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 656 毫秒
1.
The positive electrostatic environment of the active site of prolyl oligopeptidase was investigated by using substrates with glutamic acid at positions P2, P3, P4, and P5, respectively. The different substrates gave various pH rate profiles. The pKa values extracted from the curves are apparent parameters, presumably affected by the nearby charged residues, and do not reflect the ionization of a simple catalytic histidine as found in the classic serine peptidases like chymotrypsin and subtilisin. The temperature dependence of kcat/Km did not produce linear Arrhenius plots, indicating different changes in the individual rate constants with the increase in temperature. This rendered it possible to calculate these constants, i.e. the formation (k1) and decomposition (k-1) of the enzyme-substrate complex and the acylation constant (k2), as well as the corresponding activation energies. The results have revealed the relationship between the complex Michaelis parameters and the individual rate constants. Structure determination of the enzyme-substrate complexes has shown that the different substrates display a uniform binding mode. None of the glutamic acids interacts with a charged group. We conclude that the specific rate constant is controlled by k1 rather than k2 and that the charged residues from the substrate and the enzyme can markedly affect the formation but not the structure of the enzyme-substrate complexes.  相似文献   

2.
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   

3.
The rate constants for both acylation and deacylation of beta-lactamase PC1 from Staphylococcus aureus and the RTEM beta-lactamase from Escherichia coli were determined by the acid-quench method [Martin & Waley (1988) Biochem. J. 254, 923-925] with several good substrates, and, for a wider range of substrates, of beta-lactamase I from Bacillus cereus. The values of the acylation and deacylation rate constants for benzylpenicillin were approximately the same (i.e. differing by no more than 2-fold) for each enzyme. The variation of kcat./Km for benzylpenicillin with the viscosity of the medium was used to obtain values for all four rate constants in the acyl-enzyme mechanism for all three enzymes. The reaction is partly diffusion-controlled, and the rate constant for the dissociation of the enzyme-substrate complex has approximately the same value as the rate constants for acylation and deacylation. Thus all three first-order rate constants have comparable values. Here there is no single rate-determining step for beta-lactamase action. This is taken to be a sign of a fully efficient enzyme.  相似文献   

4.
M Philipp  I H Tsai  M L Bender 《Biochemistry》1979,18(17):3769-3773
The p-nitrophenyl esters of straight-chain fatty acids were used as substrates of the enzyme subtilisin Novo (EC 3.4.4.16) and its chemically produced artificial enzyme thiolsubtilisin. Subtilisin and thiolsubtilisin pH--activity profiles were determined, and kinetic effects of the active site O-S substitution were observed. Among the substrates tested, both enzymes show highest specificity with p-nitrophenyl butyrate. It was also found that subtilisin is more sensitive to changes in substrate chain length than is thiolsubtilisin. Second-order acylation rate constants (k2/Ks) are remarkably similar for both enzymes. However, thiolsubtilisin deacylation rate constants and Km values are lower than analogous subtilisin constants. While thiolsubtilisin deacylation rate constants give a pH profile identical with that of subtilisin, the pH profile of thiolsubtilisin acylation rate constants shows an active site pK value lowered from the subtilisin pK of 7.15 and exhibits an inflection point at pH 8.45, which is absent in subtilisin.  相似文献   

5.
The influence of water on the kinetics of alcoholysis of methyl propionate and n-propanol catalyzed by immobilized lipase B from Candida antarctica was studied in a continuous solid/gas reactor. In this reactor, the solid phase is composed of a packed enzymatic sample which is percolated by gaseous nitrogen, simultaneously carrying gaseous substrates to the enzyme while removing reaction products. In this system, interactions between the enzyme and nonreacting molecules are avoided, since no solvent is present, and it is thus more easy to assess the role of water. To this end, alcohol inhibition constant, substrates dissociation constants as well as acylation rate constant and ratio of acylation to deacylation rate constants have been determined as a function of water activity (a(w)). Data obtained highlight that n-propanol inhibition constant and dissociation constant of methyl propionate are a lot affected by a(w) variations whereas water has no significant effect on the catalytic acylation step nor on the ratio of acylation to deacylation rate constants. These results suggest the water-independent character of the transition step.  相似文献   

