首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Studies were made on the ultraviolet difference-spectra of glucoamylase from Rhizopus niveus [EC 3.2.1.3] specifically produced by the substrate maltose and the inhibitors, glucose, glucono-1: 5-lactone (gluconolactone), methyl beta-D-glucoside, cellubiose, and cyclohexa-, and cyclohepta-amyloses. Of these, maltose and gluconolactone produced characteristic difference spectra with a trough near 300 nm. Based on studies with a model compound for a tryptophan residue, Ac-Trp, this trough was attributed to the effect of a negative charge upon the tryptophan residue. From the concentration dependency of the difference spectra, the dissociation constants of the complexes between the enzyme and maltose, glucose, and gluconolactone were evaluated to be 1.2 mM, 51 mM, and 1.5 mM, respectively. These values are in good agreement with the values of Km or K1 obtained from the steady-state kinetics. The difference-spectrophotometric data suggested that referring to the values of subsite affinities of glucoamylase, maltose, and gluconolactone occupy mainly Subsite 1, where the non-reducing-end glucose residue of a substrate is bound in a productive form and that a tryptophan residue with shows a trough near 300 nm in difference spectra is located in this subsite.  相似文献   

2.
Inhibitory activities of 1-deoxynojirimycin and gluconolactone on Aspergillus niger glucoamylase were studied in relation to the subsite structure of the enzyme. Although both of these inhibitors are considered to bind at subsite 1 of the enzyme active site, 1-deoxynojirimycin showed competitive type inhibition but gluconolactone was a mixed type (or noncompetitive type) inhibitor for the hydrolysis of p-nitrophenyl alpha-D-glucoside. The former type of inhibition suggested that the main binding mode of the substrate was productive, but the latter, nonproductive. A possible way of explaining these apparent inconsistent results is to assume that the main binding mode of the substrate is productive and gluconolactone forms a nonproductive ternary complex with the enzyme and the substrate.  相似文献   

3.
Trp120 of Aspergillus awamori glucoamylase has previously been shown by chemical modification to be essential for activity and tentatively to be located near subsite 4 of the active site. To further test its role, restriction sites were inserted in the cloned A.awamori gene around the Trp120 coding region, and cassette mutagenesis was used to replace it with His, Leu, Phe and Tyr. All four mutants displayed 2% or less of the maximal activity (kcat) of wild-type glucoamylase towards maltose and maltoheptaose. Michaelis constants (KM) of mutants decreased 2- to 3-fold for maltose and were essentially unchanged for maltoheptaose compared with the wild type, except for a greater than 3-fold decrease for maltoheptaose with the Trp120----Tyr mutant. This mutant also bound isomaltose more strongly and had more selectivity for its hydrolysis than wild-type glucoamylase. A subsite map generated from malto-oligosaccharide substrates having 2-7 D-glucosyl residues indicated that subsites 1 and 2 had greater affinity for D-glucosyl residues in the Trp120----Tyr mutant than in wild-type glucoamylase. These results suggest that Trp120 from a distant subsite is crucial for the stabilization of the transition-state complex in subsites 1 and 2.  相似文献   

4.
Subsite mapping is a crucial procedure in the characterization of α-amylases (EC 3.2.1.1), which are extensively used in starch-based industries and in diagnosis of pancreatic and salivary glands disorders. A computer-aided method has been developed for subsite mapping of α-amylases, which substitutes the difficult, expensive, and time-consuming experimental determination of action patterns to crystal structures based energy calculations. Interaction energies between enzymes and carbohydrate substrates were calculated after short energy minimization by a molecular mechanics program. A training set of wild type and mutant amylases with known experimental action patterns of 13 enzymes of wide range of origin was used to set up the procedure. Calculations for training set resulted in good correlation in case of subsite binding energies (r2 = 0.827–0.929) and bond cleavage frequencies (r2 = 0.727–0.835). A set of eight novel barley amylase 1 mutants was used to test our model. Subsite binding energies were predicted with r2 = 0.502 correlation coefficient, while bond cleavage frequency prediction resulted in r2 = 0.538. Our computer-aided procedure may supplement the experimental subsite mapping methods to predict and understand characteristic features of α-amylases.  相似文献   

