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.
By using two reactive analogues of UDP-Glc, uridine di- and triphosphopyridoxals, we have recently probed the substrate-binding site in potato tuber UDP-Glc pyrophosphorylase [EC 2.7.7.9]. In this work, pyridoxal diphospho-alpha-D-glucose was used for the same purpose. This compound is also a reactive UDP-Glc analogue but having its reactive group on the opposite side of the pyrophosphate linkage to those of the above two compounds. The enzyme was rapidly inactivated when incubated with the compound at very low concentrations followed by reduction with sodium borohydride. The inactivation was almost completely prevented by UDP-Glc and UTP. Complete inactivation correspond to the incorporation of 1.0 mol of the reagent per mol of enzyme monomer. The label was found to be distributed in five lysyl residues (Lys-263, Lys-329, Lys-367, Lys-409, and Lys-40. All of these results were similar to those obtained previously with the other compounds, suggesting the presence of a cluster of five lysyl residues at or near the substrate-binding site of this enzyme. However, the incorporations of labels into each lysyl residue differed depending on the compounds used. The substrate retarded the incorporations in different manners. Based on the combined results of the present and previous studies, a hypothetical model is presented for the possible locations of the five lysyl residues around the substrate bound to the enzyme. This model is consistent with the kinetic properties of mutant enzymes in which the five lysyl residues were individually replaced by glutamine via site-directed mutagenesis.  相似文献   

2.
Terpenes comprise the largest class of natural products currently known. These ubiquitous molecules are synthesized by terpene synthases via complex carbocationic reactions, incorporating highly reactive intermediates. In the current study, we present a mechanistic investigation of the biosynthetic pathway for the formation of selina-4(15),7(11)-diene. We employ density functional theory to study a model carbocation system in the gas-phase, and delineate the energetic feasibility of a plausible reaction path. Our results suggests that during formation of selina-4(15),7(11)-diene, the substrate is likely folded in a conformation conducive to sequential cyclizations. We propose that a required proton transfer cannot occur intramolecularly in the gas-phase due to a high free energy barrier, and that enzyme assistance is essential for this step. Hybrid quantum mechanics-molecular mechanics docking studies suggest that enzyme intervention could be realized through electrostatic guidance.  相似文献   

3.
We have found that a chicken egg lysozyme derivative (beta-101-lysozyme) containing an L-isoaspartyl residue at position 101 has a Km for methylation by the human erythrocyte L-isoaspartyl/D-aspartyl protein methyltransferase (EC 2.1.1.77) of 183 microM, about 30 times higher than that expected from previous studies with isoaspartyl-containing peptides. In the course of investigating the reasons for this poor enzyme recognition, we found that charged residues on the carboxyl side of isoaspartyl residues had a large effect on the affinity of the enzyme for synthetic peptides. This is best illustrated by the lysozyme-related peptide YVSisoDGDG, which has a Km for methylation of 469 microM. When the penultimate aspartyl residue is replaced by a cysteinyl residue, the Km drops to 4.6 microM, comparable to other peptides of similar size. Furthermore, replacing it with a cysteic acid residue results in a Km of 104 microM, suggesting that a negative charge at this position may lead to a weaker affinity of the peptide substrate for the methyltransferase. Assays with additional synthetic peptides indicate that moving the negative charge to the first or third residue on the carboxyl side of the isoaspartyl residue has a similar but less severe effect in reducing its affinity for the methyltransferase. Enzymatic methylation has recently been proposed to be the first step in the conversion of abnormal isoaspartyl residues to aspartyl residues. The results reported here, however, along with previous evidence that protein tertiary structure can inhibit isoaspartyl methylation, suggest that only a subclass of damaged sites are capable of efficiently entering a putative repair pathway; the sites not recognized by the methyltransferase may accumulate in vivo.  相似文献   

