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 257 毫秒
1.
Despite the availability of many experimental data and some modeling studies, questions remain as to the precise mechanism of the serine proteases. Here we report molecular dynamics simulations on the acyl-enzyme complex and the tetrahedral intermediate during the deacylation step in elastase catalyzed hydrolysis of a simple peptide. The models are based on recent crystallographic data for an acyl-enzyme intermediate at pH 5 and a time-resolved study on the deacylation step. Simulations were carried out on the acyl enzyme complex with His-57 in protonated (as for the pH 5 crystallographic work) and deprotonated forms. In both cases, a water molecule that could provide the nucleophilic hydroxide ion to attack the ester carbonyl was located between the imidazole ring of His-57 and the carbonyl carbon, close to the hydrolytic position assigned in the crystal structure. In the "neutral pH" simulations of the acyl-enzyme complex, the hydrolytic water oxygen was hydrogen bonded to the imidazole ring and the side chain of Arg-61. Alternative stable locations for water in the active site were also observed. Movement of the His-57 side-chain from that observed in the crystal structure allowed more solvent waters to enter the active site, suggesting that an alternative hydrolytic process directly involving two water molecules may be possible. At the acyl-enzyme stage, the ester carbonyl was found to flip easily in and out of the oxyanion hole. In contrast, simulations on the tetrahedral intermediate showed no significant movement of His-57 and the ester carbonyl was constantly located in the oxyanion hole. A comparison between the simulated tetrahedral intermediate and a time-resolved crystallographic structure assigned as predominantly reflecting the tetrahedral intermediate suggests that the experimental structure may not precisely represent an optimal arrangement for catalysis in solution. Movement of loop residues 216-223 and P3 residue, seen both in the tetrahedral simulation and the experimental analysis, could be related to product release. Furthermore, an analysis of the geometric data obtained from the simulations and the pH 5 crystal structure of the acyl-enzyme suggests that since His-57 is protonated, in some aspects, this crystal structure resembles the tetrahedral intermediate.  相似文献   

2.
P J Tonge  P R Carey 《Biochemistry》1990,29(48):10723-10727
Resonance Raman (RR) spectroscopy has been used to obtain the vibrational spectrum of the acyl carbonyl group in a series of acylchymotrypsins and acylsubtilisins at the pH of optimum hydrolysis. The acyl-enzymes, which utilize arylacryloyl acyl groups, include three oxyanion hole mutants of subtilisin BPN', Asn155Leu, Asn155Gln, and Asn155Arg, and encompass a 500-fold range of deacylation rate constants. For each acyl-enzyme a RR carbonyl band has been identified which arises from a population of carbonyl groups undergoing nucleophilic attack in the active site. As the deacylation rate (k3) increases through the series of acyl-enzymes, the carbonyl stretching band (vC = O) is observed to shift to lower frequency, indicating an increase in single bond character of the reactive acyl carbonyl group. Experiments involving the oxyanion hole mutants of subtilisin BPN' indicate that a shift of vC = O to lower frequency results from stronger hydrogen bonding of the acyl carbonyl group in the oxyanion hole. A plot of log k3 against vC = O is linear over the range investigated, demonstrating that the changes in vC = O correlate with the free energy of activation for the deacylation reaction. By use of an empirical correlation between carbonyl frequency (vC = O) and carbonyl bond length (rC = O) it is estimated that rC = O increases by 0.015 A as the deacylation rate increases 500-fold through the series of acyl-enzymes. This change in rC = O is about 7% of that expected for going from a formal C = O double bond in the acyl-enzyme to a formal C-O single bond in the tetrahedral intermediate for deacylation. The data also allow us to estimate the energy needed to extend the acyl carbonyl group along its axis to be 950 kJ mol-1 A-1.  相似文献   

3.
The rate constants for hydrolysis of the enantiomers of amino acid p-nitrophenyl esters catalyzed by bifunctional comicellar catalysts containing the imidazolyl and hydroxyl groups have been determined at pH 7.30, 0.02 m phosphate buffer, and 25°C. The kinetic analysis suggests a reaction scheme which involves acylation followed by deacylation at the imidazolyl group. Although no appreciable cooperative catalytic efficiencies are observed between the bifunctional groups in the acylation step, it is found that the deacylation rates are thus accelerated by surfactant hydroxyl groups, and some of the stereoselective acyl transfer reaction occurs from the imidazolyl to the hydroxyl group in optically active comicellar systems.  相似文献   

