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 pH dependence for acylation of alpha-chymotrypsin by N-acetyltryptophan p-nitrophenyl-, p-nitrothiophenyl-, ethyl-, and thiolethyl esters has been studied by the stopped-flow technique. Values for the acylation rate constant, k2, and the binding constant, KS, were obtained by using measurements of phenolate release, for the p-nitrophenyl esters, and proflavin displacement, for the ethyl esters. The oxygen esters tested have slightly higher k2 values, and substantially higher KS values relative to the analogous thiol esters. Whereas k2/KS for the thiolethyl ester is higher than that for the analogous oxygen ester, the k2/KS values for oxy- and thio-p-nitrophenyl esters are nearly identical. These data are interpreted to indicate rate-determining formation of a tetrahedral intermediate in acylation of alpha-chymotrypsin by p-nitrophenyl esters, and rate-determining breakdown of such an intermediate in the case of the ethyl esters. It is also concluded that the oxygen to sulfur substitution causes a substantial increase in the proportion of nonproductive binding in these substrates. pH dependent k2 and KS values were used to calculate values for k1 and k-1, the binding and debinding rate constants for the two p-nitrophenyl compounds. This is the first such calculation based on experimentally determined acylation rate constants.  相似文献   

2.
The hydrolysis of N-acetyl-L-methionine, N-acetylglycine, N-acetyl-L-phenylalanine, and N-acetyl-L-alanine at 298.35K by porcine kidney acylase I (EC 3.5.1.14) was monitored by the heat released upon mixing of the substrate and enzyme in a differential stopped flow microcalorimeter. Values for the Michaelis constant (K(m)) and the catalytic constant (k(cat)) were determined from the progress of the reaction curve employing the integrated form of the Michaelis-Menten equation for each reaction mixture. When neglecting acetate product inhibition of the acylase, values for k(cat) were up to a factor of 2.3 larger than those values determined from reciprocal initial velocity-initial substrate concentration plots for at least four different reaction mixtures. In addition, values for K(m) were observed to increase linearly with an increase in the initial substrate concentration. When an acetate product inhibition constant of 600+/-31M(-1), determined by isothermal titration calorimetry, was used in the progress curve analysis, values for K(m) and k(cat) were in closer agreement with their values determined from the reciprocal initial velocity versus initial substrate concentration plots. The reaction enthalpies, Delta(r)H(cal), which were determined from the integrated heat pulse per amount of substrate in the reaction mixture, ranged from -4.69+/-0.09kJmol(-1) for N-acetyl-L-phenylalanine to -1.87+/-0.23kJmol(-1) for N-acetyl-L-methionine.  相似文献   

3.
Laidler (1955) showed that consideration of the effect of pH on enzymic mechanisms that obey steady-state kinetics leads to the inclusion in the equations of a ;perturbation term' that can introduce curvature into the Lineweaver-Burk plots. He also stated conditions in which this term vanishes. This term can lead to apparent activation by substrate. Further, several cases are shown in which simplification, but not disappearance, of the perturbation term can lead to linearity of Lineweaver-Burk plots. These cases arise when the ionization of groups at the active site either is unaffected or is completely prevented when the enzyme-substrate complex is formed. It is also shown that V((app.)) can vary with pH without a concomitant change in K(m(app.)) in certain cases that obey steady-state kinetics without implying that K(m)=K(s). When the perturbation term is significant, Dixon's (1953) rules for the calculation of pK values will not always apply.  相似文献   

