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 78 毫秒
1.
Kanai R  Haga K  Akiba T  Yamane K  Harata K 《Biochemistry》2004,43(44):14047-14056
Maltohexaose-producing amylase, called G6-amylase (EC 3.2.1.98), from alkalophilic Bacillus sp.707 predominantly produces maltohexaose (G6) from starch and related alpha-1,4-glucans. To elucidate the reaction mechanism of G6-amylase, the enzyme activities were evaluated and crystal structures were determined for the native enzyme and its complex with pseudo-maltononaose at 2.1 and 1.9 A resolutions, respectively. The optimal condition for starch-degrading reaction activity was found at 45 degrees C and pH 8.8, and the enzyme produced G6 in a yield of more than 30% of the total products from short-chain amylose (DP = 17). The crystal structures revealed that Asp236 is a nucleophilic catalyst and Glu266 is a proton donor/acceptor. Pseudo-maltononaose occupies subsites -6 to +3 and induces the conformational change of Glu266 and Asp333 to form a salt linkage with the N-glycosidic amino group and a hydrogen bond with secondary hydroxyl groups of the cyclitol residue bound to subsite -1, respectively. The indole moiety of Trp140 is stacked on the cyclitol and 4-amino-6-deoxyglucose residues located at subsites -6 and -5 within a 4 A distance. Such a face-to-face short contact may regulate the disposition of the glucosyl residue at subsite -6 and would govern the product specificity for G6 production.  相似文献   

2.
The role in activity of outer regions in the substrate binding cleft in alpha-amylases is illustrated by mutational analysis of Tyr(105) and Thr(212) localized at subsites -6 and +4 (substrate cleavage occurs between subsites -1 and +1) in barley alpha-amylase 1 (AMY1). Tyr(105) is conserved in plant alpha-amylases whereas Thr(212) varies in these and related enzymes. Compared with wild-type AMY1, the subsite -6 mutant Y105A has 140, 15, and <1% activity (k(cat)/K(m)) on starch, amylose DP17, and 2-chloro-4-nitrophenyl beta-d-maltoheptaoside, whereas T212Y at subsite +4 has 32, 370, and 90% activity, respectively. Thus engineering of aromatic stacking interactions at the ends of the 10-subsite long binding cleft affects activity very differently, dependent on the substrate. Y105A dominates in dual subsite -6/+4 [Y105A/T212(Y/W)]AMY1 mutants having almost retained and low activity on starch and oligosaccharides, respectively. Bond cleavage analysis of oligosaccharide degradation by wild-type and mutant AMY1 supports that Tyr(105) is critical for binding at subsite -6. Substrate binding is improved by T212(Y/W) introduced at subsite +4 and the [Y105A/T212(Y/W)]AMY1 double mutants synergistically enhanced productive binding of the substrate aglycone. The enzymatic properties of the series of AMY1 mutants suggest that longer substrates adopt several binding modes. This is in excellent agreement with computed distinct multiple docking solutions observed for maltododecaose at outer binding areas of AMY1 beyond subsites -3 and +3.  相似文献   

3.
Barley alpha-amylase 1 (AMY1) hydrolyzed amylose with a degree of multiple attack (DMA) of 1.9; that is, on average, 2.9 glycoside bonds are cleaved per productive enzyme-substrate encounter. Six AMY1 mutants, spanning the substrate binding cleft from subsites -6 to +4, and a fusion protein, AMY1-SBD, of AMY1 and the starch binding domain (SBD) of Aspergillus niger glucoamylase were also analyzed. DMA of the subsite -6 mutant Y105A and AMY1-SBD increased to 3.3 and 3.0, respectively. M53E, M298S, and T212W at subsites -2, +1/+2, and +4, respectively, and the double mutant Y105A/T212W had decreased DMA of 1.0-1.4. C95A (subsite -5) had a DMA similar to that of wild type. Maltoheptaose (G7) was always the major initial oligosaccharide product. Wild-type and the subsite mutants released G6 at 27-40%, G8 at 60-70%, G9 at 39-48%, and G10 at 33-44% of the G7 rate, whereas AMY1-SBD more efficiently produced G8, G9, and G10 at rates similar to, 66%, and 60% of G7, respectively. In contrast, the shorter products appeared with large individual differences: G1, 0-15%; G2, 8-43%; G3, 0-22%; and G4, 0-11% of the G7 rate. G5 was always a minor product. Multiple attack thus involves both longer translocation of substrate in the binding cleft upon the initial cleavage to produce G6-G10, essentially independent of subsite mutations, and short-distance moves resulting in individually very different rates of release of G1-G4. Accordingly, the degree of multiple attack as well as the profile of products can be manipulated by structural changes in the active site or by introduction of extra substrate binding sites.  相似文献   