6.
Hydrolyses catalyzed by bovine pancreatic trypsin and porcine pancreatic kallikrein were studied using synthetic peptide substrates of the type E chi-L chi 2-L chi 1 decreases Y and E chi-L chi 3-L chi 2-L chi 1 decreases Y with L chi 1 = Arg defining the hydrolysis position (indicated by the arrow). The leaving moiety Y was -OCH3, -NH-C6H4-p-NO2 and -Ala-NH2. Insight into interactions occurring between the active site of the enzymes and the acyl moiety of the substrates was gained by studying the influence on hydrolysis rate of structural variation of residues L chi 2 and L chi 3. Parallel analyses of the hydrolyses of the ester, anilide, and peptide substrates having the same acyl moiety considerably facilitated the interpretation of the kinetic data. Trypsin, but not kallikrein, displayed high reactivity even with relatively short substrates. Ac-Ala-Arg-Ala-NH2, for example, was a better substrate for trypsin than for kallikrein by a factor of 1.3 X 10(4) in terms of kcat and 5.9 X 10(4) in terms of kcat/Km. Reactivity differences of such magnitude were related to two main differences in enzyme-substrate interactions: the interaction of the arginine side chain of the substrate with the specificity pocket of the enzyme is optimal for trypsin but poor for kallikrein and the number of hydrogen bonds formed by the enzyme with the backbone section of the substrate on both sides of the specific residue is larger in the case of trypsin. The latter difference is found to be related to the structure of amino-acid residue 192 which is glutamine in trypsin and methionine in kallikrein.  相似文献   

7.
J A Mattis  J S Fruton 《Biochemistry》1976,15(10):2191-2194
Kinetic measurements have been performed on the action of papain on mansyl-Gly-Val-Glu-Leu-Gly and on mansyl-Gly-Gly-Val-Glu-Leu-Gly, both of which are cleaved solely at the Glu-Leu bond under the conditions of our experiments. Stopped-flow experiments have shown that, under conditions of enzyme excess, the enhancement of the fluorescence of the mansyl group upon association of each of the oligopeptide substrates with papain is a biphasic process. A very rapid initial increase in fluorescence is followed by a slower first-order fluorescence enhancement. The observed rate constant for the latter process is greater with the mansyl pentapeptide than with the mansyl hexapeptide. A similar biphasic fluorescence change is seen upon the interaction of the mansyl peptides with mercuripapain, but the second step is much slower than in the case of the active enzyme. The rate of the second step in the association of active papain with the mansyl paptides shows saturation with increasing enzyme concentration, supporting the view that an initial enzyme-substrate complex (ES) is converted in a first-order process to the complex (ES) that undergoes cleavage to form products. The hydrolysis of the Glu-Leu bond is associated with a first-order decrease in fluorescence, as a consequence of the formation of the mansyl peptide product, which is bound less strongly than the substrate. The rate constant for this process is about 140 times greater with the mansyl hexapeptide than with the mansyl pentapeptide, thus giving further indication of the importance of secondary enzyme-substrate interactions in the efficiency of papain catalysis. For each of the two mansyl peptides, the values of the rate constants and the apparent Michaelis constants associated with the cleavage of the Glu-Leu bond, as determined by stopped-flow measurements under conditions of enzyme excess, were the same, within the precision of the data, as those estimated from experiments under conditions of substrate excess, where the formation of Leu-Gly was determined by means of the fluorescamine reaction. This indicates that, with these substrates, the rate-limiting step in the overall catalytic process is associated with the breakdown of ES. Estimates are given of the dissociation constant of ES and of the rate constants in the interconversion of ES and ES.  相似文献   