5.
The subsite structure of Thermoactinomyces vulgaris α-amylase was estimated from its action mode and rate parameters of hydrolysis on maltooligosaccharides. These results led to the conclusion that this α-amylase has six subsites with the catalytic site located between the third and fourth subsites from the non-reducing end side. Subsite affinities were calculated to be 0.38, 5.46, 2.72 and 0.23 kcal/mol for subsites 1, 2, 5 and 6, respectively, and the sum of the affinities of subsite 3 and 4 to be ?3.41 kcal/mol. The unique action mode of this α-amylase on various substrates was interpreted in terms of the subsite structure.  相似文献   

6.
The presteady-state and steady-state kinetics of the binding and hydrolysis of substrates, maltose and isomaltose, and the transition-state analogue, gluconolactone, by glucoamylase from Aspergillus niger were investigated using initial-rate, stopped-flow and steady-state methods. The change in the intrinsic fluorescence of the enzyme was monitored. Distinct mechanistic differences were observed in the interaction of the enzyme with maltose compared to isomaltose. Hydrolysis of maltose requires a three-step mechanism, whereas that of isomaltose involves at least one additional step. The rates of an observed conformational change, which is the second discernible step of the reactions, clearly show a tighter binding of maltose compared to isomaltose, probably because the reverse rate constants differ. Compared to the non-enzymic hydrolysis the transition-state stabilization energy of glucoamylase is approximately -66 kJ/mol with maltose and only -14 kJ/mol with isomaltose. Kinetic analysis of the binding of the inhibitor, gluconolactone, implies that independent interactions of two molecules occur. One of these, apparently, is a simple, fast association reaction in which gluconolactone is weakly bound. The other resembles binding of maltose, involving a fast association followed by a conformational change. Based on the results obtained, we propose new reaction mechanisms for Aspergillus glucoamylase.  相似文献   

7.
Hydrolysis of starch or oligosaccharides by mammalian amylases, in general, results in maltose as the leaving group. The active site of these amylases harbors three aromatic residues Trp59, Tyr62, and Tyr151, which provide stacking interactions to the bound glucose moieties. We hypothesized that Tyr151, located at the S2' subsite, may influence the size of the leaving group. Therefore, using a baculovirus expression system, we generated a mutant Y151M in which the tyrosine at position 151 of human salivary amylase is replaced by a methionine. The specific activity, K(m), rate of hydrolysis, and the product distribution for Y151M were distinctly different from those of the wild-type enzyme using starch and oligosaccharides as substrates. The mutant enzyme Y151M consistently produced glucose as the minimal leaving group and exhibited a twofold increase in K(m). These results suggest that the stacking interaction at subsite S2' in the wild type plays a role in hydrolysis.  相似文献   

8.
To elucidate how temperature effects subsite mapping of a thermostable alpha-amylase from Bacillus licheniformis (BLA), a comparative study was performed by using 2-chloro-4-nitrophenyl (CNP) beta-maltooligosides with degree of polymerisation (DP) 4-10 as model substrates. Action patterns, cleavage frequencies and subsite binding energies were determined at 50 degrees C, 80 degrees C and 100 degrees C. Subsite map at 80 degrees C indicates more favourable bindings compared to the hydrolysis at 50 degrees C. Hydrolysis at 100 degrees C resulted in a clear shift in the product pattern and suggests significant differences in the active site architecture. Two preferred cleavage modes were seen for all substrates in which subsite (+2) and (+3) were dominant, but CNP-G1 was never formed. In the preferred binding mode of shorter oligomers, CNP-G2 serves as the leaving group (79%, 50%, 59% and 62% from CNP-G4, CNP-G5, CNP-G6 and CNP-G7, respectively), while CNP-G3 is the dominant hydrolysis product from CNP-G8, CNP-G9, and CNP-Gl0 (62%, 68% and 64%, respectively). The high binding energy value (-17.5 kJ/mol) found at subsite (+2) is consistent with the significant formation of CNP-G2. Subsite mapping at 80 degrees C and 100 degrees C confirms that there are no further binding sites despite the presence of longer products.  相似文献   