4.
1. N-Benzoyl-l-serine methyl ester was synthesized and evaluated as a substrate for bromelain (EC 3.4.22.4) and for papain (EC 3.4.22.2). 2. For the bromelain-catalysed hydrolysis at pH7.0, plots of [S(0)]/v(i) (initial substrate concn./initial velocity) versus [S(0)] are markedly curved, concave downwards. 3. Analysis by lattice nomography of a modifier kinetic mechanism in which the modifier is substrate reveals that concave-down [S(0)]/v(i) versus [S(0)] plots can arise when the ratio of the rate constants that characterize the breakdown of the binary (ES) and ternary (SES) complexes is either less than or greater than 1. In the latter case, there are severe restrictions on the values that may be taken by the ratio of the dissociation constants of the productive and non-productive binary complexes. 4. Concave-down [S(0)]/v(i) versus [S(0)] plots cannot arise from compulsory substrate activation. 5. Computational methods, based on function minimization, for determination of the apparent parameters that characterize a non-compulsory substrate-activated catalysis are described. 6. In an attempt to interpret the catalysis by bromelain of the hydrolysis of N-benzoyl-l-serine methyl ester in terms of substrate activation, the general substrate-activation model was simplified to one in which only one binary ES complex (that which gives rise directly to products) can form. 7. In terms of this model, the bromelain-catalysed hydrolysis of N-benzoyl-l-serine methyl ester at pH7.0, I=0.1 and 25 degrees C is characterized by K(m) (1) (the dissociation constant of ES)=1.22+/-0.73mm, k (the rate constant for the breakdown of ES to E+products, P)=1.57x10(-2)+/-0.32x10(-2)s(-1), K(a) (2) (the dissociation constant that characterizes the breakdown of SES to ES and S)=0.38+/-0.06m, and k' (the rate constant for the breakdown of SES to E+P+S)=0.45+/-0.04s(-1). 8. These parameters are compared with those in the literature that characterize the bromelain-catalysed hydrolysis of alpha-N-benzoyl-l-arginine ethyl ester and of alpha-N-benzoyl-l-arginine amide; K(m) (1) and k for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine amide hydrolysis and K(as) and k' for the serine ester hydrolysis are somewhat similar to K(m) and k(cat.) for the arginine ester hydrolysis. 9. A previous interpretation of the inter-relationships of the values of k(cat.) and K(m) for the bromelain-catalysed hydrolysis of the arginine ester and amide substrates is discussed critically and an alternative interpretation involving substantial non-productive binding of the arginine amide substrate to bromelain is suggested. 10. The parameters for the bromelain-catalysed hydrolysis of the serine ester substrate are tentatively interpreted in terms of non-productive binding in the binary complex and a decrease of this type of binding by ternary complex-formation. 11. The Michaelis parameters for the papain-catalysed hydrolysis of the serine ester substrate (K(m)=52+/-4mm, k(cat.)=2.80+/-0.1s(-1) at pH7.0, I=0.1, 25.0 degrees C) are similar to those for the papain-catalysed hydrolysis of methyl hippurate. 12. Urea and guanidine hydrochloride at concentrations of 1m have only small effects on the kinetic parameters for the hydrolysis of the serine ester substrate catalysed by bromelain and by papain.  相似文献   

5.
1. A procedure is described for determining the affinity constant K(a) and the phosphorylation constant k(p) for the inhibition by di-isopropyl phosphorofluoridate of erythrocyte acetylcholinesterase and serum cholinesterase. The procedure depends on the use of a specially designed reaction vessel with which incubation times as short as 1.2sec. could be obtained at any convenient temperature. 2. The K(a) of acetylcholinesterase decreased from 1.58 (+/-0.22)x10(-3)m at 5 degrees to 1.17 (+/-0.10)x10(-3)m at 25 degrees and the associated change in enthalpy was 2980 cal. 3. The k(p) of acetylcholinesterase increased from 11.9 (+/-0.7)min.(-1) at 5 degrees to 40.7 (+/-1.4)min.(-1) at 25 degrees , indicating an activational energy of 9600 cal. The change in entropy associated with K(a) was 23.5 cal. degree(-1) at 25 degrees . 4. At 5 degrees , the K(a) and k(p) of serum cholinesterase were 9.95 (+/-1.10)x10(-6)m and 11.2 (+/-0.63)min.(-1) respectively. 5. The 150-fold difference in the inhibitory power of di-isopropyl phosphorofluoridate for the two cholinesterases was attributed entirely to differences in affinity.  相似文献   