8.
The kinetic constants for the papain-catalyzed hydrolysis of a series of substrates with glycine or alanine in the P1 position are discussed. The substrates have N-benzoyl, N-(p-nitrobenzoyl), N-(beta-phenylpropionyl), or N-(methyloxycarbonyl)phenylalanine attached to the P1 moiety, and kinetic constants are obtained for both esters and thiono esters. The results for the hydrolysis of esters can be readily interpreted in terms of the known specificity of papain. For any glycine ester the change in kcat/Km upon substituting C=S for C=O or upon substituting an alpha-CH3 group is minimal. However, upon making both these substitutions, i.e., going from a glycine ester to an alanine thiono ester substrate, larger changes are seen for this ratio. Data for N-benzoyl- and N-(beta-phenylpropionyl)glycine and -alanine methyl thiono esters show that k2 is the parameter most affected by the double C=S and alpha-CH3 substitution. A further conclusion is that the deacylation rate constants for any pair of glycine and alanine dithioacyl papains are similar; e.g., for the intermediates based on the "good" substrates PheAla and PheGly k3 differs by only 20%. This is a surprising finding in light of the very different conformations and interactions of the bound acyl groups revealed by resonance Raman spectroscopy and raises the possibility that specific stereochemical effects, such as the oxyanion hole and general base catalysis, are not operating in the hydrolysis of dithioacyl papains.  相似文献   

9.
The specificity of porcine elastase (EC 3.4.4.7) has been studied. Ethyl esters derived from benzoyl amino acids with straight side chains are better substrates than those with branched side chains; the best substrate is norvaline ester. In the series of benzoylalanine alkyl esters the alcohol moiety markedly affects the susceptibility. The benzyl ester was found to be the best nonactivated substrate derived from monomeric amino acid. With elastase acylation is rate limiting, in contrast to chymotrypsin and trypsin where deacylation is generally the rate determining step with specific ester substrates.  相似文献   

10.
R L Kogan  T H Fife 《Biochemistry》1985,24(11):2610-2614
The second-order rate constants k2/Km for acylation of alpha-chymotrypsin by a series of N-acylimidazole derivatives of aliphatic carboxylic acids have been determined at 30 degrees C by proflavin displacement from the active site. With cyclohexyl-substituted N-acylimidazoles, the rate constants increase with increasing chain length of the acyl group; i.e., k2/Km is in the order cyclohexylcarbonyl less than cyclohexylacetyl less than beta-cyclohexylpropionyl. The latter substrate has k2/Km = 1.2 X 10(6) M-1 s-1 at pH 8.0, which appears to be a maximum value for N-acylimidazole substrates. A further increase in the chain length of the acyl group with (gamma-cyclohexylbutyryl)imidazole results in a decrease in k2/Km. Hydrophobic effects of the hydrocarbon acyl groups are of predominant importance with regard to the relative values of k2/Km for aliphatic N-acylimidazole substrates. There is a linear correlation of the logarithms of the rate constants at pH 8.0 with the hydrophobic substituent constants, pi, having a slope of 1.71 (r = 0.90). On the other hand, there is little apparent correlation with the Taft steric effect constants, Es. A four-parameter equation including both pi and Es improved the correlation only slightly [log (k2/Km) = 1.88 pi + 1.01 Es + C]. In contrast, steric effects as reflected in the Es constants are the major influence in acylation of the enzyme by corresponding p-nitrophenyl esters. There are very likely significant differences in transition-state structure with the two types of substrates.  相似文献   