4.
The overall chemical mechanism of the reaction catalyzed by the hydrolytic aldehyde dehydrogenases (ALDHs) involves three main steps: (1) nucleophilic attack of the thiol group of the catalytic cysteine on the carbonyl carbon of the aldehyde substrate; (2) hydride transfer from the tetrahedral thiohemiacetal intermediate to the pyridine ring of NAD(P)(+); and (3) hydrolysis of the resulting thioester intermediate (deacylation). Crystal structures of different ALDHs from several organisms-determined in the absence and presence of bound NAD(P)(+), NAD(P)H, aldehydes, or acid products-showed specific details at the atomic level about the catalytic residues involved in each of the catalytic steps. These structures also showed the conformational flexibility of the nicotinamide half of the cofactor, and of the catalytic cysteinyl and glutamyl residues, the latter being the general base that activates the hydrolytic water molecule in the deacylation step. The architecture of the ALDH active site allows for this conformational flexibility, which, undoubtedly, is crucial for catalysis in these enzymes. Focusing in the deacylation step of the ALDH-catalyzed reaction, here we review and systematize the crystallographic evidence of the structural features responsible for the conformational flexibility of the catalytic glutamyl residue, and for the positioning of the hydrolytic water molecule inside the ALDH active site. Based on the analysis of the available crystallographic data and of energy-minimized models of the thioester reaction intermediate, as well as on the results of theoretical calculations of the pK(a) of the carboxyl group of the catalytic glutamic acid in its three different conformations, we discuss the role that the conformational flexibility of this residue plays in the activation of the hydrolytic water. We also propose a critical participation in the water activation process of the peptide bond to which the catalytic glutamic acid in the intermediate conformation is hydrogen bonded.  相似文献   

5.
The imidazole of chromophoric p-(dimethylamino)benzoic acid, DABIm, reacts with the serine protease alpha-chymotrypsin in the pH range of 4-7 to form a stable acyl intermediate that gives very good resonance-enhanced Raman spectra. The resonance Raman and absorption spectra of the acyl enzyme intermediate have been compared with the spectra of simple model compounds such as the corresponding chromophoric methyl ester, aldehyde, and imidazole. The resonant Raman and ultraviolet absorption spectra of these simple chromophoric model compounds change considerably with the solvent. However, each of the model compounds exhibits a linear correlation between the maximum wavelength of absorption and the frequency of the carbonyl vibration. The observed values of the acyl intermediate do not fall on the line for the methyl ester but rather on the line for the aldehyde. This shows that the chromophoric serine ester of the acyl enzyme behaves differently than an ordinary ester, which cannot be explained as a solvent effect. Thermal unfolding of the acyl enzyme brings the spectroscopic parameters close to those of the model ester. We conclude that it is the specific conformation of the native enzyme and not solvent effects that change the spectroscopic properties of the acyl chromophore. It is reasonable that these changes arise from the same forces that cause the catalytic events. The carbonyl frequencies of a series of para-substituted benzoyl methyl esters show a remarkably linear correlation with the rate of deacylation of the corresponding acyl enzymes.  相似文献   

6.
Lys67 is essential for the hydrolysis reaction mediated by class C β‐lactamases. Its exact catalytic role lies at the center of several different proposed reaction mechanisms, particularly for the deacylation step, and has been intensely debated. Whereas a conjugate base hypothesis postulates that a neutral Lys67 and Tyr150 act together to deprotonate the deacylating water, previous experiments on the K67R mutants of class C β‐lactamases suggested that the role of Lys67 in deacylation is mainly electrostatic, with only a 2‐ to 3‐fold decrease in the rate of the mutant vs the wild type enzyme. Using the Class C β‐lactamase AmpC, we have reinvestigated the activity of this K67R mutant enzyme, using biochemical and structural studies. Both the rates of acylation and deacylation were affected in the AmpC K67R mutant, with a 61‐fold decrease in kcat, the deacylation rate. We have determined the structure of the K67R mutant by X‐ray crystallography both in apo and transition state‐analog complexed forms, and observed only minimal conformational changes in the catalytic residues relative to the wild type. These results suggest that the arginine side chain is unable to play the same catalytic role as Lys67 in either the acylation or deacylation reactions catalyzed by AmpC. Therefore, the activity of this mutant can not be used to discredit the conjugate base hypothesis as previously concluded, although the reaction catalyzed by the K67R mutant itself likely proceeds by an alternative mechanism. Indeed, a manifold of mechanisms may contribute to hydrolysis in class C β‐lactamases, depending on the enzyme (wt or mutant) and the substrate, explaining why different mutants and substrates seem to support different pathways. For the WT enzyme itself, the conjugate base mechanism may be well favored.  相似文献   