6.
Yang Z  Lu CD 《Journal of bacteriology》2007,189(11):3954-3959
The arginine transaminase (ATA) pathway represents one of the multiple pathways for L-arginine catabolism in Pseudomonas aeruginosa. The AruH protein was proposed to catalyze the first step in the ATA pathway, converting the substrates L-arginine and pyruvate into 2-ketoarginine and L-alanine. Here we report the initial biochemical characterization of this enzyme. The aruH gene was overexpressed in Escherichia coli, and its product was purified to homogeneity. High-performance liquid chromatography and mass spectrometry (MS) analyses were employed to detect the presence of the transamination products 2-ketoarginine and L-alanine, thus demonstrating the proposed biochemical reaction catalyzed by AruH. The enzymatic properties and kinetic parameters of dimeric recombinant AruH were determined by a coupled reaction with NAD(+) and L-alanine dehydrogenase. The optimal activity of AruH was found at pH 9.0, and it has a novel substrate specificity with an order of preference of Arg > Lys > Met > Leu > Orn > Gln. With L-arginine and pyruvate as the substrates, Lineweaver-Burk plots of the data revealed a series of parallel lines characteristic of a ping-pong kinetic mechanism with calculated V(max) and k(cat) values of 54.6 +/- 2.5 micrromol/min/mg and 38.6 +/- 1.8 s(-1). The apparent K(m) and catalytic efficiency (k(cat)/K(m)) were 1.6 +/- 0.1 mM and 24.1 mM(-1) s(-1) for pyruvate and 13.9 +/- 0.8 mM and 2.8 mM(-1) s(-1) for l-arginine. When L-lysine was used as the substrate, MS analysis suggested Delta(1)-piperideine-2-carboxylate as its transamination product. These results implied that AruH may have a broader physiological function in amino acid catabolism.  相似文献   

7.
The substrate specificity of honeybee alpha-glucosidase I, a monomeric enzyme was kinetically investigated. Unusual kinetic features were observed in the cleavage reactions of sucrose, maltose, p-nitrophenyl alpha-glucoside, phenyl alpha-glucoside, turanose, and maltodextrin (DP = 13). At relatively high substrate concentrations, the velocities of liberation of fructose from sucrose, glucose from maltose, p-nitrophenol from p-nitrophenyl alpha-glucoside, and phenol from phenyl alpha-glucoside were accelerated, and so the Lineweaver-Burk plots were convex, indicating negative kinetic cooperativity: the Hill coefficients were calculated to be 0.50, 0.64, 0.50, and 0.67 for sucrose, maltose, p-nitrophenyl alpha-glucoside, and phenyl alpha-glucoside, respectively. For the degradation of turanose and maltodextrin, the enzyme showed a sigmoidal curve in v versus s plots and thus catalyzed the reaction with positive kinetic cooperativity. The Lineweaver-Burk plots were concave and the Hill coefficients were 1.2 and 1.5 for turanose and maltodextrin, respectively. These unique properties cannot be interpreted by the reaction mechanism that Huber and Thompson proposed: (1973) Biochemistry 12, 4011-4020. The rate parameters for the hydrolysis of sucrose, maltose, p-nitrophenyl alpha-glucoside and phenyl alpha-glucoside were estimated by extrapolating the linear part of the Lineweaver-Burk plots at low substrate concentrations.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
The active site of the recombinant Talaromyces stipitatus type-C feruloyl esterase (TsFaeC) was probed using a series of C1-C4 alkyl ferulates and methyl esters of phenylalkanoic and cinnamic acids. The enzyme was active on 23 of the 34 substrates tested. Lengthening or shortening the aliphatic side chain while maintaining the same aromatic substitutions completely abolished the enzyme activity. Maintaining the phenylpropenoate structure but altering the substitutions of the aromatic ring demonstrated the importance of hydroxyl groups on meta and/or para position of the benzoic ring. The highest catalytic efficiency of TsFaeC for methyl cinnamates was shown on methyl 3,4-dihydroxy cinnamate and on its hydro form (3,4-dihydroxy-phenyl-propionate). Maintaining the ferulate structure but altering the esterified alkyl group, the comparison of k(cat) and k(cat)/K(m) values showed that the enzyme hydrolysed faster and more efficiently than ethyl ferulate. Alkyl ferulates were applied also for substrate selectivity mapping of feruloyl esterase to catalyze feruloyl group transfer to l-arabinose, using as a reaction system a ternary water-organic mixture consisting of n-hexane, t-butanol and water. The reaction parameters affecting the feruloylation rate and the conversion of the enzymatic synthesis, such as the composition of the reaction media, temperature, substrate and enzyme concentration have been investigated.  相似文献   