11.
Porter DJ  Short SA 《Biochemistry》2000,39(38):11788-11800
The catalytically active form of monofunctional yeast orotidine-5'-phosphate decarboxylase was a dimer (E(2)). The dimer equilibrium dissociation constant was 0.25 microM in 0.01 M MOPS Na(+) at pH 7.2. The bimolecular rate constant for dimer formation was 1.56 microM(-1) s(-1). The dimeric form of the enzyme was stabilized by NaCl such that the enzyme was E(2) in 100 mM NaCl at all concentrations of enzyme tested. The kinetics of binding of OMP to E(2) was governed by two ionizations (pK(1) = 6.1 and pK(2) = 7.7). From studies with substrate analogues, the higher pK was assigned to a group on the enzyme that interacted with the pyrimidinyl moiety. The value of the lower pK was dependent on the substrate analogue, which suggested that it was not exclusively the result of ionization of the phosphoryl moiety. During the decarboxylation of OMP, the fluorescence of E(2) was quenched over 20%. The enzymatic species with reduced fluorescence was a catalytically competent intermediate that had kinetic properties consistent with it being the initial enzyme-substrate complex. The stoichiometry for binding of OMP to E(2) was one OMP per enzyme monomer. The value of the first-order rate constant for conversion of the enzyme-substrate complex to free enzyme (36 s(-1)) calculated from a single turnover experiment ([E] > [S]) was slightly greater than the value of k(cat), 20 s(-1) (corrected for stoichiometry), calculated from steady-state data. In the single turnover experiments, the enzyme was E(2)*S, whereas in the steady-state turnover the experiment enzyme was E(2)*S(2). The similarity of these values suggested that the subunits were catalytically independent such that E(2)*S(2) could be treated as E*S and that conversion of the enzyme-substrate complex to E was k(cat). Kinetic data for the approach to the steady-state with OMP and E(2) yield a bimolecular association rate complex of 62 microM(-1) s(-1)and a dissociation rate constant for E*S of 60 s(-1). The commitment to catalysis was 0.25. By monitoring the effect of carbonic anhydrase on [H(+)] changes during a single turnover experiment, the initial product of the decarboxylation reaction was shown to be CO(2) not HCO(3-). UMP was released from the enzyme concomitantly with CO(2) during the conversion of E*S to E. Furthermore, the enzyme removed an enzyme equivalent of H(+) from solvent during this step of the reaction. The bimolecular rate constants for association of 6-AzaUMP and 8-AzaXMP, substrate analogues with markedly different nucleobases, had association rate constants of 112 and 130 microM(-1) s(-1), respectively. These results suggested that the nucleobase did not contribute significantly to the success of formation of the initial enzyme-substrate complex.  相似文献   

12.
A pre-steady state kinetic analysis of the stimulation by monovalent cations of the activity of bovine activated protein C (APC) and a proteolytic fragment of APC, des-1-41-light chain activated protein C (GDAPC), toward the substrate, 4-methylumbelliferyl p-guanidinobenzoate, has been undertaken. With the cations Na+ and Cs+, at least two cation sites, or classes of sites, on APC were found to be important to the kinetic effects observed. For GDAPC, with both monovalent cations investigated, a single cation-binding site, or class of sites, of kinetic importance was discovered. The most general mechanism that fits all kinetic data was a rapid equilibrium type, with the cation(s) (A) and substrate (S) binding to the enzyme in a random fashion. Cations were found to be essential activators, and only formation of the EAS or EA2S complex led to product generation. For each enzyme, stimulation of the reaction rates was found to be chiefly due to a dramatic enhancement by monovalent cations of the rate constant (k2) for acylation of the enzyme since the dissociation constant (Ks) for enzyme-substrate interactions was increased in the presence of cations, and the deacylation rate constant (k3) was not affected by these activators.  相似文献   

13.
We synthesized the following four new peptide substrates, Suc-Phe-Leu-pNA, Suc-Phe-Leu-NMec, Suc-Phe-Leu-ONPh, and Pht-Phe-Leu-pNA, and we applied the proton inventory method to their hydrolysis by papain. Useful relationships between the rate constants of the catalytic reaction have been established and contributed to the elucidation of the hydrolytic mechanism of papain. For all amide substrates, the parameter K(S) and the rate constants k(1), k(-)(1), and k(2) were estimated. Moreover, it was found that k(cat)/K(m) = k(1) for all four substrates, while two exchangeable hydrogenic sites, one in the ground state and another in the transition state, generate an inverse isotope effect during the reaction governed by this parameter. The proton inventories of both k(2) and k(3) are essentially linear, whatever the acyl moiety and/or the leaving group of the substrate. The proton inventories of K(S) are also essentially linear for all amide substrates, while the observed large isotope effect of about 3 to 9 originates from a single hydrogenic site in the product state. This latter, in agreement to both the small transition state fractionation factors found for k(cat)/K(m) (or k(1)) and the unit ground-state fractionation factors found for k(2), argues for the formation of a tetrahedral adduct during the reaction governed by the k(1) parameter. Furthermore, papain acts as a one-proton catalyst during acylation or deacylation, both of which proceed through similar concerted reaction pathways, where a nucleophilic attack is accompanied by the movement of one proton.  相似文献   