7.
The adenosine monoposphate‐forming acyl‐CoA synthetase enzymes catalyze a two‐step reaction that involves the initial formation of an acyl adenylate that reacts in a second partial reaction to form a thioester between the acyl substrate and CoA. These enzymes utilize a Domain Alternation catalytic mechanism, whereby a ~110 residue C‐terminal domain rotates by 140° to form distinct catalytic conformations for the two partial reactions. The structure of an acetoacetyl‐CoA synthetase (AacS) is presented that illustrates a novel aspect of this C‐terminal domain. Specifically, several acetyl‐ and acetoacetyl‐CoA synthetases contain a 30‐residue extension on the C‐terminus compared to other members of this family. Whereas residues from this extension are disordered in prior structures, the AacS structure shows that residues from this extension may interact with key catalytic residues from the N‐terminal domain. Proteins 2015; 83:575–581. © 2014 Wiley Periodicals, Inc.  相似文献   

8.
By using a flow system, resonance Raman spectra have been obtained of the acyl enzyme 4-amino-3-nitro-cinnamoyl-α-chymotrypsin over the pH range where it shifts from a stable to an active state. The band at 1625 cm?1 was assigned to the CC stretch in the acryloyl residue
and bands in the 1200 to 1450 cm?1 region were shown to be associated with the aromatic residue. Studies with analogues of the substrate established that the CC and the aromatic bands could independently monitor events in the
and aromatic moieties, respectively. Thus, events in the acyl enzyme leading to spectral changes involving the catalytic site could be distinguished from those involving the aromatic site. The data strongly suggested that the structure of the acylating group in the stable acyl enzyme at pH 3.0 resembled that of the substrate in aqueous solution, the conformation about the ethylenic bonds being essentially planar, trans and probably s-trans about the CCCO single bond. However, on shifting the acyl enzyme to pH values of 5.7 to 7.0 a large change occurred in the 1625 cm?1 band without detectable changes in the aromatic bands. Thus, the enzyme produces a change in a submolecular grouping which contains the bond being hydrolysed. The change occurred within the mixing and observation time of about ten seconds. Since it was identical at essentially inactive (pH 5.7) and active (pH 7.0) pH values it must precede and be independent of the rate-controlling step in deacylation. The change in the 1625 cm?1 band thus probably reflects an ionization in the acyl enzyme between pH 3.0 and 5.7. Consequently, above pH 5.7 deacylation proceeds from a structure differing from that at the stable pH of 3.0. The resonance Raman spectra of this intermediate are consistent with either deformation to reduce planarity within the acryloyl residue, or of the stabilization of resonance structures of the form
Either of these mechanisms would produce an intermediate, the carbonyl group of which would be probably intrinsically more reactive relative to the substrate in solution or to the stable acyl enzyme. The resonance Raman provides some support for Henderson's (1970) prediction that distortion occurs in the acryloyl residue prior to deacylation.  相似文献   

9.
Benzyloxycarbonyl-L-alanine p-guanidinophenyl ester behaves as a trypsin "inverse substrate," i.e., a cationic center is included in the leaving group instead of being in the acyl moiety. Using this substrate as an acyl donor, trypsin catalyzes the synthesis of peptide bonds that cannot be split by this enzyme. An optimal acyl transfer efficiency was achieved between pH 8 and 9 at 30 degrees C.The addition of as much as 50% cosolvent was shown to be of minor influence on the acyl transfer efficiency, whereas the reaction velocity decreases by more than one order of magnitude. The efficiency of H-Leu-NH(2) and H-Val-NH(2) in deacylation is almost the same for "inverse" and normal type substrates.  相似文献   