4.
Tilocca A  Gamba A  Vanoni MA  Fois E 《Biochemistry》2002,41(48):14111-14121
Large-scale Car-Parrinello molecular dynamics simulations of D-alanine oxidation catalyzed by the flavoenzyme D-amino acid oxidase have been carried out. A model of the enzyme active site was built by starting from the enzyme X-ray structure, and by testing different subsystems comprising different sets of aminoacyl residues. In this process, the stability of the enzyme-substrate complex was taken as a measure of the accuracy of the model. The activated transfer of the amino acid alpha-hydrogen from the substrate to the flavin N5 position was then induced by constraining a suitable transfer reaction coordinate, and the free energy profile of the reaction was calculated. The evolution of electronic and structural properties of both enzyme-bound substrate and flavin cofactor along the reaction path is consistent with a hydride-transfer mechanism. The calculated free energy barrier for this process (13 kcal/mol) is in excellent agreement with the activation energy value derived from the experimentally determined rate constant for the corresponding enzyme-catalyzed reaction. The electronic distribution of the reduced flavin shows that the transferred electrons tend to be centered near the C4a position rather than delocalized over the flavin pyrimidine ring. This feature is mechanistically relevant in that such an electronic distribution may promote the subsequent enzyme-catalyzed reduction of molecular oxygen to yield hydrogen peroxide via a postulated flavin 4a-peroxide intermediate. These results also show that a first-principles molecular dynamics approach is suitable to study the mechanism of complex enzymatic processes, provided that a smaller, yet reliable, subsystem of the enzyme can be identified, and special computational techniques are employed to enhance the sampling of the reactive event.  相似文献   

5.
Wu N  Gillon W  Pai EF 《Biochemistry》2002,41(12):4002-4011
The crystal structures of orotidine 5'-monophosphate decarboxylases from four different sources have been published recently. However, the detailed mechanism of catalysis of the most proficient enzyme known to date remains elusive. As the ligand-protein interactions at the orotate binding site are crucial to the understanding of this enzyme, we mutated several of the residues surrounding the aromatic part of the substrate, individually and in combination. The ensuing effects on enzyme structure and stability were characterized by X-ray crystallography of inhibitor, product, or substrate complexes and by chemical denaturation with guanidine hydrochloride, respectively. The results are consistent with the residues K42D70K72D75B being charged and forming an 'alternate charge network' around the reactive part of the substrate. In addition to exerting charge-charge repulsion on the orotate carboxylate, Asp70 also makes a crucial contribution to enzyme stability. Consequently, orotidine 5'-monophosphate decarboxylases seem to require the presence of a negative charge at this position for catalysis as well as for correct and stable folding.  相似文献   

6.
Alginate is a major constituent of mature biofilms produced by Pseudomonas aeruginosa. The penultimate step in the biosynthesis of alginate is the conversion of some beta-D-mannuronate residues in the polymeric substrate polymannuronan to alpha-L-guluronate residues in a reaction catalyzed by C5-mannuronan epimerase. Specificity studies conducted with size-fractionated oligomannuronates revealed that the minimal substrate contained nine monosaccharide residues. The maximum velocity of the reaction increased from 0.0018 to 0.0218 s(-1) as the substrate size increased from 10 to 20 residues, and no additional increase in kcat was observed for substrates up to 100 residues in length. The Km decreased from 80 microM for a substrate containing fewer than 15 residues to 4 microM for a substrate containing more than 100 residues. In contrast to C5-mannuronan epimerases that have been characterized in other bacterial species, P. aeruginosa C5-mannuronan epimerase does not require Ca2+ for activity, and the Ca2+-alginate complex is not a substrate for the enzyme. Analysis of the purified, active enzyme by inductively coupled plasma-emission spectroscopy revealed that no metals were present in the protein. The pH dependence of the kinetic parameters revealed that three residues on the enzyme which all have a pKa of approximately 7.6 must be protonated for catalysis to occur. The composition of the polymeric product of the epimerase reaction was analyzed by 1H NMR spectroscopy, which revealed that tracts of adjacent guluronate residues were readily formed. The reaction reached an apparent equilibrium when the guluronate composition of the polymer was 75%.  相似文献   