9.
MaltodExtrin (high-d.p. malto-oligosaccharides) was found to produce a trough at 303 nm in the difference spectrum of glucoamylase (E.C. 3.2.1.3) from Rhizopus niveus upon binding with the enzyme; this trough disappears upon hydrolysis. The trough, which was ascribed to a change, in the electrostatic environment of a tryptophan residue at the terminal subsite of the enzyme, was found closely related to the formation of the enzyme-substrate complex. The kinetics of binding of maltodextrin and maltotriose to the enzyme were studied at pH 4.5. and 5°, by monitoring the trough by the stopped-flow method. The result was consistent with a two-step mechanism, in which a fast, bimolecular association is followed by a slower, uni-molecular isomerization-process. The latter process involves an environmental change of the tryptophan residue, and is considered to be closely connected to the formation of the productive complex essential for the catalysis.  相似文献   

10.
Hrmova M  Fincher GB 《Carbohydrate research》2007,342(12-13):1613-1623
Higher plant, family GH3 beta-D-glucan glucohydrolases exhibit exo-hydrolytic and retaining (e-->e) mechanisms of action and catalyze the removal of single glucosyl residues from the non-reducing termini of beta-D-linked glucosidic substrates, with retention of anomeric configuration. The broad specificity beta-D-glucan glucohydrolases are likely to play roles in cell wall re-modelling, turn-over of cell wall components and possibly in plant defence reactions against pathogens. Crystal structures of the barley beta-D-glucan glucohydrolase, obtained from both native enzyme and from the enzyme in complex with a substrate analogues and mechanism-based inhibitors, have enabled the basis of substrate specificity, the mechanism of catalysis, and the role of domain movements during the catalytic cycle to be defined in precise molecular terms. The active site of the enzyme forms a shallow 'pocket' that is located at the interface of two domains of the enzyme and accommodates two glucosyl residues. The propensity of the enzyme to hydrolyze a broad range of substrates with (1-->2)-, (1-->3)-, (1-->4)- and (1-->6)-beta-D-glucosidic linkages is explained from crystal structures of the enzyme in complex with non-hydrolysable S-glycoside substrate analogues, and from molecular modelling. During binding of gluco-oligosaccharides, the glucosyl residue at subsite -1 is locked in a highly constrained position, but the glucosyl residue at the +1 subsite is free to adjust its position between two tryptophan residues positioned at the entry of the active site pocket. The flexibility at subsite +1 and the projection of the remainder of the substrate away from the pocket provide a structural rationale for the capacity of the enzyme to accommodate and hydrolyze glucosides with different linkage positions and hence different overall conformations. While mechanism-based inhibitors with micromolar Ki constants bind in the active site of the enzyme and form esters with the catalytic nucleophile, transition-state mimics bind with their 'glucose' moieties distorted into the 4E conformation, which is critical for the nanomolar binding of these inhibitors to the enzyme. The glucose product of the reaction, which is released from the non-reducing termini of substrates, remains bound to the beta-D-glucan glucohydrolase in the -1 subsite of the active site, until a new substrate molecule approaches the enzyme. If dissociation of the glucose from the enzyme active site could be synchronized throughout the crystal, time-resolved Laue X-ray crystallography could be used to follow the conformational changes that occur as the glucose product diffuses away and the incoming substrate is bound by the enzyme.  相似文献   

11.
The binding of glucono-1,5-lactone (gluconolactone) with almond beta-glucosidase was studied at pH 5.0 and 25 degrees C, in the absence and presence of glucose, by monitoring the enzyme fluorescence as a probe. From the results of fluorometric titration, the dissociation constant Kd and the maximum fluorescence intensity increase (percent) of the enzyme-gluconolactone complex relative to the enzyme alone, delta Fmax, were determined to be 12.7 microM and 14.7%, respectively. From the study of the temperature dependence of Kd, delta G degrees, delta H degrees and delta S degrees for the binding were evaluated to be -6.7 kcal mol-1, -3.5 kcal mol-1, and 10.8 e.u. (cal mol-1 deg-1), respectively, at 25 degrees C. The analysis of the fluorometric titration data in the presence of glucose revealed that these ligands bind competitively to the enzyme, probably at the same site. The results of a stopped-flow kinetic study are consistent with the following two-step mechanism: (formula; see text) which indicates that gluconolactone (L) and the enzyme (E) transiently form a loosely bound complex, ELtr (k-1/k+1 = 4.5 mM), in the first rapid bimolecular association step, and ELtr is converted into a more tightly bound complex EL (k+2 = 94 s-1, k-2 = 0.36 s-1) in the subsequent slow unimolecular process. The fluorescence intensity increase occurs solely in the latter step.  相似文献   

