, but ionization of the nucleotide does not appear to be important for activity. Within the hairpin ribozyme crystal structure, A10 and A38 line opposite edges of a solvent-excluded cavity adjacent to the 5'-OH nucleophile. The results are inconsistent with the model of ribozyme chemistry in which A38 acts as a general acid-base catalyst, and suggest that the hairpin ribozyme uses an alternative mechanism to achieve catalytic rate enhancement that utilizes functional groups within a solvent-excluded cleft in the ribozyme active site.  相似文献   

9.
S K Law  T M Minich  R P Levine 《Biochemistry》1984,23(14):3267-3272
The binding of [3H]glycerol and [3H]putrescine to C3 was studied in a fluid-phase system using trypsin as the C3 convertase. The binding of glycerol showed little variation in the pH range between 6.0 and 10.0. The binding of putrescine (pKa = 9.0) is rather ineffective below pH 7.5 but becomes more efficient as the pH of the reaction mixture increases. These results agree with the contention that the final step of the binding reaction is the transfer of the acyl group of the exposed thio ester of C3 to a nucleophile since the nucleophilicity of hydroxyl groups is rather independent of pH whereas only the unprotonated form of amino groups is nucleophilic. The inefficient reaction of amino groups with the exposed thio ester of C3 is also supported by the study of the inhibitory activity of serine and its two derivatives, N-acetylserine and O-methylserine, to the binding of [3H]glycerol to C3. N-Acetylserine showed an inhibitory activity equivalent to that of serine, whereas O-methylated serine showed only minimal activity. It can be concluded, therefore, that serine reacts with the thio ester of C3 by its hydroxyl group but not by its alpha-amino group. The ability of the alcohol group of various alkanes to inhibit the binding of [3H]glycerol to C3 was also studied. The primary alcohols inhibit the binding reaction with an efficiency that is similar to glycerol, and there are no significant differences in the binding efficiencies of methanol, ethanol, 1-propanol, and 1-butanol.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
Kinetic regularities of the enzymatic acyl group transfer reactions have been studied using ampicillin synthesis catalyzed by E. coli penicillin acylase as an example. It was shown that ampicillin synthesis proceeds through the formation of an acylenzyme–nucleophile complex capable of undergoing hydrolysis. The relative nucleophile reactivity of 6-aminopenicillanic acid (6-APA) is a complex parameter dependent on the nucleophile concentration. The kinetic analysis showed that the maximum yield of antibiotic being synthesized depended only on the nucleophile reactivity of 6-APA, the ratio between the enzyme reactivities with respect to the target product and acyl donor, and the initial concentrations of reagents. The parameters characterizing the nucleophile reactivity of 6-APA have been determined. The algorithm of modeling the enzymatic synthesis has been elaborated. The proposed algorithm allows the kinetics of the process not only in homogeneous, but also in heterogeneous (aqueous solution–precipitate) systems to be quantitatively predicted and described based on experimental values of parameters of the reaction. It was shown that in heterogeneous aqueous solution–precipitate systems PA-catalyzed ampicillin synthesis proceeds much more efficiently compared to the homogeneous solution.  相似文献   

11.
(a) The reactivity of substituted acetates as substrates for CoA transferase increases sharply with increasing basicity and exhibits a slope of 1.0 in a plot of either log kappacat or log (kappacat/Km) against pKa (betanuc = 1.0). This result shows that the catalyzed reaction, which involves both carboxylate activation and leaving group transfer, does not proceed through a fully concerted reaction mechanism in the rate-determining step. The result is consistent with a stepwise reaction mechanism that proceeds through an anhydride intermediate. (b) Equilibrium constants for thiol ester formation, either bound to the enzyme or free in solution, show the same dependence on the basicity of carboxylate ions (betaeq = 1.0) and are independent of acidity when expressed in terms of the carboxylic acid. Thus, the polar environment around substituents on the acyl group is the same for carboxylic acids, thiol esters, and oxygen esters. (c) The interaction of the terminal CH3CO group of acetoacetate with the active site causes a 200,000-fold increase in kappacat/Km, corresponding to a decrease in delta G++ OF 7.2 kcal/mol compared with an unsubstituted acid of the same pK. The binding energy of the coenzyme A moiety of the substrate is utilized to interact with the active site and cause a 10(4) to 10(6)-fold increase in kappacat, corresponding to a decrease in delta G++ of 6 to 9 kcal/mol, compared with fragments of the coenzyme A moiety added separatly or together. (d) The exchange of labeled coenzyme A into acyl-CoA substrates was found to be greater than or equal to 10(5) slower than substrate turnover.  相似文献   