7.
O'Brien G  Quinsey NS  Whisstock JC  Pike RN 《Biochemistry》2003,42(50):14939-14945
The classical complement pathway, which plays a vital role in preventing infection, is initiated by the action of the serine proteases C1r and C1s. We have examined the hydrolysis of substrates representing cleavage sequences in the physiological substrates for C1s, C2 and C4. These studies showed that the P(1)'-P(4)' substrate residues of C2 and C4 conferred greater affinity of substrate for enzyme and also induced a sigmoidal dependence of enzyme velocity on substrate concentration. This indicates that the substrate gave rise to homotropic positive cooperative behavior in the enzyme. When C1s was in complex with C1q and C1r, as would occur under physiological conditions, the same behavior was observed, indicating that this mechanism is relevant in the complement pathway in vivo. We further investigated the requirements of C1s for prime side amino acids by examining a substrate library in which each of the P(1)'-P(4)' positions had been substituted by different classes of amino acids. This revealed that the P(1)' position was a major determinant of the selectivity of the enzyme, while certain substitutions at the P(1)'-P(4)' positions abolished the allosteric behavior, indicating that contact residues at these positions in the C1s enzyme must mediate the cooperativity. The studies reported here highlight the importance of prime subsites in C1s for interaction with its cognate substrates in the complement pathway and therefore yield greater understanding of the mechanism of interaction between this vital protease and its physiological substrates.  相似文献   

8.
The pseudorabies virus protein kinase prefers model substrates containing arginyl residues on the amino-terminal side of a target seryl or threonyl residue. We have defined this substrate specificity more precisely in experiments using a new series of synthetic model peptides. When the number of arginyl residues was varied from two to four in substrates of the type RnASVA it was found that peptides with four arginyl residues constituted the best substrates, although the most marked decrease in Km was seen on increasing the number of arginyl residues from two to three. The effect of varying the number of 'spacer' alanyl residues from zero to three was investigated in peptides of the type R4AmSVA, and the peptide with one alanyl residue was found to be the best substrate, making R4X the optimal amino-terminal environment for this enzyme. A similar substrate specificity was observed with the herpes simplex type 1 protein kinase. Protein kinase C was found to have a quite similar substrate preference to the viral enzyme as far as the number and position of the amino-terminal basic residues was concerned; but, unlike the viral protein kinase, it also requires carboxy-terminal basic residues in optimal peptide substrates, and can tolerate the substitution of lysyl for arginyl residues. The cyclic AMP-dependent protein kinase, like the viral enzyme, had favourable kinetic constants for this series of peptides, but differed from the latter in being able to catalyze the phosphorylation of the peptides with two to four arginyl residues with similar efficiency. Studies with the protein, clupeine Y1, as substrate indicated that the pseudorabies virus protein kinase can tolerate arginyl residues on the carboxyl-terminal side of its target residue when there are suitable amino-terminal arginyl determinants. In this respect the virus protein kinase resembled protein kinase C but differed from the cyclic AMP-dependent protein kinase which cannot tolerate such carboxyl-terminal basic residues. The relationship of substrate specificity with model peptides to the ability of the pseudorabies virus protein kinase to phosphorylate proteins in vitro and in vivo is discussed.  相似文献   

9.
We have developed a cellular automata model of an enzyme reaction with a substrate in water. The model produces Michaelis-Menten kinetics with good Lineweaver-Burk plots. The variation in affinity parameters predicts that, in general, hydrophobic substrates are more reactive with enzymes, this attribute being more important than the relationship between enzyme and substrate. The ease of generation and the illustrative value of the model lead us to believe that cellular automata models have a useful role in the study of dynamic phenomena such as enzyme kinetics.  相似文献   

10.
The previous notion that the amino acid side chain at position 104 of subtilisins is involved in the binding of the side chain at position P4 of the substrate has been investigated. The amino acid residue Val104 in subtilisin 309 has been replaced by Ala, Arg, Asp, Phe, Ser, Trp and Tyr by site-directed mutagenesis. It is shown that the P4 specificity of this enzyme is not determined solely by the amino acid residue occupying position 104, as the enzyme exhibits a marked preference for aromatic groups in P4, regardless of the nature of the position-104 residue. With hydrophilic amino acid residues at this position, no involvement is seen in binding of either hydrophobic or hydrophilic amino acid residues at position P4 of the substrates. The substrate with Asp in P4 is an exception, as the preference for this substrate is increased dramatically by introduction of an arginine residue at position 104 in the enzyme, presumably due to a substrate-induced conformational change. However, when position 104 is occupied by hydrophobic residues, it is highly involved in binding of hydrophobic amino acid residues, either by increasing the hydrophobicity of S4 or by determining the size of the pocket. The results suggest that the amino acid residue at position 104 is mobile such that it is positioned in the S4 binding site only when it can interact favourably with the substrate's side chain at position P4.  相似文献   