12.
The structure of the complex between a catalytically compromised family 10 xylanase and a xylopentaose substrate has been determined by X-ray crystallography and refined to 3.2 A resolution. The substrate binds at the C-terminal end of the eightfold betaalpha-barrel of Pseudomonas fluorescens subsp. cellulosa xylanase A and occupies substrate binding subsites -1 to +4. Crystal contacts are shown to prevent the expected mode of binding from subsite -2 to +3, because of steric hindrance to subsite -2. The loss of accessible surface at individual subsites on binding of xylopentaose parallels well previously reported experimental measurements of individual subsites binding energies, decreasing going from subsite +2 to +4. Nine conserved residues contribute to subsite -1, including three tryptophan residues forming an aromatic cage around the xylosyl residue at this subsite. One of these, Trp 313, is the single residue contributing most lost accessible surface to subsite -1, and goes from a highly mobile to a well-defined conformation on binding of the substrate. A comparison of xylanase A with C. fimi CEX around the +1 subsite suggests that a flatter and less polar surface is responsible for the better catalytic properties of CEX on aryl substrates. The view of catalysis that emerges from combining this with previously published work is the following: (1) xylan is recognized and bound by the xylanase as a left-handed threefold helix; (2) the xylosyl residue at subsite -1 is distorted and pulled down toward the catalytic residues, and the glycosidic bond is strained and broken to form the enzyme-substrate covalent intermediate; (3) the intermediate is attacked by an activated water molecule, following the classic retaining glycosyl hydrolase mechanism.  相似文献   

13.
Subsite affinity maps of long substrate binding clefts in barley alpha-amylases, obtained using a series of maltooligosaccharides of degree of polymerization of 3-12, revealed unfavorable binding energies at the internal subsites -3 and -5 and at subsites -8 and +3/+4 defining these subsites as binding barriers. Barley alpha-amylase 1 mutants Y105A and T212Y at subsite -6 and +4 resulted in release or anchoring of bound substrate, thus modifying the affinities of other high-affinity subsites (-2 and +2) and barriers. The double mutant Y105A-T212Y displayed a hybrid subsite affinity profile, converting barriers to binding areas. These findings highlight the dynamic binding energy distribution and the versatility of long maltooligosaccharide derivatives in mapping extended binding clefts in alpha-amylases.  相似文献   

14.
The kinetics of glucose-fructose oxidoreductase from Zymomonas mobilis   总被引:5,自引:0,他引:5  
Glucose-fructose oxidoreductase operates by a classic ping-pong mechanism with a single site for all substrates: glucose, fructose, gluconolactone and sorbitol. The Km values for these substrates were determined. The values of kcat are 200 s-1 and 0.8 s-1 for the forward and reverse directions respectively. The overall catalytic process consists of two half-reactions with alternate reduction of NADP+ and oxidation of NADPH tightly bound to the enzyme. Reduction of enzyme-NADP+ by glucose and oxidation of enzyme-NADPH by gluconolactone involve single first-order processes. The values of the rate constants at saturating substrate are 2100 s-1 and 8 s-1 respectively; deuterium isotope effects indicate that these are for the hydrogen transfer step. Oxidation of enzyme-NADPH by fructose is first order with a limiting rate constant of at least 430 s-1. The reaction of enzyme-NADP+ with sorbitol is biphasic, with rate constants for both phases less than 1 s-1. This behaviour is explained by a mechanism in which the slow cyclisation of the acyclic form of fructose follows its dissociation from the enzyme. The rate-determining steps for the overall reaction are probably dissociation of gluconolactone in the forward direction and hydrogen transfer from sorbitol to enzyme-bound NADP+ in the reverse direction.  相似文献   