12.
The kinetically controlled condensation reaction of Z-Gly-Trp-Met-OR(1) (R(1): Et, Al, Cam) and H-Asp-(OR(2))-Phe-NH(2) (R(2): H, Bu(t)) catalyzed by alpha-chymotrypsin deposited onto polyamide in organic media was studied. The effect of the drying process of the enzyme-support preparation, substrate concentrations, reaction medium, acyl donor, and nucleophile structure on both enzymatic activity and pentapeptide yield was investigated. The immobilized preparation directly equilibrated at a(w) = 0.113, gave higher enzymatic activities than dried with vacuum first, and then equilibrated at a(w) = 0.113. The addition of triethylamine to the reaction medium increased dramatically the enzymatic activity. However, the pentapeptide yield was affected neither by the drying procedure nor by the addition of triethylamine. The donor ester Z-Gly-Trp-Met-OAl gave initial reaction rates 2.6 times higher than the conventional ethyl ester derivative but rendered similar yields. The best results were obtained using Z-Gly-Trp-Met-OCam as acyl-donor ester; 80% yield and initial reaction rates 4 times higher than the ethyl ester derivative. In all cases, acetonitrile containing Tris-HCl 50 mM pH 9 buffer (0.5% v/v) and triethylamine (0.5% v/v) was found to be the best reaction system. Under these conditions, it was possible to use the nucleophile H-Asp-Phe-NH(2) with beta-unprotected aspartic acid residue. In this case, 50% yield was obtained, but economic considerations could lead to select it as nucleophile. Finally, the fragment condensation reaction was carried out at gram scale, obtaining a 39% yield which included the reaction, removal of protecting groups and purification steps. (c) 1997 John Wiley & Sons, Inc. Biotechnol Bioeng 56: 456-463, 1997.  相似文献   

13.
HlyC, hemolysin-activating lysine-acyltransferase, catalyses the acylation (from acyl-acyl carrier protein [ACP]) of Escherichia coli prohemolysin (proHlyA) on the epsilon-amino groups of specific lysine residues, 564 and 690 of the 1024 amino acid primary structure, to form hemolysin (HlyA). Isothermal titration calorimetry was used to measure the thermodynamic properties of the protein acylation of proHlyA-derived structures, altered by substantial deletions and separation of the acylation sites into two different peptides and site directed mutation analyses of acylation sites. Acylation of proHlyA-derived proteins catalyzed by HlyC was overall an exothermic reaction driven by a negative enthalpy. The reaction, whose kinetics are compatible to a ping-pong mechanism, is composed of two partial reactions. The first, the formation of an acyl-HlyC intermediate, was entropically driven, most likely by noncovalent complex formation between acyl-ACP and HlyC; enthalpy-driven acyl transfer followed, resulting in acyl-HlyC and ACPSH product formation. The second partial reaction was an energetically unfavorable acyl transfer from acyl-enzyme intermediate to the final acyl acceptor, a proHlyA derivative. Overall the acylation of proHlyA-derived proteins catalyzed by HlyC was driven by the energetics of the acyl enzyme intermediate reaction. Of the two acylation sites, intactness of the site equivalent to proHlyA K564 was more important for acylation reaction thermodynamic stability.  相似文献   