14.
The hydrolysis of sodium 3-dansylamidomethyl-7-beta (thienyl-2')-acetamido-ceph-3-em-4-oate, catalyzed by the beta-lactamase of Staphylococcus aureus PC1, has previously been shown (Anderson, E. G., and Pratt, R. F. (1981) J. Biol. Chem. 256, 11401-11404) to follow the reaction scheme Formula; see text. where ES' is an enzyme-substrate complex in which the substrate has undergone nucleophilic attack at the beta-lactam carbonyl group and P is product. Acid quenching of the reaction mixture has now been shown to yield, in amounts predicted by the rate constants, a covalent enzyme-substrate complex. The liability of this complex in alkaline solution is suggestive of that of an ester. Together, all of these results prove that the turnover of this apparently normal substrate by a class A beta-lactamase involves an acyl-enzyme intermediate. In the case of another fluorescent substrate, dansylcephalexin, no intermediate analogous to ES' accumulated during catalysis; presumably here, acylation of the enzyme is rate-determining. The pH profiles (pH 4-9) of the pre-steady state rate constants for hydrolysis of the former substrate have also been determined. Binding (1/K8) is pH invariant except at low pH where it weakens, probably because of substrate protonation and/or a protein conformational change. The rate constants, k2, k-2, and k3, are pH invariant at low pH but decrease at higher pH in a way which can be described by ionization of an essential acid of pKa around 7.7. This may be the same acid for each constant, being either an active participant at the active site, or a more distant acid which controls an essential conformational change.  相似文献   

15.
Kinetic constants for the hydrolysis by porcine tissue beta-kallikrein B and by bovine trypsin of a number of peptides related to the sequence of kininogen (also one containing a P2 glycine residue instead of phenylalanine) and of a series of corresponding arginyl peptide esters with various apolar P2 residues have been determined under strictly comparative conditions. kcat and kcat/Km values for the hydrolysis of the Arg-Ser bonds of the peptides by trypsin are conspicuously high. kcat for the best of the peptide substrates, Ac-Phe-Arg-Ser-Val-NH2, even reaches kcat for the corresponding methyl ester, indicating rate-limiting deacylation also in the hydrolysis of a peptide bond by this enzyme. kcat/Km for the hydrolysis of the peptide esters with different nonpolar L-amino acids in P2 is remarkably constant (range 1.7), as it is for the pair of the above pentapeptides with P2 glycine or phenylalanine. kcat for the ester substrates varies fivefold, however, being greatest for the P2 glycine compounds. Obviously, an increased potential of a P2 residue for interactions with the enzyme lowers the rate of deacylation. In contrast to results obtained with chymotrypsin and pancreatic elastase, trypsin is well able to tolerate a P3 proline residue. In the hydrolysis of peptide esters, tissue kallikrein is definitely superior to trypsin. Conversely, peptide bonds are hydrolyzed less efficiently by tissue kallikrein and the acylation reaction is rate-limiting. The influence of the length of peptide substrates is similar in both enzymes and indicates an extension of the substrate recognition site from subsite S3 to at least S'3 of tissue kallikrein and the importance of a hydrogen bond between the P3 carbonyl group and Gly-216 of the enzymes. Tissue kallikrein also tolerates a P3 proline residue well. In sharp contrast to the behaviour of trypsin is the very strong influence of the P2 residue in tissue-kallikrein-catalyzed reactions. kcat/Km varies 75-fold in the series of the dipeptide esters with nonpolar L-amino acid residues in P2, a P2 glycine residue furnishing the worst and phenylalanine the best substrate, whereas this exchange in the pentapeptides changes kcat/Km as much as 730-fold. This behaviour, together with the high value of kcat/Km for Ac-Phe-Arg-OMe of 3.75 X 10(7) M-1 s-1, suggests rate-limiting binding (k1) in the hydrolysis of the best ester substrates.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

