首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
B Asbóth  L Polgár 《Biochemistry》1983,22(1):117-122
X-ray diffraction studies suggested that the tetrahedral intermediate formed during the catalysis by serine and thiol proteinases can be stabilized by hydrogen bonds from the protein to the oxyanion of the intermediate [cf. Kraut, J. (1977) Annu. Rev. Biochem. 46, 331-358; Drenth, J., Kalk, K.H., & Swen, H.M. (1976) Biochemistry 15, 3731-3738]. To obtain evidence in favor or against this hypothesis, we synthesized thiono substrates (the derivatives of N-benzoyl-glycine methyl ester and N-acetylphenylalanine ethyl ester) containing a sulfur in place of the carbonyl oxygen atom of the scissile ester bond. We anticipated that this relatively subtle structural change specifically directed to the oxyanion binding site should produce serious catalytic consequences owing to the different properties of oxygen and sulfur if transition-state stabilization in the oxyanion hole is indeed important. In fact, while in alkaline hydrolysis the chemical reactivities of oxygen esters and corresponding thiono esters proved to be similar, neither chymotrypsin nor subtilisin hydrolyzed the thiono esters at a measurable rate. This result substantiates the crucial role of the oxyanion binding site in serine proteinase catalysis. On the basis of the similar values of the binding constants found for oxygen esters and their thiono counterparts, it can be concluded that the substitution of sulfur for oxygen significantly influences transition state stabilization but not substrate binding. The thiol proteinases papain and chymopapain react with the oxygen and thiono esters of N-benzoylglycine at similar rates. Apparently, in these reactions the above stabilizing mechanism is absent or not important, which is a major mechanistic difference between the catalyses by serine and thiol proteinases.  相似文献   

2.
A series of arylalkanoate esters and alpha-acetamidoarylalkanoate esters were tested as substrates for alpha-chymotrypsin and subtilisin BPN'. Chymotrypsin hydrolysed N-acetyl-l-phenylalanine methyl ester and methyl 4-phenylbutyrate faster than their respective higher and lower homologues, whereas methyl 2-acetamido-6-phenylhexanoate and methyl 6-phenylhexanoate were better substrates for subtilisin than their lower homologues. N-Acetyl-l-tryptophan methyl ester and its analogue, N-acetyl-3-(1-naphthyl)-alanine methyl ester, were hydrolysed 23 times faster by chymotrypsin than by subtilisin. These results indicate that the binding site of alpha-chymotrypsin is roughly 1.1nm (11A) long and curved, whereas that of subtilisin is a longer system and less curved. The stereo-specificity during the hydrolysis of typical substrates by both enzymes was found to vary over a wide range. The enhancing effect of the alpha-acetamido group in the l-series of substrates and the detrimental effect in the d-series of substrates also varies considerably.  相似文献   

3.
The interaction of a variety of substrates with Pseudomonas aeruginosa native amidase (E.C. 3.5.1.4), overproduced in an Escherichia coli strain, was investigated using difference FTIR spectroscopy. The amides used as substrates showed an increase in hydrogen bonding upon association in multimers, which was not seen with esters. Evidence for an overall reduction or weakening of hydrogen bonding while amide and ester substrates are interacting with the enzyme is presented. The results describe a spectroscopic approach for analysis of substrate–amidase interaction and in situ monitoring of the hydrolysis and transferase reaction when amides or esters are used as substrates.  相似文献   

