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 peptide synthesis from N-acetyl-L-tyrosine ethyl ester and amino acid amides was realized using α-chymotrypsin as a catalyst in ethanol or acetonitrile containing small amounts of water. In these reaction systems, the precipitates of phosphate salt, which was used as a component of buffer solution, are considered to act as carriers of chymotrypsin. It was found that peptide formation is competitive with hydrolysis of the substrate ester, but the secondary synthesis of the peptide from the hydrolysate was also considered to proceed. The yield of the peptide after 24 h reaction was strongly dependent on the water concentration; maximum yields of the peptide were obtained at water concentrations below 10% (v/v). The addition of tertiary amines, such as triethyl amine, markedly increased the peptide yield, probably due to the increase in the concentration of the nucleophilic amine components by neutralization of hydrohalides of amino acid amides. The effect of reaction temperature and the reactions with CT immobilized on PVA, chitosan, or TEAE-cellulose are also described.  相似文献   

2.
To differentiate esterases from lipases at the structure–function level, we have compared the kinetic properties and structural features of sequence-related esterase 1 from rabbit liver (rLE) and bile-salt-activated lipase from bovine pancreas (bBAL). In contrast to rLE, bBAL hydrolyses water-insoluble medium and long chain esters as vinyl laurate, trioctanoin and olive oil. Conversely, rLE and bBAL are both active on water-soluble short chain esters as vinyl acetate, vinyl propionate, vinyl butyrate, tripropionin, tributyrin and p-nitrophenyl butyrate. However, the enzymes show distinctive kinetic behaviours. rLE displays maximal activity at low substrate concentration, below the critical micelle concentration, whereas bBAL acts preferencially on emulsified esters, at concentration exceeding the solubility limit. Comparison of the 3D structures of rLE and bBAL shows, in particular, that the peptide loop at positions 116–123 in bBAL is deleted in rLE. This peptide segment interacts with a bile salt molecule thus inducing a conformational transition which gives access to the active site. Inhibition studies and manual docking of a bulky ester molecule as vinyl laurate in the catalytic pocket of rLE and bBAL show that the inability of the esterase to hydrolyse large water-insoluble esters is not due to steric hindrance. It is hypothesized that esterases lack specific hydrophobic structures involved both in the stabilization of the lipase–lipid adsorption complex at interfaces and in the spontaneous transfer of a single substrate molecule from interface to the catalytic site.  相似文献   

3.
In recent years, the reductive amination of ketones in the presence of amine dehydrogenases emerged as an attractive synthetic strategy for the enantioselective preparation of amines starting from ketones, an ammonia source, a reducing reagent and a cofactor, which is recycled in situ by means of a second enzyme. Current challenges in this field consists of providing a broad synthetic platform as well as process development including enzyme immobilization. In this contribution these issues are addressed. Utilizing the amine dehydrogenase EsLeuDH-DM as a mutant of the leucine dehydrogenase from Exigobacterium sibiricum, a range of aryl-substituted ketones were tested as substrates revealing a broad substrate tolerance. Kinetics as well as inhibition effects were also studied and the suitability of this method for synthetic purpose was demonstrated with acetophenone as a model substrate. Even at an elevated substrate concentration of 50?mM, excellent conversion was achieved. In addition, the impact of water-miscible co-solvents was examined, and good activities were found when using DMSO of up to 30% (v/v). Furthermore, a successful immobilization of the EsLeuDH-DM was demonstrated utilizing a hydrophobic support and a support for covalent binding, respectively, as a carrier.  相似文献   