14.
Z Y Zhang  R L Van Etten 《Biochemistry》1991,30(37):8954-8959
The kcat and Km values for the bovine heart low molecular weight phosphotyrosyl protein phosphatase catalyzed hydrolysis of 16 aryl phosphate monoesters and of five alkyl phosphate monoesters having the structure Ar(CH2)nOPO3H2 (n = 1-5) were measured at pH 5.0 and 37 degrees C. With the exception of alpha-naphthyl phosphate and 2-chlorophenyl phosphate, which are subject to steric effects, the values of kcat are effectively constant for the aryl phosphate monoesters. This is consistent with the catalysis being nucleophilic in nature, with the existence of a common covalent phosphoenzyme intermediate, and with the breakdown of this intermediate being rate-limiting. In contrast, kcat for the alkyl phosphate monoesters is much smaller and the rate-limiting step for these substrates is interpreted to be the phosphorylation of the enzyme. A single linear correlation is observed for a plot of log (kcat/Km) vs leaving group pKa for both classes of substrates at pH 5.0: log (kcat/Km) = -0.28pKa + 6.88 (n = 19, r = 0.89), indicating a uniform catalytic mechanism for the phosphorylation event. The small change in effective charge (-0.28) on the departing oxygen of the substrate is similar to that observed in the specific acid catalyzed hydrolysis of monophosphate monoanions (-0.27) and is consistent with a strong electrophilic interaction of the enzyme with this oxygen atom in the transition state. The D2O solvent isotope effect and proton inventory experiments indicate that only one proton is "in flight" in the transition state of the phosphorylation process and that this proton transfer is responsible for the reduction of effective charge on the leaving oxygen.  相似文献   

15.
Pseudomonas sp. lipase-catalyzed enantioselective acylation procedure using acid anhydrides as acyl donors was exploited for the resolution of 2-aryloxy-1-propanols carrying different substituents on the benzene ring. These primary alcohols, which belong to primary alcohols with an oxygen atom at the stereocenter, were resolved generally with moderate to good enantioselectivity (E of up to 55) through the acylation with hexanoic anhydride in diisopropyl ether at 25 °C in a short reaction time. With the alcohol substrate, which gave a low enantioselectivity in the acylation at ordinary temperature, the selectivity proved to be enhanced by conducting the reaction at low temperature (−10 °C). By this acylation procedure employing the acid anhydride, enantiomerically pure (R)-2-phenoxy-1-propanol was prepared in a gram-scale reaction.  相似文献   

16.
Analysis of nucleophile separation via lipase catalyzed reactions has been developed on the basis of competitive enzymatic kinetics. Ester synthesis as well as ester interchange reactions catalyzed by lipases in organic media have been analyzed according to a transfer reaction of the acyl group from/to the enzyme. The reversible reactions are conveniently simulated from the knowledge of the a competitive factor of the enzymatic system and of the final equilibrium conditions. The model which is proposed describes the reaction profile in a predictive way. Modelling of alcohol kinetic separation and resolution is given.  相似文献   

17.
The reaction pathways of several Friedel–Crafts acylations involving phenyl aromatic compounds were studied using density functional theory. The reactions were related to the Friedel–Crafts polycondensation of polyaryletherketones. In particular, the acylation of benzene with benzoyl chloride to form benzophenone and variations on this reaction were investigated. The acylation of benzene by one molecule of terephthaloyl chloride or isophthaloyl chloride as well as acylations at the m-, o-, and p-positions of diphenyl ether with one molecule of benzoyl chloride were studied. Adding an additional acyl chloride group to the electrophile appeared to have little influence on the reaction pathway, although the activation energy for the C–C bond-forming steps that occurred when isophthaloyl choride was used was different to the activation energy observed when terephthaloyl chloride was used. Upon changing the nucleophile to diphenyl ether, the reactivity changed according to the trend predicted on based on the o-, p-directing effects of the ether group. The deprotonation step that restored aromaticity varied widely according to the reaction. The rate-determining step in all of the studied reactions was the formation of the acylium ion, followed in importance by either the formation of the Wheland intermediate or the abstraction of hydrogen, depending on the reactivity of the nucleophile.  相似文献   