4.
P Carter  L Abrahmsén  J A Wells 《Biochemistry》1991,30(25):6142-6148
A mutant of the serine protease, subtilisin BPN', in which the catalytic His64 is replaced by Ala (H64A), is very specific for substrates containing a histidine, presumably by the substrate-bound histidine assisting in catalysis [Carter, P., & Wells, J.A. (1987) Science (Washington, D.C.) 237, 394-399]. Here we probe the catalytic mechanism of H64A subtilisin for cleaving His and non-His substrates. We show that the ratio of aminolysis to hydrolysis is the same for ester and amide substrates as catalyzed by the H64A subtilisin. This is consistent with formation of a common acyl-enzyme intermediate for H64A subtilisin, analogous to the mechanism of the wild-type enzyme. However, the catalytic efficiencies (kcat/KM) for amidase and esterase activities with His-containing substrates are reduced by 5000-fold and 14-fold, respectively, relative to wild-type subtilisin BPN, suggesting that acylation is more compromised than deacylation in the H64A mutant. High concentrations of imidazole are much less effective than His substrates in promoting hydrolysis by the H64A variant, suggesting that the His residue on the bound (not free) substrate is involved in catalysis. The reduction in catalytic efficiency kcat/KM for hydrolysis of the amide substrate upon replacement of the oxyanion stabilizing asparagine (N155G) is only 7-fold greater for wild-type than H64A subtilisin. In contrast, the reductions in kcat/KM upon replacement of the catalytic serine (S221A) or aspartate (D32A) are about 3000-fold greater for wild-type than H64A subtilisin, suggesting that the functional interactions between the Asp32 and Ser221 with the substrate histidine are more compromised in substrate-assisted catalysis.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
To investigate how the conformational flexibility of subtilisin affects its ability to discriminate between enantiomeric amino acid and ester substrates for the subtilisin-catalyzed reaction in an organic solvent, the flexibility around the active site and the surface of subtilisin was estimated from the mobility of a spin label bound to subtilisin by ESR spectroscopy. Many studies on enzyme flexibility focus on the active site. Both the surface and active site flexibility play an important role in the enantioselectivity enhancement of the enzyme-catalyzed reaction. It was found, however, that the different behavior observed for the enantioselectivity between the amino acid and ester substrates could be correlated with the flexibility around the surface rather than the flexibility at the active site of subtilisin. In other words, for the ester substrates, the greater flexibility around the surface of subtilisin induced by a conformational change resulting from the presence of an additive such as DMSO is essential for the enantioselectivity enhancement. This model is also supported by the Michaelis-Menten kinetic parameters for each enantiomeric substrate. Our findings provide insight into the enantioselectivity enhancement for the resolution of enantiomers for enzyme-catalyzed reactions in organic solvents.  相似文献   

6.
Subtilisin BPN' hydrolysed N-acetyl-l-3-(2-naphthyl)-alanine methyl ester, N-acetyl-l-leucine methyl ester and N-acetyl-l-valine methyl ester, faster than alpha-chymotrypsin. Of eight ;locked' substrates tested, only methyl 5,6-benzindan-2-carboxylate was hydrolysed faster by subtilisin, whereas the other esters were better substrates for chymotrypsin. Compared with the values for chymotrypsin, the stereospecific ratios during the hydrolysis of the optically active locked substrates by subtilisin were decreased by one and two orders of magnitude for bi- and tri-cyclic substrates respectively. The polar groups adjacent to the alpha-carbon atom of locked substrates did not contribute significantly to the reactivity of the more active optical isomers, but had a detrimental effect on the less active antipodes during hydrolysis by both the enzymes. These studies show that the binding site of subtilisin BPN' is longer and broader than that of alpha-chymotrypsin.  相似文献   

7.
Limited proteolysis of carboxypeptidase A from bovine pancreas with subtilisin Carlsberg generates a stable intermediate, carboxypeptidase S, whose esterase and peptidase activities are increased and decreased, respectively, under standard assay conditions. Carboxypeptidase S was isolated by affinity chromatography. Sequence analysis shows that it is cleaved solely at the Ala154-Gly155 bond. Its enzymatic properties were determined under stopped-flow conditions with Dns-Gly-Ala-Phe and its ester analogue Dns-Gly-Ala-OPhe. For both substrates, the Km values are increased 30-40-fold. The kcat value for peptide hydrolysis is virtually unaffected whereas that for ester hydrolysis is increased 10-fold. The magnitude of the Km effect is equivalent to a loss of 9 kJ/mol of binding energy and likely reflects a disruption of the network of hydrogen bonds that links Tyr-248 and Arg-145 to the backbone carbonyls of Ala-154 and Gly-155. The difference in kcat effects for the two substrate classes is related to differences in the chemical nature of the rate-determining step. Product release is rate determining for catalytic hydrolysis of ester substrates, and hence, the increase in kcat indicates that dissociation of products is facilitated as a result of the Ala154-Gly155 bond scission. The changes in enzymatic activity accompanying limited proteolysis are due to conformational alterations in the vicinity of the active center of the molecule. The affinity of a monoclonal antibody, mAb 100, directed toward the antigenic determinant located between residues 209 and 218 in carboxypeptidase A is diminished considerably for carboxypeptidase S.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