11.
The full-length and ectodomain forms of beta-site APP cleavage enzyme (BACE) have been cloned, expressed in Sf9 cells, and purified to homogeneity. This aspartic protease cleaves the amyloid precursor protein at the beta-secretase site, a critical step in the Alzheimer's disease pathogenesis. Comparison of BACE to other aspartic proteases such as cathepsin D and E, napsin A, pepsin, and renin revealed little similarity with respect to the substrate preference and inhibitor profile. On the other hand, these parameters are all very similar for the homologous enzyme BACE2. Based on a collection of decameric substrates, it was found that BACE has a loose substrate specificity and that the substrate recognition site in BACE extends over several amino acids. In common with the aspartic proteases mentioned above, BACE prefers a leucine residue at position P1. Unlike cathepsin D etc., BACE accepts polar or acidic residues at positions P2'0 and P1 but prefers bulky hydrophobic residues at position P3. BACE displays poor kinetic constants toward its known substrates (wild-type substrate, SEVKM/DAEFR, K(m) = 7 microm, K(cat) = 0.002 s(-1); Swedish mutant, SEVNL/DAEFR, K(m) = 9 microm, K(cat) = 0.02 s(-1)). A new substrate (VVEVDA/AVTP, K(m) = 1 microm, K(cat) = 0.004) was identified by serendipity.  相似文献   

12.
Asparaginase II from Escherichia coli is well-known member of the bacterial class II amidohydrolases. Enzymes of this family utilize a peculiar catalytic mechanism in which a pair of threonine residues play pivotal roles. Another common feature is a mobile surface loop that closes over the active site when the substrates is bound. We have studied the motion of the loop by stopped-flow experiments using the fluorescence of tryptophan residues as the spectroscopic probe. With wild-type enzyme the fluorescence of the only tryptophan, W66, was monitored. Here asparagine induced a rapid closure of the loop. The rate constants of the process (100-150 s(-1) at 4 degrees C) were considerably higher than those of the rate-limiting catalytic step. A more selective spectroscopic probe was generated by replacing W66 with tyrosine and Y25, a component of the loop, with tryptophan. In the resulting enzyme variant, k(cat) and the rate of loop movement were reduced by factors of 10(2) and >10(3), respectively, while substrate binding was unaffected. This indicates that the presence of tyrosine in position 25 is essential for both loop closure and catalysis. Numerical simulations of the observed transients are consistent with a model where loop closure is an absolute prerequisite for substrate turnover.  相似文献   

13.
The structural changes taking place in the enzyme thymidine phosphorylase (TPase, also known as PD-ECGF) that are required to achieve catalytic competence upon binding thymidine and phosphate have been simulated by means of targeted molecular dynamics (tMD). The hinge regions were characterized by structural homology comparisons with pyrimidine nucleoside phosphorylase, whose X-ray structure has been solved both in a closed and in an open form. The rearrangement of residues around the substrate that was observed during the tMD trajectory suggested that His-85 could be playing an important role in the catalytic mechanism. A quantum mechanical study of the reaction in the presence of the most relevant active site residues was then performed at the semiempirical level. The results revealed that His-85 could be involved in the protonation of the pyrimidine base at the O2 position to yield the enol tautomer of the base. To establish the role of this oxygen atom in the reaction, ground states, transition states, and final products were studied using higher level ab initio methods starting from both thymidine and 2-thiothymidine as alternative substrates. Comparison of both transition states showed that replacing the oxygen at position 2 of the pyrimidine base by sulfur should accelerate the reaction rate. Consistent with this result, 2-thiothymidine was shown to be a better substrate for TPase than the natural substrate, thymidine. For simulating the final step of the reaction, tMD simulations were used to study domain opening upon product formation considering both the enol and keto tautomers of thymine. Product release from the enzyme was easiest in the simulation that incorporated the keto tautomer of thymine, suggesting that the enol intermediate spontaneously tautomerizes back to the more energetically stable keto form. These results highlight a previously unreported role for His-85 in the catalytic mechanism of TPase and can have important implications for the design of novel TPase inhibitors.  相似文献   