9.
1. The oxidation of NN-dimethyl-p-phenylenediamine (DPD) by inorganic oxidants and by caeruloplasmin was studied. Some experiments were also made with NNN'N'-tetramethyl-p-phenylenediamine (TPD). 2. E(mM) (550) of the first free radical oxidation product of DPD (DPD(+)) was 9.8 and E(mM) (563) of the corresponding product of TPD (TPD(+)) was 12.5. 3. The non-enzymic decomposition of DPD(+) was studied with respect to temperature, pH, concentration and DPD/DPD(+) ratio, thus defining conditions for enzyme experiments under which DPD(+) extinction at 550mmu was proportional to enzyme activity. 4. Rates of oxidation of DPD to DPD(+) by caeruloplasmin were constant over a range of DPD concentrations. At low DPD concentrations a lag period occurred, which was eliminated by addition of DPD(+). 5. A lag period was not observed with TPD, but at low TPD concentrations the rate of TPD(+) formation was greater when TPD(+) was added. This suggests that TPD(+) may compete weakly as a substrate with TPD and may be oxidized further by the enzyme before a non-enzymic reaction with TPD to form more TPD(+). 6. With DPD sulphate or acetate or TPD sulphate as substrate, Lineweaver-Burk plots were curved. With DPD hydrochloride the chloride ion caused inhibition at higher concentrations, opposing the curvature. 7. Curved Lineweaver-Burk plots were interpreted in terms of two types of substrate binding site with different K(m) values but similar V(max.) values. 8. The apparent thermodynamic changes associated with enzyme-substrate-complex formation at the sites with higher K(m) suggest that considerable conformational change may occur on binding at these sites. 9. With substrate concentrations at which only the low-K(m) sites are involved 2mol. of DPD(+)/mol. of caeruloplasmin are formed before a steady state is established. At higher substrate concentrations up to 3.2mol. of DPD(+)/mol. of caeruloplasmin are formed at this initial stage. 10. Results are discussed in relation to caeruloplasmin structures in which (a) two valence-changing and two permanently cuprous copper atoms are more accessible than the remaining four copper atoms or (b) binding of substrate at one site hinders access of substrate to another site.  相似文献   

10.
Specificity and kinetics of triose phosphate isomerase from chicken muscle   总被引:12,自引:6,他引:6  
The isolation of crystalline triose phosphate isomerase from chicken breast muscle is described. The values of k(cat.) and K(m) for the reaction in each direction were determined from experiments over wide substrate-concentration ranges, and the reactions were shown to obey simple Michaelis-Menten kinetics. With d-glyceraldehyde 3-phosphate as substrate, k(cat.) is 2.56x10(5)min(-1) and K(m) is 0.47mm; with dihydroxyacetone phosphate as substrate, k(cat.) is 2.59x10(4)min(-1) and K(m) is 0.97mm. The enzyme-catalysed exchange of the methyl hydrogen atoms of the ;virtual substrate' monohydroxyacetone phosphate with solvent (2)H(2)O or (3)H(2)O was shown. This exchange is about 10(4)-fold slower than the corresponding exchange of the C-3 hydrogen of dihydroxyacetone phosphate. The other deoxy substrate, 3-hydroxypropionaldehyde phosphate, was synthesized, but is too unstable in aqueous solution for analogous proton-exchange reactions to be studied.  相似文献   