8.
As part of a search for peptides that have specificity for selected protein kinases, the possibility that adenosine cyclic 3',5'-phosphate dependent protein kinase (A-kinase) recognizes the hydrogen-bonding potential of its peptide substrates was investigated. A-Kinase catalyzes the phosphorylation of five N alpha-methylated and four depsipeptide derivatives of Leu-Arg-Arg-Ala-Ser-Leu-Gly (peptide 1) at rates that differ by at least 7 orders of magnitude. These peptide 1 analogues each lack the ability to donate a hydrogen bond at selected positions in the peptide chain. If a particular amide hydrogen of a peptide amide is involved in hydrogen bonding, which is important for enzyme recognition, the prediction is that peptides which contain an ester or a N-methylated bond at that position in peptide 1 will be comparatively poor substrates. In contrast, if a depsipeptide has a reactivity comparable to that of peptide 1 but the analogous N-methylated peptide has a poor reactivity with A-kinase, the result might indicate that the N-methyl group causes unfavorable steric effects. The depsipeptide that lacks a Leu6 amide proton is a good substrate for A-kinase, but the corresponding N-methylated peptide is phosphorylated far less efficiently. This result and others presented in this paper suggest that although enzyme-substrate hydrogen bonding may play some role in A-kinase catalysis of phosphoryl group transfer, other explanations are necessary to account for the relative reactivities of N alpha-methylated and depsi-containing peptide 1 analogues.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
Determination of individual rate constants for enzyme-catalyzed reactions is central to the understanding of their mechanism of action and is commonly obtained by stopped-flow kinetic experiments. However, most natural substrates either do not fluoresce/absorb or lack a significant change in their spectra while reacting and, therefore, are frequently chemically modified to render adequate molecules for their spectroscopic detection. Here, isothermal titration calorimetry (ITC) was used to obtain Michaelis–Menten plots for the trypsin-catalyzed hydrolysis of several substrates at different temperatures (278–318 K): four spectrophotometrically blind lysine and arginine N-free esters, one N-substituted arginine ester, and one amide. A global fitting of these data provided the individual rate constants and activation energies for the acylation and deacylation reactions, and the ratio of the formation and dissociation rates of the enzyme–substrate complex, leading also to the corresponding free energies of activation. The results indicate that for lysine and arginine N-free esters deacylation is the rate-limiting step, but for the N-substituted ester and the amide acylation is the slowest step. It is shown that ITC is able to produce quality kinetic data and is particularly well suited for those enzymatic reactions that cannot be measured by absorption or fluorescence spectroscopy.  相似文献   

10.
From stereochemical considerations and model building the following conclusions were drawn for the stereochemistry of the catalytic steps of chymotrypsin and subtilisin. (1) In contrast to previous stereochemical investigations, rotation of 120° or more of the oxygen atom of the “reactive” serine residue is not possible in the course of the reaction with specific substrates. (2) During catalysis the serine oxygen atom is approximately in the position found in the crystalline enzyme, i.e. at a distance of about 3 Å from the nitrogen atom of the catalytically important histidine residue. (3) The detailed stereochemical mechanism involves the formation of a strained tetrahedral intermediate and a strained acylenzyme. The strain energy is supplied by the formation of a hydrogen bond between the enzyme and a specific substrate. (4) The geometry of proton transfers in the intimate encounter complex of chymotrypsin is slightly but significantly different from that of subtilisin.  相似文献   