14.
l-3-Hydroxyacyl-CoA dehydrogenase reversibly catalyzes the conversion of l-3-hydroxyacyl-CoA to 3-ketoacyl-CoA concomitant with the reduction of NAD(+) to NADH as part of the beta-oxidation spiral. In this report, crystal structures have been solved for the apoenzyme, binary complexes of the enzyme with reduced cofactor or 3-hydroxybutyryl-CoA substrate, and an abortive ternary complex of the enzyme with NAD(+) and acetoacetyl-CoA. The models illustrate positioning of cofactor and substrate within the active site of the enzyme. Comparison of these structures with the previous model of the enzyme-NAD(+) complex reveals that although significant shifting of the NAD(+)-binding domain relative to the C-terminal domain occurs in the ternary and substrate-bound complexes, there are few differences between the apoenzyme and cofactor-bound complexes. Analysis of these models clarifies the role of key amino acids implicated in catalysis and highlights additional critical residues. Furthermore, a novel charge transfer complex has been identified in the course of abortive ternary complex formation, and its characterization provides additional insight into aspects of the catalytic mechanism of l-3-hydroxyacyl-CoA dehydrogenase.  相似文献   

15.
The formation of the reactive enzyme–substrate complex of formate dehydrogenase has been investigated by molecular dynamics techniques accounting for different conformational states of the enzyme. Simulations revealed that the transport of substrate to the active site through the substrate channel proceeds in the open conformation of enzyme due to the crucial role of the Arg284 residue acting as a vehicle. However, formate binding in the active site of the open conformation leads to the formation of a nonproductive enzyme–substrate complex. The productive Michaelis complex is formed only in the closed enzyme conformation after the substrate and coenzyme have bound, when required rigidity of the binding site and reactive formate orientation due to interactions with Arg284, Asn146, Ile122, and His332 residues is attained. Then, the high occupancy (up to 75%) of the reactive substrate–coenzyme conformation is reached, which was demonstrated by hybrid quantum mechanics/molecular mechanics simulations using various semiempirical Hamiltonians.  相似文献   

16.
Fang M  Macova A  Hanson KL  Kos J  Palmer DR 《Biochemistry》2011,50(40):8712-8721
MenD catalyzes the thiamin diphosphate-dependent decarboxylative carboligation of α-ketoglutarate and isochorismate. The enzyme is essential for menaquinone biosynthesis in many bacteria and has been proposed to be an antibiotic target. The kinetic mechanism of this enzyme has not previously been demonstrated because of the limitations of the UV-based kinetic assay. We have reported the synthesis of an isochorismate analogue that acts as a substrate for MenD. The apparent weaker binding of this analogue is advantageous in that it allows accurate kinetic experiments at substrate concentrations near K(m). Using this substrate in concert with the dead-end inhibitor methyl succinylphosphonate, an analogue of α-ketoglutarate, we show that MenD follows a ping-pong kinetic mechanism. Using both the natural and synthetic substrates, we have measured the effects of 12 mutations of residues at the active site. The results give experimental support to previous models and hypotheses and allow observations unavailable using only the natural substrate.  相似文献   