15.
A metal chelating ligand is bonded to alpha-, beta-, and gamma-cyclodextrin by the reaction of diethylenetraminepentaacetic dianhydride with the corresponding 6-mono- and 2-mono(amine)cyclodextrin. Adding Dy(III) to the cyclodextrin derivatives causes shifts in the (1)H-NMR spectra of substrates such as propranolol, tryptophan, aspartame, carbinoxamine, pheniramine, doxylamine, and 1-anilino-8-naphthalenesulfonate. The Dy(III)-induced shifts enhance the enantiomeric resolution in the NMR spectra of several substrates. Enhancements in enantiomeric resolution using cyclodextrin derivatives with the amine tether are compared to previously described compounds in which the chelating ligand is attached through an ethylenediamine tether. In general, the Dy(III) complex of the 6-beta-derivative with the amine tether is a more effective chiral resolving agent than the complex with the ethylenediamine tether. The opposite trend is observed with the 2-beta-derivatives. The presence of the chelating ligand in the 2-beta-derivative hinders certain substrates from entering the cavity. For cationic substrates, evidence suggests that a cooperative association involving inclusion in the cavity and association with the Dy(III) unit occurs. Enhancements in enantiomeric resolution in the spectrum of tryptophan are greater for the secondary alpha- and gamma-derivatives than the beta-derivative.  相似文献   

16.
Kinetics of human polymerase beta binding to gapped DNA substrates having single stranded (ss) DNA gaps with five or two nucleotide residues in the ssDNA gap has been examined, using the fluorescence stopped-flow technique. The mechanism of the recognition does not depend on the length of the ssDNA gap. Formation of the enzyme complex with both DNA substrates occurs by a minimum three-step reaction, with the bimolecular step followed by two isomerization steps. The results indicate that the polymerase initiates the association with gapped DNA substrates through the DNA-binding subsite located on the 8-kDa domain of the enzyme. This first association step is independent of the length of the ssDNA gap and is characterized by similar rate constants for both examined DNA substrates. The subsequent, first-order transition occurs at the rate of approximately 600-1200 s(-1). This is the major docking step accompanied by favorable free energy changes in which the 31-kDa domain engages in interactions with the DNA. The 5'-terminal PO(4)(-) group downstream from the primer is not a specific recognition element of the gap. However, the phosphate group affects the enzyme orientation in the complex with the DNA, particularly, for the substrate with a longer gap.  相似文献   

17.
Asp176, Glu179 and Glu180 of Aspergillus awamori glucoamylase appeared by differential labeling to be in the active site. To test their functions, they were replaced by mutagenesis with Asn, Gln and Gln respectively, and kinetic parameters and pH dependencies of all enzyme forms were determined. Glu179----Gln glucoamylase was not active on maltose or isomaltose, while the kcat for maltoheptaose hydrolysis decreased almost 2000-fold and the KM was essentially unchanged from wild-type glucoamylase. The The Glu180----Gln mutation drastically increased the KM and moderately decreased the kcat with maltose and maltoheptaose, but affected isomaltose hydrolysis less. Difference in substrate activation energies between Glu180----Gln and wild-type glucoamylases indicate that Glu180 binds D-glucosyl residues in subsite 2. The Asp176----Asn substitution gave moderate increases and decreases in KM and kcat respectively, and therefore similar increases in activation energies for the three substrates. This and the differences in subsite binding energies between Asp176----Asn and wild-type glucoamylases suggest that Asp176 is near subsite 1, where it stabilizes the transition state and interacts with Trp120 at subsite 4. Glu179 and Asp176 are thus proposed as the general catalytic acid and base of pKa 5.9 and 2.7 respectively. The charged Glu180 contributes to the high pKa value of Glu179.  相似文献   