10.
A novel strategy for a more efficient synthesis of difficult sequence‐containing peptides, the S‐acyl isopeptide method, was developed and successfully applied. A model pentapeptide Ac–Val–Val–Cys–Val–Val–NH2 was synthesized via its water‐soluble S‐acyl isopeptide using an S‐acyl isodipeptide unit, Boc–Cys(Fmoc–Val)–OH. An S‐acyl isopeptide possessing excellent water solubility could be readily and quantitatively converted to the native peptide via an SN intramolecular acyl migration reaction at pH 7.4. Thus, the S‐acyl isopeptide method provides a useful tool in peptide chemistry. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
P J Tonge  P R Carey 《Biochemistry》1989,28(16):6701-6709
By use of resonance Raman (RR) spectroscopy, the population of the reactive carbonyl group in active acylchymotrypsins has been characterized and correlated with acyl-enzyme reactivity. RR spectra have been obtained, with a flow system and 324- and 337.5-nm excitation, at low and active pH for six acylchymotrypsins, viz., (indoleacryloyl)-, (4-amino-3-nitrocinnamoyl)-, (furylacryloyl)-, [( 5-ethylfuryl)-acryloyl]-, (thienylacryloyl)-, and [( 5-methylthienyl)acryloyl]chymotrypsin. These acyl-enzymes represent a 100-fold range of deacylation rate constants. Good RR spectral quality has enabled us to obtain the vibrational spectrum of the carbonyl group at low and active pH in each acyl-enzyme. The measured pKa of the spectroscopic changes in the carbonyl region is identical with that for the deacylation kinetics, showing that the RR carbonyl features reflect the ionization state of His-57. A carbonyl population has been observed in the active acyl-enzymes in which the carbonyl oxygen atom of the reactive acyl linkage is hydrogen-bonded in the active site. The proportion of this hydrogen-bonded population, with respect to other observed non-hydrogen-bonded species, together with the degree of polarization of the carbonyl bond, as monitored by vC = 0, has been correlated with the deacylation rate constants of the acyl-enzymes. It is proposed that the hydrogen-bonded carbonyl species is located at or near the oxyanion hole and represents the ground state from which deacylation occurs. An increase in the proportion of the hydrogen-bonded population and an increase in polarization of the carbonyl bond result in an increase in deacylation rate constant.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
5‐Aminolevulinate synthase (ALAS) controls the rate‐limiting step of heme biosynthesis in mammals by catalyzing the condensation of succinyl‐coenzyme A and glycine to produce 5‐aminolevulinate, coenzyme‐A (CoA), and carbon dioxide. ALAS is a member of the α‐oxoamine synthase family of pyridoxal 5′‐phosphate (PLP)‐dependent enzymes and shares high degree of structural similarity and reaction mechanism with the other members of the family. The X‐ray crystal structure of ALAS from Rhodobacter capsulatus reveals that the alkanoate component of succinyl‐CoA is coordinated by a conserved arginine and a threonine. The functions of the corresponding acyl‐CoA‐binding residues in murine erthyroid ALAS (R85 and T430) in relation to acyl‐CoA binding and substrate discrimination were examined using site‐directed mutagenesis and a series of CoA‐derivatives. The catalytic efficiency of the R85L variant with octanoyl‐CoA was 66‐fold higher than that of the wild‐type protein, supporting the proposal of this residue as key in discriminating substrate binding. Substitution of the acyl‐CoA‐binding residues with hydrophobic amino acids caused a ligand‐induced negative dichroic band at 420 nm in the CD spectra, suggesting that these residues affect substrate‐mediated changes to the PLP microenvironment. Transient kinetic analyses of the R85K variant‐catalyzed reactions confirm that this substitution decreases microscopic rates associated with formation and decay of a key reaction intermediate and show that the nature of the acyl‐CoA tail seriously affect product binding. These results show that the bifurcate interaction of the carboxylate moiety of succinyl‐CoA with R85 and T430 is an important determinant in ALAS function and may play a role in substrate specificity.  相似文献   

13.
Penicillin-binding protein 5 (PBP 5) from Escherichia coli is a well-characterized d-alanine carboxypeptidase that serves as a prototypical enzyme to elucidate the structure, function, and catalytic mechanism of PBPs. A comprehensive understanding of the catalytic mechanism underlying d-alanine carboxypeptidation and antibiotic binding has proven elusive. In this study, we report the crystal structure at 1.6 A resolution of PBP 5 in complex with a substrate-like peptide boronic acid, which was designed to resemble the transition-state intermediate during the deacylation step of the enzyme-catalyzed reaction with peptide substrates. In the structure of the complex, the boron atom is covalently attached to Ser-44, which in turn is within hydrogen-bonding distance to Lys-47. This arrangement further supports the assignment of Lys-47 as the general base that activates Ser-44 during acylation. One of the two hydroxyls in the boronyl center (O2) is held by the oxyanion hole comprising the amides of Ser-44 and His-216, while the other hydroxyl (O3), which is analogous to the nucleophilic water for hydrolysis of the acyl-enzyme intermediate, is solvated by a water molecule that bridges to Ser-110. Lys-47 is not well-positioned to act as the catalytic base in the deacylation reaction. Instead, these data suggest a mechanism of catalysis for deacylation that uses a hydrogen-bonding network, involving Lys-213, Ser-110, and a bridging water molecule, to polarize the hydrolytic water molecule.  相似文献   