11.
Two mechanisms recently proposed to account for the action of serine in proteases differ on the matter of hydrogen bonding between enzyme and substrate in the Michaelis complex and acyl enzyme. One mechanism involves the removal of hydrogen bonds as a means of raising the free energy of the Michaelis complex toward that of the transition state and thereby effecting catalysis. This report shows that the kinetic parameters k2Ks and k3 for the α-chymotrypsin-catalyzed hydrolysis of a series of halogen-substituted phenyl acetates respond to electron with-drawal and concomitant changes in hydrogen bond energy in a nearly identical fashion. Reference to thermodynamic studies of hydrogen bonding in free solution shows that these data are not consistent with hydrogen bond removal in the Michaelis complex and acyl enzyme but do support the alternative mechanism, in which hydrogen bonding of substrate is maintained along the entire reaction coordinate.  相似文献   

12.
Analogous soluble and insoluble derivatives of subtilisin Novo (EC 3.4.21.14) were prepared by coupling the enzyme to CNBr-activated DEAE-dextran and DEAE-Sephadex, respectively. The DEAE-dextran-subtilisin displayed pH optima and Km values for ester hydrolysis similar to subtilisin, whereas the pH versus activity profiles obtained with DEAE-Sephadex-subtilisin were shifter towards the alkaline pH region and the Km values were increased. Compared with subtilisin, DEAE-dextran-subtilisin showed a 40-65% reduction of kcat for hydrolysis of N-acetyl-L-tyrosine ethyl ester, p-tosyl-L-arginine methyl ester and benzyloxycarbonyl-glycyl-L-tyrosinamide and its maximum velocities for digestion of casein and clupein also amounted to 40-60% of the subtilisin values. With Deae-sephadex-subtilisin, in contrast, the maximum velocity of hydrolysis decreased to a greater extent for polypeptide substrates compared to ester substrates. The present results indicate that the chemical nature of a support can effect intrinsic properties of a matrix-bound enzyme in addition to the steric and diffusional effects usually observed with polymer-attached enzymes.  相似文献   

13.
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   

14.
Wang M  Boddy CN 《Biochemistry》2008,47(45):11793-11803
The final step in polyketide synthase-mediated biosynthesis of macrocyclic polyketides is thioesterase (TE)-catalyzed cyclization of a linear polyketide acyl chain. TEs are highly specific in the chemistry they catalyze. Understanding the molecular basis for substrate specificity of TEs is crucial for engineering these enzymes to macrocyclize non-native linear substrates. We investigated the role of hydrogen bonding interactions in the substrate specificity of formation of an acyl-enzyme intermediate for the TE from the 6-deoxyerythronolide B biosynthetic pathway. Thirteen single site-directed mutants were constructed, via removal of side chain hydrogen bonding groups from the binding cavity. Specificity constants for four different substrates with and without hydrogen bond donors and acceptors were determined for the five active mutants. The relative magnitude of specificity constants for substrates did not change for the mutant TEs. Circular dichroism spectroscopy was used to show that the majority of the catalytically inactive mutants did not fold. Two mutations were identified that enabled mutant TEs to form a folded but catalytically inactive tertiary structure. Our data do not support a role for hydrogen bonding in mediating substrate specificity of bacterial polyketide synthase TEs. The highly conserved polar residues in the binding cavity appear to stabilize the unusual substrate channel, which passes through the enzyme. We propose that hydrophobic interactions between the binding cavity and substrate drive substrate specificity, as is seen in many protein-carbohydrate recognition events. This hypothesis is in agreement with high-resolution structural data for nonhydrolyzable acyl-enzyme intermediates from the picromycin TE.  相似文献   