16.
Hays JL  Watowich SJ 《Biochemistry》2004,43(32):10570-10578
Although oligomerization of receptor tyrosine kinases (RTKs) is necessary for receptor activation and signaling, a quantitative understanding of how oligomerization mediates these critical processes does not exist. We present a comparative thermodynamic analysis of functionally active dimeric and functionally inactive monomeric soluble analogues of the c-MET RTK, which clearly reveal that oligomerization regulates the binding affinity and binding kinetics of the kinase toward ATP and tyrosine-containing peptide substrates. Thermodynamic binding data for oligomeric c-MET were obtained from the dimeric TPR-MET oncoprotein, a functionally active fusion derivative of the c-MET RTK. This naturally occurring oncoprotein contains the cytoplasmic domain of c-MET fused to a coiled coil dimerization domain from the nuclear pore complex. Comparative data were obtained from a soluble monomeric kinase compromising the c-MET cytoplasmic domain (cytoMET). Significantly, under equilibrium binding conditions, the oligomeric phosphorylated kinase showed a significantly lower dissociation constant (K(d,dimer) = 11 microM) for a tyrosine-containing peptide derived from the C-terminal tail of the c-MET RTK when compared to the phosphorylated monomeric kinase cytoMET (K(d,monomer) = 140 microM). Surprisingly, equilibrium dissociation constants measured for the kinase and ATP were independent of the oligomerization state of the kinase (approximately 10 microM). Stopped-flow analysis of peptide substrate binding showed that the association rate constants (k(2)) differed 2-fold and dissociation rate constants (k(-2)) differed 10-fold when phosphorylated TPR-MET was compared to phosphorylated cytoMET. ATP binding abrogated the differences in k(2) rates observed between the two oligomeric states of the c-MET cytoplasmic domain. These results clearly imply that oligomerization induces important thermodynamic and conformational changes in the substrate binding regions of the c-MET protein and provide quantitative mechanistic insights into the necessary role of oligomerization in RTK activation.  相似文献   

17.
Anderson JW  Pratt RF 《Biochemistry》2000,39(40):12200-12209
Bacterial cell walls are cross-linked in the final step of biosynthesis by specific D-alanyl-D-alanine(DD)-peptidases/transpeptidases. The natural substrates of these enzymes should therefore be segments of peptidoglycan, but high specificity for such structures has yet to be demonstrated. The binding of dipeptides to the extended substrate binding site of the Streptomyces R61 DD-peptidase has been studied by means of a fluorescent beta-lactam probe. It was found that dipeptides of structure Gly-L-Xaa have affinity for a subsite adjacent to the beta-lactam binding site. Hydrophobic peptides such as Gly-L-Met and Gly-L-aminocaprylic acid had the greatest affinity for this site, with dissociation constants in each case of 0.19 mM. A combination of this motif with the C-terminal D-alanyl-D-alanine moiety required of a DD-peptidase substrate yielded a new substrate, glycyl-L-alpha-amino-epsilon-pimelyl-D-alanyl-D-alanine. Steady-state kinetic measurements established this compound as the most specific peptide substrate yet discovered for a DD-peptidase by at least 3 orders of magnitude (k(cat) = 69 s(-1), K(m) = 7.9 microM, k(cat)/K(m) = 8.7 x 10(6) s(-1) M(-1)); acylation was rate-determining at saturation. This substrate, presumably not coincidentally, contains the acyl donor and acceptor moieties, appropriately separated, of the Streptomyces peptidoglycan structure. This general method of approach should be of value in the search for specific substrates and inhibitors (antibiotics) of other DD-peptidases.  相似文献   