14.
The motions of water molecules, the acyl moiety, the catalytic triad, and the oxyanion binding site of acyl-chymotrypsin were studied by means of a stochastic boundary molecular dynamics simulation. A water molecule that could provide the nucleophilic OH? for the deacylation stage of the catalysis was found to be trapped between the imidazole ring of His-57 and the carbonyl carbon of the acyl group. It makes a hydrogen bond with the Nε2 of His-57 and is heldin place through a network of hydrogen-bonded water molecules in theactive site. The water molecule was found as close as 2.8 Å to the carbonyl carbon. This appears to be due to the constraints imposed by nonbonded interaction in the active site. Configurations were found in which one hydrogen of the trapped water shared a bifurcated hydrogen bond with His-57-Nε2 and Ser-195-0γ with the water oxygen very close to the carbonyl carbon. The existence of such a water molecule suggests that large movement of the His-57 imidazole ring between positions suitable for providing general-base catalyzed assistance and for providing general-acid catalyzed assistance may notbe required during the reaction. The simulation indicates that the side chains of residues involved in catalysis (i.e., His-57, Ser-195, and Asp-102) are significantly less flexible than other side chains in the protein. The 40% reduction in rms fluctuations is consistent with a comparable reduction calculated from the temperature factors obtained in the X-ray crystal-lographic data of γ-chymotrypsin. The greater rigidity of active site residues seems to result from interconnected hydrogen bonding networks among the residues and between the residues and the solvent water in the active site. © Wiley-Liss, Inc.  相似文献   

15.
Cell-free preparations of Chlorella pyrenoidosa catalyze the transfer of the fatty acyl moiety of fatty acyl CoA derivatives to sulfoquinovosyl monoglyceride to form sulfoquinovosyl diglyceride. This reaction is stimulated by Triton X-100 concentrations of up to 0.6 mg/ml and has a pH optimum of 7.7. Similar Chlorella preparations catalyze the stepwise removal of both fatty acyl groups from sulfoquinovosyl diglyceride to form sulfoquinovosyl monoglyceride and then sulfoquinovosyl glycerol. This reaction is inhibited by both calcium and magnesium. The nonionic surfactant Triton X-100 inhibits the enzymatic deacylation at concentrations of less than 0.5 mg/ml but stimulates it at higher concentrations. The pH optimum for the deacylation of sulfoquinovosyl glycerides is 8.2, with little activity observed below pH 8. The enzymatic activities for both the transacylation and deacylation reactions are associated with a 30,000 g particulate fraction of Chlorella. Sulfoquinovosyl glycerol was found not to be an acceptor of the fatty acyl moiety of fatty acyl CoA derivatives. Methods are described for the preparation of sulfoquinovosyl monoglyceride, sulfoquinovose, and 3-sulfo-1,2-propanediol.  相似文献   

16.
Cycloheptaamylose cinnamate, an intermediate in the hydrolysis of m-nitrophenyl cinnamate by cycloheptaamylose, was isolated in pure form. The deacylation of acyl-cycloamyloses (cinnamate and acetate) catalyzed by noncovalently complexed 6-nitrobenzimidazole (1) was studied. The reaction was enzyme-like. Saturation of acyl-cycloamylose by 1 was observed; the rate and dissociation constants were determined from Lineweaver-Burk plots. The catalyzed reaction rates at neutral pH were two to three times larger than those of the spontaneous reactions for cycloheptaamylose or cyclohexaamylose cinnamate, respectively. The catalytic effect of 1 on the deacylation rate of cyclohexaamylose cinnamate became smaller as the pH of the solution was raised. The deacylation of cyclohexaamylose acetate was followed by nmr spectroscopy, whereas the deacylation of cycloamylose cinnamates was followed by uv spectroscopy and extraction of trans-cinnamic acid with ether. Thermodynamic parameters for the rates of deacylation of cycloamylose cinnamates and dissociation constants of cycloamylose cinnamate-1 complexes were obtained and discussed.  相似文献   