4.
Various structural features of polyamines which are responsible for the acceleration of the hydrolysis of ATP to ADP and Pi at pH 3–4 were surveyed by means of kinetic studies, leading to the following conclusions: 1) The ethyleneimine chain of the polyamines should be as long as possible; 2) the number of methylene carbon atoms between the two adjacent nitrogen atoms of the polyamine has to be two; 3) the terminal groups of the ethyleneimine chain should be primary amino groups.The rate of ATP hydrolysis in the presence of pentaethylenehexamine (pentaen), which possesses the above properties, was found to be 15 times as high as that of hydrolysis at pH 3.5 in the absence of amines. The kinetic data support the previous assumption that there is formation of an ATP-pentaen complex in the hydrolysis reaction. The formation constant of the complex has been calculated to be K = 1.9 × 104M−1 at pH 3.5 and 50°C from the kinetic data. From the temperature dependence of the rates for pentaen or tetraethylenepentamine, the thermodynamic data of these reactions have been obtained.On the other hand, it has been found that pentaen enhances the hydrolysis of GTP and UTP as well as ATP. No phosphate ester bonds of AMP, p-nitrophenylphosphate and α-d-glucose-1-phosphate were hydrolyzed. Therefore, it may be concluded that hydrolysis of the phosphate ester bond with the polyamine is characteristic of ATP, GTP and UTP.  相似文献   

5.
The cleavage of β-cyclodextrin trans-cinnamate (1) was accelerated by amines such as quinuclidine and piperidine by 27- and 13-fold, respectively. The reaction involves complex formation of 1 with the amines, and proceeds via nucleophilic attack by the neutral amine, which was shown by the production of amide in the reaction of 1 with piperidine. Quinuclidine exhibited real catalysis of hydrolysis without production of amide. The present finding indicates that the rates of the rate-determining deacylation step in the cyclodextrin-accelerated hydrolyses of phenyl esters can be made larger than the rates of uncatalyzed hydrolyses by an amine such as quinuclidine, resulting in the use of cyclodextrin as a true catalyst and as a better enzyme model.  相似文献   

6.
During complex formation between α2-macroglobulin and trypsin the internal thiol esters (one in each of the four Mr 180,000 subunits) are activated. In this activated state (nascent α2-macroglobulin-trypsin complex) addition of low Mr amines lead to their covalent incorporation into α2M. Evidence is presented showing that covalent binding of added amines occurs at the γ-carbonyl group of the Glx-residue in the thiol ester sequence:
  相似文献   

7.
The kinetics of the enzymatic transesterification between a mixture of triglycerides (oils) and methanol for biodiesel production in a bis(2-ethylhexyl) sodium sulfosuccinate (AOT)/isooctane reversed micellar system, using recombinant cutinase from Fusarium solani pisi as a catalyst, was investigated. In order to describe the results that were obtained, a mechanistic scheme was proposed, based on the literature and on the experimental data. This scheme includes the following reaction steps: the formation of the active enzyme–substrate complex, the addition of an alcohol molecule to the complex followed by the separation of a molecule of the fatty acid alkyl ester and a glycerol moiety, and release of the active enzyme. Enzyme inhibition and deactivation effects due to methanol and glycerol were incorporated in the model. This kinetic model was fitted to the concentration profiles of the fatty acid methyl esters (the components of biodiesel), tri-, di- and monoglycerides, obtained for a 24 h transesterification reaction performed in a stirred batch reactor under different reaction conditions of enzyme and initial substrates concentration.  相似文献   

8.
Beef liver esterase. II. Kinetic properties   总被引:1,自引:0,他引:1  
The kinetic parameters, kcat and KM, in beef liver esterase-catalyzed hydrolysis were determined for about 100 substrates, which can be classified in several groups: (1) In the ethyl ester series of fatty acids KM decreases with elongation of the acid, while kcat has a maximum value with pentanoate. (2) Alkyl acetates are better substrates as the alkyl moiety is longer, whereas esters with branched alkyl groups become worse substrates. (3) Aryl esters are very good substrates. (4) Esters of dicarboxylic acids are good substrates, but only one ester group is cleaved by the enzyme. Fumarate diester is susceptible to esterase hydrolysis, while maleate is not. (5) Esters of hydrophobic amino acids are very good substrates; the enzyme is not stereoselective and both the l and d stereoisomers are readily hydrolyzed. Branching at the β-carbon atom leads to loss of activity, and blocking of the amino group abolishes it. Fluoride ion and dl-malate esters are potent competitive inhibitors of the enzymic reaction. The optimal pH was found to lie between 8 and 8.5. The reaction rate increased between 5 and 40 °C then dropped sharply. The activity decreased at high salt concentration.  相似文献   