4.
Rockey WM  Laederach A  Reilly PJ 《Proteins》2000,40(2):299-309
The Lamarckian genetic algorithm of AutoDock 3.0 was used to dock alpha-maltotriose, methyl alpha-panoside, methyl alpha-isopanoside, methyl alpha-isomaltotrioside, methyl alpha-(6(1)-alpha-glucopyranosyl)-maltoside, and alpha-maltopentaose into the closed and, except for alpha-maltopentaose, into the open conformation of the soybean beta-amylase active site. In the closed conformation, the hinged flap at the mouth of the active site closes over the substrate. The nonreducing end of alpha-maltotriose docks preferentially to subsites -2 or +1, the latter yielding nonproductive binding. Some ligands dock into less optimal conformations with the nonreducing end at subsite -1. The reducing-end glucosyl residue of nonproductively-bound alpha-maltotriose is close to residue Gln194, which likely contributes to binding to subsite +3. In the open conformation, the substrate hydrogen-bonds with several residues of the open flap. When the flap closes, the substrate productively docks if the nonreducing end is near subsites -2 or -1. Trisaccharides with alpha-(1-->6) bonds do not successfully dock except for methyl alpha-isopanoside, whose first and second glucosyl rings dock exceptionally well into subsites -2 and -1. The alpha-(1-->6) bond between the second and third glucosyl units causes the latter to be improperly positioned into subsite +1; the fact that isopanose is not a substrate of beta-amylase indicates that binding to this subsite is critical for hydrolysis.  相似文献   

5.
The active centre of porcine pancreatic alpha-amylase contains five subsites. Their occupancy has been studied using as a substrate maltooligosaccharide of various chain lengths (maltose up to maltoheptaose), some of their p- and o-nitrophenylated derivatives, and 412-residue amylose. Quantitative analysis of the digestion products allowed the determination of the subsite occupancy for the various productive complexes, the bond cleavage frequency and respective kcati (where i is the binding mode). The catalytic efficiency (kcat/Km) increases with chain length from maltose (2 M-1 X S-1) up to amylose (1.06 X 10(7) M-1 X S-1). The kinetic parameters of p-nitrophenylmaltoside hydrolysis are quite close to those of maltose, and the ortho compound behaves as maltotriose. Determination of binding energy of glucose residue at the various subsites calculated according to the method of Hiromi et al. (Hiromi, K., Nitta, Y., Numata, C. and Ono, S. (1973) Biochim. Biophys. Acta 302, 362-375) did not give consistent results. A method is proposed based on certain properties of porcine pancreatic alpha-amylase, especially the non-interaction of the p-nitrophenyl moiety of the maltose derivative with subsites 1 and 2, and the o-nitrophenyl group which interacts in a similar way to a glucose residue at the reducing end, and on the grounds that the amylase-amylose complexes are of the productive type. In addition, binding energy differences were calculated from substrates with the same chain length. The subsite energy profile is characterized by a low value at subsite 3 which confirms this subsite as the catalytic one. Another consequence is that the hydrolysis rate constant of productive complexes (kintn) (where n is the number of glucose or glucose equivalent residues for a given substrate) varies with chain length which is in conflict with the hypothesis of Hiromi et al.  相似文献   