15.
We used substrate mapping to develop a rule that predicts which enantiomer of chiral carboxylic acid esters reacts faster in hydrolyses catalyzed by lipase from Candida rugosa (CRL, triacylglycerol hydrolase, E. C. 3.1.1.3). This rule, based on the size of the substituents at the stereocenter, is not reliable for crude CRL. It predicts the favoured enantiomer for only 23 out of 34 examples, 68% reliability. However, this rule is completely reliable for purified CRL; it predicts the favoured enantiomer for all 16 examples correctly. The examples include arylpropanoicacids, aryloxypropanoic acids, α-halophenylacetic acids, mandelic acid and O-methylmandelic acid. Further, purified CRL did not catalyse the hydrolysis of N-CBZ-phenylalanine methyl ester and N-CBZ-norleucine methyl ester. These two substrates were exceptions to the rule with crude CRL as the catalyst. Besides eliminating several exceptions, purification also raised the enantioselectivity of CRL toward carboxylic acid esters. To provide a structural basis for this proposed rule we examined the x-ray crystal structure of CRL containing transition state analogs of ester hydrolysis. We suggest that the large substituent of chiral carboxylic acids binds in a tunnel that normally binds the alkyl chain of a fatty acid. The phenyl rings of Phe 345 and Phe 415 lie close to the stereocenter, thereby fixing the orientation of the medium substituent. The three-dimensional orientation of these proposed binding sites is consistent with the rule derived from substrate mapping.  相似文献   

16.
Steady-state kinetic analysis has been used to confirm the catalytic mechanism of lyophilized subtilisin suspended in a variety of organic solvents. Specifically, this article demonstrates that partial reactions can occur between subtilisin and ester substrates in organic solvents. Partitioning of common intermediates between competing acceptors at a constant ratio of products has also been described. The decomposition of a common intermediate formed from different substrates at the same rate is also further evidence of an acyl-enzyme mechanism for subtilisin suspended in anhydrous solvents. Partitioning of a common intermediate to give two products at a constant total rate, and saturation kinetics at varying substrate concentrations, complete a kinetic investigation of the enzyme mechanism. All the data generated support the formation of a stable acyl enzyme during the transesterification reaction catalzyed by subtilisin in the solvents used.  相似文献   

17.
S Naderi  K Fotherby 《Steroids》1983,41(3):397-417
The hydrolysis of 108 esters of norethisterone (17 alpha-ethynyl-17 beta-hydroxyestr-4-en-3-one) was studied in vitro using a rabbit liver preparation. Introduction of a double or triple bond into a straight-chain ester did not inhibit hydrolysis but a marked reduction in hydrolysis was produced on replacement of a methylene group by an oxygen atom. Hydrolysis was inhibited by substituents at C2 of the ester chain except in short chain esters. Cyclopropylcarboxylate and cyclobutylcarboxylate were readily hydrolysed and introduction of a furan ring into the side-chain did not affect hydrolysis. No hydrolysis occurred with a cholesteryl carbonate ester or with a pentamethyldisilyloxy ether. Forty-nine esters of levonorgestrel (13 beta-ethyl-17 alpha-ethynyl-17 beta-hydroxygon-4-en-3-one) were also studied. In general, the pattern of hydrolysis for these esters was similar to that of the norethisterone esters. However, with few exceptions the levonorgestrel esters were hydrolyzed more slowly. For those esters for which information regarding the biological activity was available, there was no correlation between the potency of the esters and their rate of hydrolysis in vitro.  相似文献   