9.
Reactivities of several amine derivatives with the Folin and Lowry reagents were examined. Tertiary amines reacted with the Folin reagent to produce a blue color, and secondary amines having a 2-hydroxyethyl group reacted with the Folin reagent only in the presence of Cu2+, i.e., with the Lowry reagent. On the other hand, primary and quarternary amines and amine N-oxides produced no color with either reagent. Reactivities of tertiary amines were greatly influenced by the nature of the N-substituted groups, and the color yield of those forming stable chelate complexes with metals was strongly inhibited by the presence of Cu2+, indicating that the formation of a stable complex with Cu2+ reduces the reactivity of tertiary amino nitrogen. The requirement of Cu2+ for the color development with secondary amines having a 2-hydroxyethyl group may be due to the formation of weak chelate complex with Cu2+.  相似文献   

10.
We have studied the enzymatic hydrolysis of solutions and emulsions of vinyl propionate, vinyl butyrate and tripropionin by lipases of various origin and specificity. Kinetic studies of the hydrolysis of short-chain substrates by microbial triacylglycerol lipases from Rhizopus oryzae, Mucor miehei, Candida rugosa, Candida antarctica A and by (phospho)lipase from guinea-pig pancreas show that these lipolytic enzymes follow the Michaelis–Menten model. Surprisingly, the activity against solutions of tripropionin and vinyl esters ranges from 70% to 90% of that determined against emulsions. In contrast, a non-hyperbolic (sigmoidal) dependence of enzyme activity on ester concentration is found with human pancreatic lipase, triacylglycerol lipase from Humicola lanuginosa (Thermomyces lanuginosa) and partial acylglycerol lipase from Penicillium camembertii and the same substrates. In all cases, no abrupt jump in activity (interfacial activation) is observed at substrate concentration corresponding to the solubility limit of the esters. Maximal lipolytic activity is always obtained in the presence of emulsified ester. Despite progress in the understanding of structure–function of lipases, interpretation of the mode of action of lipases active against solutions of short-chain substrates remains difficult. Actually, it is not known whether these enzymes, which possess a lid structure, are in open or/and closed conformation in the bulk phase and whether the opening of the lid that gives access to the catalytic triad is triggered by interaction of the enzyme molecule with monomeric substrates or/and multimolecular aggregates (micelles) both present in the bulk phase. From the comparison of the behaviour of lipases used in this study which, in some cases, follow the Michaelis–Menten model and, in others, deviate from classical kinetics, it appears that the activity of classical lipases against soluble short-chain vinyl esters and tripropionin depends not only on specific interaction with single substrate molecules at the catalytic site of the enzyme but also on physico-chemical parameters related to the state of association of the substrate dispersed in the aqueous phase. It is assumed that the interaction of lipase with soluble multimolecular aggregates of tripropionin or short-chain vinyl esters or the formation of enzyme–substrate mixed micelles with ester bound to lipase, might represent a crucial step that triggers the structural transition to the open enzyme conformation by displacement of the lid.  相似文献   