11.
The kinetics of the cellulase-catalyzed conversion of soluble cellulose into glucose have been studied over a range of substrate concentrations and temperatures, and at pH values ranging from 4.75 to 7.0. Lineweaver-Burk plots were linear and led to V = 6.2muM/s and K(m) = 13.1 mM at pH 5.8 and 25.0 degrees C. The pK values corresponding to the free enzyme are 4.8 and 6.8 and are consistent with carboxyl and imidazole groups as the active ionizing species. These pK values were little changed in the enzyme-substrate intermediate that reacts in the ratedetermining step, suggesting that the ionizing groups are still free in this intermediate. The activation energy corresponding to V/K(m) is 80.6 kJ/mol, and that corresponding to V is 38.7 kJ/mol. The corresponding entropies of activation are 21 J K(-1) mol(-1) and -157 J K(-1) mol(-1), respectively.  相似文献   

12.
To elucidate the interaction between substrate inhibition and substrate transglycosylation of retaining glycoside hydrolases (GHs), a steady-state kinetic study was performed for the GH family 3 glucan (1-->3)-beta-glucosidase from the white-rot fungus Phanerochaete chrysosporium, using laminarioligosaccharides as substrates. When laminaribiose was incubated with the enzyme, a transglycosylation product was detected by thin-layer chromatography. The product was purified by size-exclusion chromatography, and was identified as a 6-O-glucosyl-laminaribiose (beta-D-Glcp-(1-->6)-beta-D-Glcp-(1-->3)-D-Glc) by 1H NMR spectroscopy and electrospray ionization mass spectrometry analysis. In steady-state kinetic studies, an apparent decrease of laminaribiose hydrolysis was observed at high concentrations of the substrate, and the plots of glucose production versus substrate concentration were thus fitted to a modified Michaelis-Menten equation including hydrolytic and transglycosylation parameters (K(m), K(m2), k(cat), k(cat2)). The rate of 6-O-glucosyl-laminaribiose production estimated by high-performance anion-exchange chromatography coincided with the theoretical rate calculated using these parameters, clearly indicating that substrate inhibition of this enzyme is fully explained by substrate transglycosylation. Moreover, when K(m), k(cat), and affinity for glucosyl-enzyme intermediates (K(m2)) were estimated for laminarioligosaccharides (DP=3-5), the K(m) value of laminaribiose was approximately 5-9 times higher than those of the other oligosaccharides (DP=3-5), whereas the K(m2) values were independent of the DP of the substrates. The kinetics of transglycosylation by the enzyme could be well interpreted in terms of the subsite affinities estimated from the hydrolytic parameters (K(m) and k(cat)), and a possible mechanism of transglycosylation is proposed.  相似文献   

13.
Case A  Huskey WP  Stein RL 《Biochemistry》2003,42(16):4727-4732
Investigating the selectivity that an enzyme expresses toward its substrates can be technically challenging if reaction of these substrates is not accompanied by a conveniently monitored change in some physicochemical property. In this paper, we describe a simple method for determining steady-state kinetic parameters for enzymatic turnover of such "silent" substrates. According to this method, silent substrate S is allowed to compete for enzymic reaction with signal-generating substrate S*, whose conversion to product can be conveniently monitored. Full reaction progress curves are collected under conditions of [S*](o) < K(m)* and [S](o) >or= 3K(m). Progress curves collected under these conditions are characterized by an initial lag phase of duration tau that is followed by the pseudo-first-order reaction of S. Steady-state kinetic parameters for the silent substrate can be obtained by one of two methods. One method combines least-squares fitting with numerical integration of appropriate rate equations to analyze the progress curves, while the other method relies on direct graphical analysis in which K(m) is the value of [S](o) that reduces the control velocity by a factor of 2 and V(max) is shown to simply equal the ratio [S](o)/tau. We use these methods to analyze the alpha-chymotrypsin-catalyzed hydrolysis of silent substrate Suc-Ala-Phe-AlaNH(2) with signal generator Suc-Ala-Phe-pNA. From the curve-fitting method, k(c) = 0.9 +/- 0.2 s(-1) and K(m) = 0.4 +/- 0.1 mM, while by direct graphical analysis, k(c) = 1.1 +/- 0.1 s(-1) and K(m) = 0.51 +/- 0.03 mM. As validation of this new method, we show agreement of these values with those determined independently by HPLC analysis of the hydrolysis of Suc-Ala-Phe-AlaNH(2) by alpha-CT, where k(c) = 1.1 +/- 0.1 s(-1) and K(m) = 0.5 +/- 0.1 mM.  相似文献   