18.
An esterase was isolated from influenza C virus with a specific activity from 1.7-5 U/mg protein, and its substrate specificity was tested with various naturally occurring O-acylated sialic acids, synthetic carbohydrate acetates, and other esters. The enzyme hydrolyses only acetic acid esters at significant rates. The non-natural substrates 4-methyl-umbelliferyl acetate, 4-nitrophenyl acetate, and alpha-naphthyl acetate are cleaved at highest hydrolysis rates, followed by the natural substrate N-acetyl-9-O-acetylneuraminic acid. The esterase also acts on N-glycoloyl-9-O-acetylneuraminic acid and, much slower, on N-acetyl-4-O-acetylneuraminic acid; N-acetyl-7-O-acetylneuraminic acid is not hydrolysed. 2-Deoxy-2,3-didehydro-N-acetyl-9-O-acetylneuraminic acid is also a substrate for this enzyme, however, 6-O-acetylated N-acetylmannosamine and glucose are not. Esterification of the carboxyl function of sialic acids strongly reduces or prevents esterase action on O-acetyl groups. The carboxyl ester is not hydrolysed. The relative cleavage rates also depend on the type of the non-sialic acid part of the molecule. N-Acetyl-9-O-acetylneuraminic acid as component of sialyllactose and rat serum glycoprotein shows hydrolysis rates close to the free form of this sugar, while acetyl ester groups of bovine submandibular gland mucin and rat erythrocytes are hydrolysed at slower rates. Gangliosides and 4-O-acetylated glycoproteins are no substrates for the purified enzyme. A slow hydrolysis is observed by incubation of 9-O-acetylated GD1a with intact influenza C viruses. As other natural acetyl esters (acetyl-CoA and acetylthiocholine iodide) are not hydrolysed, the enzyme can be classified as sialate 9(4)-O-acetylesterase (EC 3.1.1.53).  相似文献   

19.
A C Storer  P R Carey 《Biochemistry》1985,24(24):6808-6818
The kinetic constants for the papain-catalyzed hydrolysis of the methyl thiono esters of N-benzoylglycine and N-(beta-phenylpropionyl)glycine are compared with those for the corresponding methyl ester substrates. The k2/Ks values for the thiono esters are 2-3 times higher than those for the esters, and both show bell-shaped pH dependencies with similar pKa's (approximately 4 and 9). The k3 values for the thiono esters are 30-60 times less than those for the esters and do not exhibit a pH dependency. Solvent deuterium isotope effects on k2/Ks and k3 were measured for the ester and thiono ester substrates of both glycine derivatives. Each thiono ester substrate showed an isotope effect similar to that for the corresponding ester substrate. Moreover, use of the proton inventory technique indicated that, as for esters, one proton is transferred in the transition state for deacylation during reactions involving thiono esters and the degree of heavy atom reorganization in the transition state is very similar in both cases. The k3 values for the hydrolysis of a series of para-substituted N-benzoylglycine esters were found to correlate with the k3 values for the corresponding para-substituted thiono esters [Carey, P. R., Lee, H., Ozaki, Y., & Storer, A. C. (1984) J. Am. Chem. Soc. 106, 8258-8262], showing that the rate-determining step for the deacylation of both thiolacyl and dithioacyl enzymes probably involves the disruption of a contact between the substrate's glycinic nitrogen atom and the sulfur of cysteine-25. It is concluded that the hydrolysis of esters and thiono esters proceeds by essentially the same reaction pathway. Due to an oxygen-sulfur exchange process the product released in the case of the N-(beta-phenylpropionyl)glycine thiono ester substrate is the dioxygen acid; however, for the N-benzoylglycine thiono ester substrate, the thiol acid is the initial product. This thiol acid then acts as a substrate for papain and reacylates the enzyme to eventually give the dioxygen acid product. It is shown that thiol acids are excellent substrates for papain.  相似文献   

20.
Contrary to common protease substrates, the hydrolysis of 4-guanidinophenyl esters of the Boc-Xaa-OGp type by trypsin and trypsin-like proteases performs easily and independently of the structure and chirality of the acyl moiety. The hydrolysis of this new class of substrate mimetics, previously called inverse substrates, is enabled by the highly specific leaving group. However, the mechanism cannot be explained on the basis of the conventional binding model which defines the interactions between the protease and its substrate. Hydrolysis and aminolysis kinetics, protein-ligand docking, and molecular dynamics studies have been carried out in order to get insight into the catalytic mechanism which holds for these substrate mimetics. The experimental and theoretical results obtained for the serine protease trypsin suggest a novel extended kinetic model. It explains the hydrolysis of these types of protease substrates and accounts for the structural consequences for their aminolysis.  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