11.
Guinea pig liver transglutaminase (TGase) reacts with 0.1 mM N-Cbz-L-Glu(gamma-p-nitrophenyl ester)Gly (5, prepared herein, K(M) = 0.02 mM) to undergo rapid acylation that can be followed spectrophotometrically at 400 nm (pH 7.0, 25 degrees C). Deacylation of the transiently formed thiolester acyl enzyme intermediate via catalytic aminolysis was studied in the presence of six primary amines of widely varying basicity (pK(NH+) = 5.6-10.5). Steady-state kinetic studies were performed to measure k(cat) and K(M) values for each amine substrate. A Br?nsted plot constructed through the correlation of log(k(cat)/K(M)) and pK(NH+) for each amine substrate displays a linear free-energy relationship with a slope beta(nuc) = -0.37 +/- 0.08. The shallow negative slope is consistent with a general-base-catalyzed deacylation mechanism in which a proton is removed from the amine substrate during its rate-limiting nucleophilic attack on the thiolester carbonyl. Kinetic isotope effects were measured for four acceptor substrates (water, kie = 1.1 +/- 0.1; aminoacetonitrile, kie = 5.9 +/- 1.2; glycine methyl ester, kie = 3.4 +/- 0.7; N-Ac-L-lysine methyl ester, kie = 1.1 +/- 0.1) and are consistent with a proton in flight at the rate-limiting transition state. The active site general-base implicated by these kinetic results is believed to be His-334, of the highly conserved TGase Cys-His-Asp catalytic triad.  相似文献   

12.
Conjugate addition of n-butyl amine to d-glucose derived alpha,beta-unsaturated ester 4 afforded beta-amino esters 5a,b that on reduction of ester group, 1,2-acetonide deprotection, and reductive amination led to the formation of corresponding N-butyl 1-deoxy-D-gluco-homonojirimycin 2c and N-butyl 1-deoxy-L-ido-homonojirimycin 2d which were found to be selective beta-glucosidase inhibitors with an IC(50) value in millimolar range.  相似文献   

13.
Acetyl esterases from carbohydrate esterase family 7 exhibit unusual substrate specificity. These proteins catalyze the cleavage of disparate acetate esters with high efficiency, but are unreactive to larger acyl groups. The structural basis for this distinct selectivity profile is unknown. Here, we investigate a thermostable acetyl esterase (TM0077) from Thermotoga maritima using evolutionary relationships, structural information, fluorescent kinetic measurements, and site directed mutagenesis. We measured the kinetic and structural determinants for this specificity using a diverse series of small molecule enzyme substrates, including novel fluorogenic esters. These experiments identified two hydrophobic plasticity residues (Pro228, and Ile276) surrounding the nucleophilic serine that impart this specificity of TM0077 for small, straight-chain esters. Substitution of these residues with alanine imparts broader specificity to TM0077 for the hydrolysis of longer and bulkier esters. Our results suggest the specificity of acetyl esterases have been finely tuned by evolution to catalyze the removal of acetate groups from diverse substrates, but can be modified by focused amino acid substitutions to yield enzymes capable of cleaving larger ester functionalities.  相似文献   

14.
Reactive extraction of succinic acid from aqueous solutions with various tertiary amines dissolved in 1-octanol and n-heptane has been studied as a function of the acid concentration and the chain length of tertiary amine. When 1-octanol was used as diluent, the extractability of the tertiary amine was proportional to the chain length of tertiary amine at the same concentration of amines. The loading values were also proportional to the chain length of tertiary amine. However, when n-heptane was used as diluent, the extractability and loading values of the tertiary amine were inversely proportional to the chain length of amines. It was due to the aggregate formation of the acid-amine complex.  相似文献   