14.
The effect of galactose on the inactivation of purified beta-galactosidase from the black bean, Kestingiella geocarpa, in 5 M urea at 50 degrees C and at pH 4.5, was determined. Lineweaver-Burk plots of initial velocity data in the presence and absence of urea and galactose were used to determine the relevant K(m) and V(max) values, with p-nitrophenyl beta-D-galactopyranoside (PNPG) as substrate, S. The inactivation data were analysed using the Tsou equation and plots. Plots of ln([P](infinity) - [P](t) ) against time in the presence of urea yielded the inactivation rate constant, A. Plots of A vs [S] at different galactose concentrations were zero order showing that A was independent of [S]. Plots of [P](infinity) vs [S] were used to determine the mode of inhibition of the enzyme by galactose, and slopes and intercepts of the 1/[P](infinity) vs. 1/[S] yielded k(+0) and k '(+0), the microscopic rate constants for the free enzyme and the enzyme-substrate complex, respectively. Plots of k(+0) and k '(+0) vs. galactose concentrations showed that galactose protected the free enzyme and not the enzyme-substrate complex against urea inactivation via a noncompetitive mechanism at low galactose concentrations and a competitive pattern of inhibition at high galactose concentrations. The implication of the different modes of inhibition in protecting the free enzyme was discussed.  相似文献   

15.
High-resolution crystallographic analysis of a complex of the serine-carboxyl proteinase sedolisin with pseudo-iodotyrostatin revealed two molecules of this inhibitor bound in the active site of the enzyme, marking subsites from S3 to S3('). The mode of binding represents two products of the proteolytic reaction. Substrate specificity of sedolisin was investigated using peptide libraries and a new peptide substrate for sedolisin, MCA-Lys-Pro-Pro-Leu-Glu#Tyr-Arg-Leu-Gly-Lys(DNP)-Gly, was synthesized based on the results of the enzymatic and crystallographic studies and was shown to be efficiently cleaved by the enzyme. The kinetic parameters for the substrate, measured by the increase in fluorescence upon relief of quenching, were: k(cat)=73+/-5 s(-1), K(m)=0.12+/-0.011 microM, and k(cat)/K(m)=608+/-85 s(-1)microM(-1).  相似文献   

16.
Cytosine deaminase (CDA) from Escherichia coli was shown to catalyze the deamination of isoguanine (2-oxoadenine) to xanthine. Isoguanine is an oxidation product of adenine in DNA that is mutagenic to the cell. The isoguanine deaminase activity in E. coli was partially purified by ammonium sulfate fractionation, gel filtration, and anion exchange chromatography. The active protein was identified by peptide mass fingerprint analysis as cytosine deaminase. The kinetic constants for the deamination of isoguanine at pH 7.7 are as follows: k(cat) = 49 s(-1), K(m) = 72 μM, and k(cat)/K(m) = 6.7 × 10(5) M(-1) s(-1). The kinetic constants for the deamination of cytosine are as follows: k(cat) = 45 s(-1), K(m) = 302 μM, and k(cat)/K(m) = 1.5 × 10(5) M(-1) s(-1). Under these reaction conditions, isoguanine is the better substrate for cytosine deaminase. The three-dimensional structure of CDA was determined with isoguanine in the active site.  相似文献   

17.
gamma-glutamyl Transferase fron Sheep brain cortex capillaries was studied from the point of view of transport of aminoacids across blood brain barrier. Excess substrate inhibition was competitive and observed both with donor (glutathione) and various acceptors (methionine, alanine, tryptophan) but not with arginine. Excess glutathione inhibition of transfer reaction is concomitant with an increase of total reaction (transfer + hydrolysis + autotranspeptidation). With regard to aminoacids, the greater the K'm the stronger the inhibition. This inhibition is the result of formation of a dead complex. Lineweaver-Burk plots 1/v versus 1/[acceptor] give straight lines meeting at the same point, whereas 1/v verus 1/[donor] plots are roughly parallel for high aminoacid concentrations and become secant for the low ones. Replots of slopes vs. 1/[acceptor] are not linear: the lower the aminoacid affinity the more pronounced the slope replot curvature. Thus kinetic patterns are consistent with a branched ping-pong mechanism including a ternary complex (Enzyme-acceptor-H2O) at high or low relative concentration, which balances the two branches. The estimated value of kinetic parameters does not support the hypothesis of major implication of the enzyme in brain uptake of aminoacids.  相似文献   