18.
Asgeirsson B  Cekan P 《FEBS letters》2006,580(19):4639-4644
Temperature imposes limits on where life can thrive and this is evident in the evolution of the basic structural properties of proteins. Cold-adaptation of enzymes is one example, where the catalytic rate constant (k(cat)) is increased compared with hot-acclimated homologous under identical assay conditions. Trypsin I from Atlantic cod (Gadus morhua) has catalytic efficiency (k(cat)/K(m)) for amide hydrolysis that is 17-fold larger than observed for bovine trypsin. Here, the individual rate-constants for association of substrate (k(1)), dissociation of substrate (k(-1)), and acylation of the enzyme (k(2)) have been determined using benzoyl-Arg-p-nitroanilide or benzyloxycarbonyl-Gly-Pro-Arg-p-nitroanilide as substrates. Rather unexpectedly, by far the largest difference (37-fold increase) was observed in k(1), the rate constant for binding of substrate. The cold-adaptation of the dissociation and catalytic steps were not as prominent (increased by 3.7-fold). The length of substrate did have an effect by increasing the reaction rate by 70-fold, and again, the step most affected was the initial binding-step.  相似文献   

19.
Steady-state and presteady-state kinetic parameters for plasmins derived from the two rabbit plasminogen isozymes have been determined. Steady-state kinetic experiments with N-α-tosyl-l-arginine methyl ester indicate that each isozyme has a similar V. Plasmin isozyme 2 has a higher Km value for this substrate as well as a higher Ki, for the competitive inhibitor, benzamidine-HCl. Presteady-state kinetic experiments, using the p-nitrophenyl esters of p-(methylethylsulfoniummethyl)benzoate, p-(pyridiniummethyl) benzoate, p-(thiouroniummethyl)benzoate and p-(guanidinium)benzoate, indicate that each plasmin has similar rate constants of acylation (k2). However, values of the dissociation constant (KS) indicate that plasmin isozyme 1 has a greater binding affinity for these substrates than does isozyme 2. The magnitude of this difference varies with the substrate and is the greatest for those containing analogs of the guanidino moiety of l-arginine.  相似文献   

20.
We investigated the effects of mutations at positions 164 and 179 of the TEM(pUC19) beta-lactamase on turnover of substrates. The direct consequence of some mutations at these sites is that clinically important expanded-spectrum beta-lactams, such as third-generation cephalosporins, which are normally exceedingly poor substrates for class A beta-lactamases, bind the active site of these mutant enzymes more favorably. We employed site-saturation mutagenesis at both positions 164 and 179 to identify mutant variants of the parental enzyme that conferred resistance to expanded-spectrum beta-lactams by their enhanced ability to turn over these antibiotic substrates. Four of these mutant variants, Arg(164) --> Asn, Arg(164) --> Ser, Asp(179) --> Asn, and Asp(179) --> Gly, were purified and the details of their catalytic properties were examined in a series of biochemical and kinetic experiments. The effects on the kinetic parameters were such that either activity with the expanded-spectrum beta-lactams remained unchanged or, in some cases, the activity was enhanced. The affinity of the enzyme for these poorer substrates (as defined by the dissociation constant, K(s)) invariably increased. Computation of the microscopic rate constants (k(2) and k(3)) for turnover of these poorer substrates indicated either that the rate-limiting step in turnover was the deacylation step (governed by k(3)) or that neither the acylation nor deacylation became the sole rate-limiting step. In a few instances, the rate constants for both the acylation (k(2)) and deacylation (k(3)) of the extended-spectrum beta-lactamase were enhanced. These results were investigated further by molecular modeling experiments, using the crystal structure of the TEM(pUC19) beta-lactamase. Our results indicated that severe steric interactions between the large 7beta functionalities of the expanded-spectrum beta-lactams and the Omega-loop secondary structural element near the active site were at the root of the low affinity by the enzyme for these substrates. These conclusions were consistent with the proposal that the aforementioned mutations would enlarge the active site, and hence improve affinity.  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