15.
T A Alston  R H Abeles 《Biochemistry》1987,26(13):4082-4085
L-Histidine methyl ester inactivates histidine decarboxylase in a time-dependent manner. The possibility was considered that an irreversible reaction between enzyme and inhibitor occurs [Recsei, P. A., & Snell, E. E. (1970) Biochemistry 9, 1492-1497]. We have confirmed time-dependent inactivation by histidine methyl ester and have investigated the structure of the enzyme-inhibitor complex. Upon exposure to either 8 M guanidinium chloride or 6% trichloroacetic acid, unchanged histidine methyl ester is recovered. Formation of the complex involves Schiff base formation, most likely with the active site pyruvyl residue [Huynh, Q. K., & Snell, E. E. (1986) J. Biol. Chem. 261, 4389-4394], but does not involve additional irreversible covalent interaction between inhibitor and enzyme. Complex formation is a two-step process involving rapidly reversible formation of a loose complex and essentially irreversible formation of a tight complex. For the formation of the tight complex, Ki = 80 nM and koff = 2.5 X 10(-4) min-1. Time-dependent inhibition was also observed with L-histidine ethyl ester, L-histidinamide, and DL-3-amino-4-(4-imidazolyl)-2-butanone. No inactivation was observed with glycine methyl ester or histamine. We propose that in the catalytic reaction the carboxyl group of the substrate is in a hydrophobic region. The unfavorable interaction between the carboxylate group and the hydrophobic region facilitates decarboxylation [Crosby, J., Stone, R., & Liehard, G. E. (1970) J. Am. Chem. Soc. 92, 2891-2900]. With histidine methyl ester this unfavorable interaction is no longer present; hence, there is tight binding.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
Apparent kinetic constants (Km and Vmax values) were determined for human liver acyl-CoA: glycine acyltransferase (glycine-N-acylase) towards isobutyryl-CoA, 2-methyl butyryl-CoA, isovaleryl-CoA, butyryl-CoA, hexanoyl-CoA, octanoyl-CoA, and decanoyl-CoA. These acyl-CoA esters were selected because of their relevance to the human diseases with cellular accumulation of these esters, i.e., especially to metabolic defects in the acyl-CoA dehydrogenation steps of the branched-chain amino acids, lysine, 5-hydroxy lysine, tryptophan, and fatty acid oxidation pathways. With the acyl-CoA ester as the fixed substrate, the Km value for glycine ranged from 0.5 to 2.9 mole/liter, and with glycine as fixed substrate, the Km values for the acyl-CoA esters varied from 0.3 to 5.6 mmole/liter. It is concluded that the substrate concentration is decisive for the glycine conjugate formation and that the occurrence in urine of acylglycines reflects an intramitochondrial accumulation of the corresponding acyl-CoA ester.  相似文献   

17.
Sedimentation equilibrium studies on arylsulfate sulfohydrolase II (EC 3.1.6.1) from Aspergillus oryzae under nondissociating conditions have resulted in a revised molecular weight of 94,900 ± 7100. Sedimentation equilibrium and gel electrophoresis data collected in the presence of the dissociating agents, urea and sodium dodecylsulfate demonstrate that the native enzyme is composed of two identical subunits as suggested by previous studies employing an irreversible inhibitor.The pH dependencies of the kinetic parameters V and VKm for the enzymic hydrolysis of 4-nitrophenyl sulfate indicate that two groups of pKa 4.7 and 6.0 control the activity of the enzyme. The product inorganic sulfate was shown to be a linear competitive inhibitor of the enzyme at pH 4.0, implying that it is a last released product along the reaction pathway. Inhibition by the phenol product was not observed. Enzymic hydrolysis of 4-nitrophenyl sulfate in 18O enriched water revealed that one atom of solvent oxygen is incorporated per molecule of inorganic sulfate, which is consistent with a mechanism featuring sulfur-oxygen bond cleavage. Evidence is presented based on stopped-flow kinetics, partitioning experiments in the presence of amine nucleophiles, and 18O exchange studies that collectively suggest that the breakdown of a covalent sulfuryl enzyme intermediate probably is not the rate-limiting step along the reaction pathway.The substrate specificity of the enzyme was examined by testing a variety of sulfate and phosphate esters as inhibitors of the hydrolysis of 4-nitrophenyl sulfate. The Cbz-l-Phe-l-Tyrosine-O-sulfate methyl ester serves as a substrate for the enzyme. Apparently substrate activity requires an aromatic sulfate ester whose binding is enhanced by incorporating the aromatic moiety in a hydrophobic matrix.  相似文献   