17.
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   

18.
The acyl‐AMP forming family of adenylating enzymes catalyze two‐step reactions to activate a carboxylate with the chemical energy derived from ATP hydrolysis. X‐ray crystal structures have been determined for multiple members of this family and, together with biochemical studies, provide insights into the active site and catalytic mechanisms used by these enzymes. These studies have shown that the enzymes use a domain rotation of 140° to reconfigure a single active site to catalyze the two partial reactions. We present here the crystal structure of a new medium chain acyl‐CoA synthetase from Methanosarcina acetivorans. The binding pocket for the three substrates is analyzed, with many conserved residues present in the AMP binding pocket. The CoA binding pocket is compared to the pockets of both acetyl‐CoA synthetase and 4‐chlorobenzoate:CoA ligase. Most interestingly, the acyl‐binding pocket of the new structure is compared with other acyl‐ and aryl‐CoA synthetases. A comparison of the acyl‐binding pocket of the acyl‐CoA synthetase from M. acetivorans with other structures identifies a shallow pocket that is used to bind the medium chain carboxylates. These insights emphasize the high sequence and structural diversity among this family in the area of the acyl‐binding pocket. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

19.
Chromophoric [4-(dimethylamino)cinnamoyl]imidazole reacts with the serine protease alpha-chymotrypsin to form an acyl enzyme. At pHs below 4.0, the acyl enzyme turns over very slowly to yield the free acid. During this slow deacylation it is possible to obtain a very good resonance Raman spectrum of the acyl intermediate by using the 350.7-nm line of the krypton laser. The resonance Raman carbonyl frequency of the covalently bonded substrate and its wavelength at maximum intensity in the absorption spectrum of the acyl enzyme have been taken and used to monitor the active site environment. A comparison has been made of the absorption and Raman spectra of the acyl enzyme and those of the corresponding chromophoric methyl ester, aldehyde, and imidazole model compounds. A linear correlation is found between the wavelength of maximum absorption and the Raman frequency of the carbonyl group over a wide range of solvent conditions for each of the model compounds. By combining the Raman carbonyl frequency with the absorption maximum, we can determine that the bond order changes in the carbonyl bond of the bound substrate are not due to changes in the solvent, since the carbonyl frequency and the absorption maximum of the acyl enzyme do not fall on any of the linear correlations for the model compounds. The unusual spectroscopic properties of the bound substrate appear to be due to some specific enzyme-induced change in the substrate when it is bound at the active site. Thermal unfolding of the acyl enzymes changes both the carbonyl frequency of the acyl enzyme and its absorption maximum to completely different values.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
Serpin family protein proteinase inhibitors trap proteinases at the acyl-intermediate stage of cleavage of the serpin as a proteinase substrate by undergoing a dramatic conformational change, which is thought to distort the proteinase active site and slow deacylation. To investigate the extent to which proteinase catalytic function is defective in the serpin-proteinase complex, we compared the pH dependence of dissociation of several serpin-proteinase acyl-complexes with that of normal guanidinobenzoyl-proteinase acyl-intermediate complexes. Whereas the apparent rate constant for dissociation of guanidinobenzoyl-proteinase complexes (k(diss, app)) showed a pH dependence characteristic of His-57 catalysis of complex deacylation, the pH dependence of k(diss, app) for the serpin-proteinase complexes showed no evidence for His-57 involvement in complex deacylation and was instead characteristic of a hydroxide-mediated deacylation similar to that observed for the hydrolysis of tosylarginine methyl ester. Hydroxylamine enhanced the rate of serpin-proteinase complex dissociation but with a rate constant for nucleophilic attack on the acyl bond several orders of magnitude slower than that of hydroxide, implying limited accessibility of the acyl bond in the complex. The addition of 10-100 mm Ca(2+) ions stimulated up to 80-fold the dissociation rate constant of several serpin-trypsin complexes in a saturable manner at neutral pH and altered the pH dependence to a pattern characteristic of His-57-catalyzed complex deacylation. These results support a mechanism of kinetic stabilization of serpin-proteinase complexes wherein the complex is trapped as an acyl-intermediate by a serpin conformational change-induced inactivation of the proteinase catalytic function, but suggest that the inactive proteinase conformation in the complex is in equilibrium with an active proteinase conformation that can be stabilized by the preferential binding of an allosteric ligand such as Ca(2+).  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