18.
Krogh BO  Shuman S 《Biochemistry》2000,39(21):6422-6432
Vaccinia topoisomerase forms a covalent DNA-(3'-phosphotyrosyl)-enzyme intermediate at sites containing the sequence 5'-CCCTT downward arrow. The covalently bound topoisomerase can religate the CCCTT strand to a 5'-OH-terminated polynucleotide or else transfer the strand to a non-DNA nucleophile such a water or glycerol. Here, we report that vaccinia topoisomerase also catalyzes strand transfer to hydrogen peroxide. The observed alkaline pH-dependence of peroxidolysis is consistent with enzyme-mediated attack by peroxide anion on the covalent intermediate. The reaction displays apparent first-order kinetics. From a double-reciprocal plot of k(obs) versus [H(2)O(2)] at pH 10, we determined a rate constant for peroxidolysis of 6.3 x 10(-)(3) s(-)(1). This rate is slower by a factor of 200 than the rate of topoisomerase-catalyzed strand transfer to a perfectly aligned 5'-OH DNA strand but is comparable to the rate of DNA strand transfer across a 1-nucleotide gap. Strand transfer to 2% hydrogen peroxide is 300 times faster than strand transfer to 20% glycerol and approximately 2000 times faster than topoisomerase-catalyzed hydrolysis of the covalent intermediate. Hydroxylamine is also an effective nucleophile in topoisomerase-mediated strand transfer (k(obs) = 6.4 x 10(-)(4) s(-)(1)). The rates of the peroxidolysis, hydroxylaminolysis, glycerololysis, and hydrolysis reactions catalyzed by the mutant enzyme H265A were reduced by factors of 100-700, in accordance with the 100- to 400-fold rate decrements in DNA cleavage and religation by H265A. We surmise that vaccinia topoisomerase catalyzes strand transfer to DNA and non-DNA nucleophiles via a common reaction pathway in which His-265 stabilizes the scissile phosphate in the transition state rather than acting as a general acid or base.  相似文献   

19.
The quantitative analysis of peptide synthesis via transfer of the acyl moiety from the activated donor (S) to the nucleophile (N), catalysed by proteases forming an acylenzyme intermediate, has been continued. The new kinetic model takes into account the hydrolysis of an acylenzyme-nucleophile complex (EAN). The intensity of the hydrolysis is characterized by parameter gamma equal to the ratio of the rate constant of EAN hydrolysis and the rate constant of peptide formation. The ability of the EAN complex to hydrolyse leads to a decrease in the apparent nucleophile reactivity (beta) of the aminocomponent. As a result, the maximal fractional conversions of S and N to the peptide decrease, and the apparent nucleophile reactivity becomes dependent on the nucleophile concentration. The pattern of parameter gamma influence on maximal fractional conversions depends on which component is in an excess. It is with the donor excess that hydrolysis of the EAN complex affects the peptide yield dramatically. Analytical expressions for the estimation of maximal product concentration were obtained and their accuracy evaluated.  相似文献   

20.
M W Washabaugh  W P Jencks 《Biochemistry》1988,27(14):5044-5053
Rate constants for C(2)-proton exchange from thiamin, N(1')-methylthiamin, and several 3-substituted-4-methylthiazolium ions catalyzed by D2O and deuterioxide ion were determined by 1H NMR at 30 degrees C and ionic strength 2.0 M. Values of pKa for the thiazolium ions, including thiamin itself, were found to be in the range pKa = 17-19; the pKa values for N(1')-protonated thiamin and free thiamin C(2)-H in H2O are 17.7 and 18.0, respectively. The pKa value for N(1')-protonated thiamin was calculated from the observed rate constant for the pD-independent reaction with D2O after correction for a secondary solvent deuterium isotope effect of kH2O/kD2O = 2.6. The pKa value for free thiamin was calculated from the rate constant for catalysis by OD- after correction by a factor of 3.3 = 8/2.4 for an 8-fold negative deviation of kOD from the Br?nsted plot of slope 1.0 for general base catalysis and a secondary solvent isotope effect of kOD/kOH = 2.4. Values of k-a = 2 X 10(10) and 3 X 10(9) M-1 s-1 were assumed for diffusion-controlled protonation of the C(2) ylide in the reverse direction by H3O+ and H2O, respectively. The Hammett rho I value for the exchange reaction catalyzed by deuterioxide ion or D2O is 8.4 +/- 0.2. There is no positive deviation of the rate constants for free or N(1')-substituted thiamin analogues in either Hammett correlation. This shows that the aminopyrimidinyl group does not provide significant intramolecular catalysis of nonenzymic C(2)-proton removal in the coenzyme.  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