17.
Chondroitinase B from Flavobacterium heparinum is the only known lyase that cleaves the glycosaminoglycan, dermatan sulfate (DS), as its sole substrate. A recent co-crystal structure of chondroitinase B with a disaccharide product of DS depolymerization has provided some insight into the location of the active site and suggested potential roles of some active site residues in substrate binding and catalysis. However, this co-crystal structure was not representative of the actual enzyme-substrate complex, because the disaccharide product did not have the right length or the chemical structure of the minimal substrate (tetrasaccharide) involved in catalysis. Therefore, only a limited picture of the functional role of active site residues in DS depolymerization was presented in previous structural studies. In this study, by docking a DS tetrasaccharide into the proposed active site of the enzyme, we have identified novel roles of specific active site amino acids in the catalytic function of chondroitinase B. Our conformational analysis also revealed a unique, symmetrical arrangement of active site amino acids that may impinge on the catalytic mechanism of action of chondroitinase B. The catalytic residues Lys-250, Arg-271, His-272, and Glu-333 along with the substrate binding residues Arg-363 and Arg-364 were mutated using site-directed mutagenesis, and the kinetics and product profile of each mutant were compared with recombinant chondroitinase B. Mutating Lys-250 to alanine resulted in inactivation of the enzyme, potentially attributable to the role of the residue in stabilizing the carbanion intermediate formed during enzymatic catalysis. The His-272 and Glu-333 mutants showed diminished enzymatic activity that could be indicative of a possible role for one or both residues in the abstraction of the C-5 proton from the galactosamine. In addition, the Arg-364 mutant had an altered product profile after exhaustive digestion of DS, suggesting a role for this residue in defining the substrate specificity of chondroitinase B.  相似文献   

18.
Hao Hu  Haiyan Liu  Yunyu Shi 《Proteins》1997,27(4):545-555
Different pathways of the metal-induced isomerization of D-xylose to D-xylulose are investigated and compared in detail using energy minimization and molecular dynamics simulation. Two theoretical models are constructed for the reaction: in vacuum and in the enzyme D-xylose isomerase. The vacuum model is constructed based on the X-ray structure of the active site of D-xylose isomerase. It contains the atoms directly involved in the reaction and is studied using a semi-empirical molecular orbital method (PM3). The model in the enzyme includes the effects of the enzyme environment on the reaction using a combined quantum mechanical and molecular mechanical potential. For both models, the structures of the reactants, products, and intermediate complexes along the isomerization pathway are optimized. The effects of the position of the “catalytic Mg2+ ion” on the energies of the reactions are studied. The results indicate: 1) in vacuum, the isomerization reaction is favored when the catalytic metal cation is at site A, which is remote from the substrate; 2) in the enzyme, the catalytic metal cation, starting from site A, moves and stays at site B, which is close to the substrate; analysis of the charge redistribution of the active site during the catalytic process shows that the metal ion acts as a Lewis acid to polarize the substrate and catalyze the hydride shift; these results are consistent with previous experimental observations; and 3) Lys183 plays an important role in the isomerization reaction. The ϵ-NH3+ group of its side chain can provide a proton to the carboxide ion of the substrate to form a hydroxyl group after the hydride shift step. This role of Lys183 has not been suggested before. Based on our calculations, we believe that this is a reasonable mechanism and consistent with site-directed mutation experiments. © 1997 Wiley-Liss Inc.  相似文献   

19.
An extensive screening among microorganisms for the presence of post-proline-specific endopeptidase activity was performed. This activity was found among ordinary bacteria from soil samples but not among fungi and actinomycetes. This result is in contrast to the previous notion that this activity is confined to the genus Flavobacterium. A proline endopeptidase was isolated from a Xanthomonas sp. and characterized with respect to physicochemical and enzymatic properties. The enzyme is composed of a single peptide chain with a molecular weight of 75,000. The isoelectric point is 6.2. It is inhibited by diisopropylfluorophosphate and may therefore be classified as a serine endopeptidase. The activity profile is bell shaped with an optimum at pH 7.5. By using synthetic peptide substrates and intramolecular fluorescence quenching it was possible to study the influence of substrate structure on the rate of hydrolysis. The enzyme specifically hydrolyzed Pro-X peptide bonds. With Glu at position X, low rates of hydrolysis were observed; otherwise the enzyme exhibited little preference for particular amino acid residues at position X. A similar substrate preference was observed with respect to the amino acid residue preceding the prolyl residue in the substrate. The enzyme required a minimum of two amino acid residues toward the N terminus from the scissile bond, but further elongation of the peptide chain by up to six amino acid residues caused only a threefold increase in the rate of hydrolysis. Attempts to cleave at the prolyl residues in oxidized RNase failed, indicating that the enzyme does not hydrolyze long peptides, a peculiar property it shares with other proline-specific endopeptidases.  相似文献   

20.
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