18.
Electron spin resonance spectroscopy (ESR) was used to study the effects of amphiphilic amines of the carbamate, amide, and ester type and amine oxide on the photosynthetic system of spinach chloroplasts. The ESR signal II connected to the photosynthetic center PS II donor side was observed to diminish in the presence of amines, whereas that of PS I remained unchanged. The inhibition of PS II increased with the increasing of amine concentration. In the presence of amines, the light: dark chloroplast ESR signals ratio as well as the intensity of the ESR signal of unbound Mn2+ increased. It is suggested that the amphiphilic amines affect the structure of PS II and the electron transfer to PS I. The effects of the amines tested on the photosynthetic system correlate with their potency to perturb the lipid membrane structure.  相似文献   

19.
Quinohemoprotein amine dehydrogenase (QHNDH) possesses a cysteine tryptophylquinone (CTQ) prosthetic group that catalyzes the oxidative deamination of primary amines. In addition to CTQ, two heme c cofactors are present in QHNDH that mediate the transfer of the substrate-derived electrons from CTQ to an external electron acceptor. Steady-state kinetic assays yielded relatively small k(cat) values (<6 s(-1)), and the rate-limiting step appears to be the interprotein electron transfer from heme in QHNDH to the external electron acceptor. Transient kinetic studies of the CTQ-dependent reduction of heme in QHNDH by amine substrates yielded different rate constants for different substrates (72, 190, and 162 s(-1) for methylamine, butylamine, and benzylamine, respectively). Deuterium kinetic isotope effect (KIE) values of 5.3, 3.9, and 8.5 were observed, respectively, for the reactions of methylamine, butylamine, and benzylamine. These results suggest that the abstraction of a proton from the alpha-methylene group of the substrate, which occurs concomitant with CTQ reduction, is the rate-limiting step in the CTQ-dependent reduction of hemes in QHNDH by these amine substrates. In contrast, the reaction of 2-phenylethylamine with QHNDH does not exhibit a significant KIE ((H)k(3)/(D)k(3) = 1.05) and exhibits a much smaller rate constant of 16 s(-1). This suggests that for 2-phenylethylamine, the rate-limiting step in the single-turnover reaction is either hydrolysis of the imine reaction intermediate from CTQ or product release prior to intraprotein electron transfer. Analysis of the products of the reactions of QHNDH with chiral deuterated 2-phenylethylamines demonstrated that the enzyme abstracts the pro-S proton of the substrate in a highly stereospecific manner. Inspection of the crystal structure of phenylhydrazine-inhibited QHNDH suggests that Asp33(gamma) is the residue that performs the proton abstraction. On the basis of these results, kinetic and chemical reaction mechanisms for QHNDH are proposed and discussed in the context of the crystal structure of the enzyme.  相似文献   

20.
The effect of solvent on stereoselectivity in the nucleophilic addition reaction of various optically active amines to N-methylphenylalanine N-carboxyanhydride has been investigated. In m-dimethoxybenzene as solvent, (S)-valine, (S)-leucine, and (S)-phenylalanine ethyl esters reacted preferentially with (R)-N-methylphenylalanine N-carboxyanhydride, but the stereoselectivity decreased considerably in nitrobenzene and dimethylacetamide as solvents. In the latter solvents, the dipolar interactions between an amine and an N-carboxyanhydride and the orientation of the substituent of N-carboxyanhydride were seriously affected, hence the stereoselectivity decreased. As a consequence, the enantiomer selection by the terminal amine of a growing chain in the nucleophilic addition-type polymerization of α-amino acid N-carboxyanhydride can be controlled by the choice of solvent. (S)-Proline ethyl ester and (S)-α-phenylethylamine reacted preferentially with (S)-N-methylphenylalanine N-carboxyanhydride in m-dimethoxybenzene, and this type of selectivity did not change in nitrobenzene. But in dimethylacetamide the stereoselectivity decreased. In the transition state of the reaction of these amines and N-methylphenylalanine N-carboxyanhydride dipolar interactions are operating, which should be destroyed by dimethylacetamide but may not be affected by nitrobenzene.  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