18.
The primary specificity residue of a substrate or an inhibitor, called the P(1) residue, is responsible for the proper recognition by the cognate enzyme. This residue enters the S(1) pocket of the enzyme and establishes contacts (up to 50%) inside the proteinase substrate cavity, strongly affecting its specificity. To analyze the influence on bovine alpha-chymotrypsin substrate activity, aromatic non-proteinogenic amino acid residues in position P(1) with the sequence Ac-Phe-Ala-Thr-X-Anb(5,2)-NH(2) were introduced: L-pyridyl alanine (Pal), 4-nitrophenylalanine - Phe(p-NO(2)), 4-aminophenylalanine - Phe(p-NH(2)), 4-carboxyphenylalanine Phe(p-COOH), 4-guanidine phenylalanine - Phe(p-guanidine), 4-methyloxycarbonyl-phenylalanine - Phe(p-COOMe), 4-cyanophenylalanine - Phe(p-CN), Phe, Tyr. The effect of the additional substituent at the phenyl ring of the Phe residue was investigated. All peptides contained an amide of 5-amino-2-nitrobenzoic acid, which served as a chromophore. Kinetic parameters (k(cat), K(M) and k(cat)/K(M)) of the peptides synthesized with bovine alpha-chymotrypsin were determined. The highest value of the specificity constant k(cat)/K(M), reaching 6.0 x 10(5) [M(-1)xs(-1)], was obtained for Ac-Phe-Ala-Thr-Phe(p-NO(2))-Anb(5,2)-NH(2). The replacement of the acetyl group with benzyloxycarbonyl moiety yielded a substrate with the value of k(cat) more than three times higher. Peptide aldehydes were synthesized with selected residues (Phe, Pal, Tyr, Phe(p-NO(2)) in position P(1) and potent chymotrypsin inhibitors were obtained. The dissociation constant (K(i)) with the experimental enzyme determined for the most active peptide, Tos-Phe-Ala-Thr-Phe(p-NO(2))-CHO, amounted to 1.12 x 10(-8) M.  相似文献   

19.
1. Lineweaver-Burk plots for glutamate dehydrogenase, glucose 6-phosphate dehydrogenase and several other enzymes show one or more abrupt transitions between apparently linear sections. These transitions correspond to abrupt increases in the apparent K(m) and V(max.) with increasing concentration of the varied substrate. 2. The generalized reciprocal initial-rate equation for a multi-site enzyme requires several restrictions to be put on it in order to generate such plots. These mathematical conditions are explored. 3. It is shown that the effective omission of a term in the denominator of the reciprocal initial-rate equation represents a minimal requirement for generation of abrupt transitions. This corresponds in physical terms to negative co-operativity followed by positive co-operativity affecting the catalytic rate constant for the reaction. 4. Previous models for glutamate dehydrogenase cannot adequately account for the results. On the other hand, the model based on both negative and positive co-operativity gives a good fit to the experimental points. 5. The conclusions are discussed in relation to current knowledge of the structure and mechanism of glutamate dehydrogenase.  相似文献   

20.
For (S)-thiirancarboxylic acid a second-order rate constant of k2nd = 222 M(-1) min(-1) for the irreversible inhibition of papain was determined. The ethyl and methyl ester do not inhibit the enzyme time-dependently. An improved synthesis of enantiomerically pure thiirancarboxylic acid is described. It is shown that thiirancarboxylates can be substrates for serine proteases (alpha-chymotrypsin) and esterases (pig liver esterase) and even for metallo proteases (thermolysin).  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