18.
Family 3 beta-D-glucan glucohydrolases are distributed widely in higher plants. The enzymes catalyze the hydrolytic removal of beta-D-glucosyl residues from nonreducing termini of a range of beta-D-glucans and beta-D-oligoglucosides. Their broad specificity can be explained by x-ray crystallographic data obtained from a barley beta-D-glucan glucohydrolase in complex with nonhydrolyzable S-glycoside substrate analogs and by molecular modeling of enzyme/substrate complexes. The glucosyl residue that occupies binding subsite -1 is locked tightly into a fixed position through extensive hydrogen bonding with six amino acid residues near the bottom of an active site pocket. In contrast, the glucosyl residue at subsite +1 is located between two Trp residues at the entrance of the pocket, where it is constrained less tightly. The relative flexibility of binding at subsite +1, coupled with the projection of the remainder of bound substrate away from the enzyme's surface, means that the overall active site can accommodate a range of substrates with variable spatial dispositions of adjacent beta-D-glucosyl residues. The broad specificity for glycosidic linkage type enables the enzyme to perform diverse functions during plant development.  相似文献   

19.
Binding onto cellobiohydrolase II from Trichoderma reesei of glucose, cellobiose, cellotriose, derivatized and analogous compounds, is monitored by protein-difference-absorption spectroscopy and by titration of ligand fluorescence, either at equilibrium or by the stopped-flow technique. The data complete earlier results [van Tilbeurgh, H., Pettersson, L. G., Bhikhabhai, R., De Boeck, H. and Claeyssens, M. (1985) Eur. J. Biochem. 148, 329-334] indicating an extended active center, with putative subsites ABCD. Subsite A specifically complexes with beta-D-glucosides and D-glucose; at 25 degrees C the latter influences the concomitant binding of other ligands at neighbouring sites. For several ligands this cooperative effect for binding (at 0.33 M glucose and temperature range 4-37 degrees C) was characterized by a substantial increase of the enthalpic term (delta delta H = -35 kJ mol-1). Glucose (0.33 M) decreases the association and dissociation rate parameters of 4-methylumbelliferyl beta-D-cellobioside by one order of magnitude: k+ = (3.6 +/- 0.5) x 10(-5) M-1 s-1 versus (5.1 +/- 0.1) x 10(-6) M-1 s-1 (in the absence of glucose) and k- = (1.3 +/- 0.1) s-1 versus (14.0 +/- 0.3) s-1. As deduced from substrate-specificity studies and inhibition experiments, subsite B interacts with terminal non-reducing glucopyranosyl residues of oligomeric ligands and substrates, whereas catalytic (hydrolytic) cleavage occurs between C and D. Association constants 10-100 times higher than those for cellobiose or its glycosides were observed for D-glucopyranosyl-(1----4)-beta-D-xylopyranose and cellobionolactone derivatives, suggesting 'transition-state'-type binding for these ligands at subsite C. Although subsite D can accomodate a bulky chromophoric group (MeUmb) its preference for a glucosyl residue is reflected in the lower binding enthalpy of cellotriose (-34 kJ mol-1) as compared to cellobiose (-28.3 kJ mol-1) and MeUmb(Glc)2 (-11.6 kJ mol-1). This model indicates that oligomeric ligands (substrates) interact through cooperativity of their subunits at the extended binding site of cellobiohydrolase II.  相似文献   

20.
The radical anions (SCN)2.- and Br2.- produced during a pulse radiolysis of the respective potassium salts have been used to study the tryptophan residues of the glucoenzyme, glucoamylase I (EC 3.2.1.3.). At neutral pH, Br2.- reacted with the tryptophan residues of glucoamylase I as expected from previous studies of proteins and free amino acids. However, (SCN)2.- at neutral and high pH was surprisingly unreactive towards the native enzyme. Reaction did occur, however, between (SCN)2.- and glucoamylase from which one-third of the covalently bound carbohydrate had been removed, producing a tryptophyl radical. Reaction also occured between (SCN)2.- and glucoamylase I inactivated by treatment with sodium dodecyl sulphate, but the tryptophan residues were not involved. It is concluded from the results that two 'types' of tryptophan residues are found in glucoamylase I; both are attacked by Br2.- but only one type is attacked by (SCN)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