6.
4-Nitrophenyl penta-N-acetyl-β-chitopentaoside [(GlcNAc)5-pNP] was hydrolyzed by a family GH-19 class II barley chitinase, and the enzymatic reaction was monitored by real-time ESIMS. The wild-type enzyme hydrolyzed (GlcNAc)5-pNP producing predominantly (GlcNAc)3-pNP and a lesser amount of (GlcNAc)2-pNP, indicating that the (GlcNAc)5 portion of the substrate binds predominantly to subsites −2 ∼ +3 and less frequently to −3 ∼ +2. However, (GlcNAc)2-pNP was mainly produced from (GlcNAc)5-pNP by mutated enzymes, in which Trp72 and Trp82 located at +3/+4 were substituted with alanine (W72A and W72A/W82A), indicating that the (GlcNAc)5 portion of the substrate binds predominantly to subsites −3 ∼ +2 of the mutants. The mutations of the tryptophan residues resulted in a significant shift of the substrate-binding mode to the glycon side, supporting the idea that the indole side chain of Trp72 interacts with the 4-nitrophenyl moiety of the substrate at subsite +4.  相似文献   

7.
The environments of the binding subsites in Asp 101-modified lysozyme, in which glucosamine or ethanolamine is covalently bound to the carboxyl group of Asp 101, were investigated by chemical modification and nuclear magnetic resonance spectroscopy. Trp 62 in each of the native and the modified lysozymes was nitrophenylsulfenylated. The yield of the nitrophenylsulfenylated derivative from the lysozyme modified with glucosamine at Asp 101 (GlcN-lysozyme) was considerably lower than those from native lysozyme and from the lysozyme modified with ethanolamine at Asp 101 (EtN-lysozyme). These results suggest that Trp 62 in GlcN-lysozyme is less susceptible to nitrophenylsulfenylation. Kinetic analyses of the [Trp 62 and Asp 101]-doubly modified lysozymes indicated that the nitrophenylsulfenylation of Trp 62 in the native lysozyme, EtN-lysozyme, or GlcN-lysozyme decreased the sugar residue affinity at subsite C while increasing the binding free energy change by 2.7 kcal/mol, 1.5 kcal/mol, or 0.1 kcal/mol, respectively. Although the profile of tryptophan indole NH resonances in the 1H-NMR spectrum for EtN-lysozyme was not different from that for the native lysozyme, the indole NH resonance of Trp 62 in GlcN-lysozyme was apparently perturbed in comparison with that of native lysozyme. These results suggest that the environment of subsite C in GlcN-lysozyme is considerably different from those in native lysozyme and EtN-lysozyme. The glucosamine residue attached to Asp 101 may contact the sugar residue binding site of the lysozyme, affecting the environment of subsite C.  相似文献   

8.
Rice BGlu1 (Os3BGlu7) is a glycoside hydrolase family 1 β‐glucosidase that hydrolyzes cellooligosaccharides with increasing efficiency as the degree of polymerization (DP) increases from 2 to 6, indicating six subsites for glucosyl residue binding. Five subsites have been identified in X‐ray crystal structures of cellooligosaccharide complexes with its E176Q acid‐base and E386G nucleophile mutants. X‐ray crystal structures indicate that cellotetraose binds in a similar mode in BGlu1 E176Q and E386G, but in a different mode in the BGlu1 E386G/Y341A variant, in which glucosyl residue 4 (Glc4) interacts with Q187 instead of the eliminated phenolic group of Y341. Here, we found that the Q187A mutation has little effect on BGlu1 cellooligosaccharide hydrolysis activity or oligosaccharide binding in BGlu1 E176Q, and only slight effects on BGlu1 E386G glycosynthase activity. X‐ray crystal structures showed that cellotetraose binds in a different position in BGlu1 E176Q/Y341A, in which it interacts directly with R178 and W337, and the Q187A mutation had little effect on cellotetraose binding. Mutations of R178 and W337 to A had significant and nonadditive effects on oligosaccharide hydrolysis by BGlu1, pNPGlc cleavage and cellooligosaccharide inhibition of BGlu1 E176Q and BGlu1 E386G glycosynthase activity. Hydrolysis activity was partially rescued by Y341 for longer substrates, suggesting stacking of Glc4 on Y341 stabilizes binding of cellooligosaccharides in the optimal position for hydrolysis. This analysis indicates that complex interactions between active site cleft residues modulate substrate binding and hydrolysis.  相似文献   

9.
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   

10.
Cyclodextrin glycosyltransferase (CGTase) enzymes from various bacteria catalyze the formation of cyclodextrins from starch. The Bacillus stearothermophilus maltogenic alpha-amylase (G2-amylase is structurally very similar to CGTases, but converts starch into maltose. Comparison of the three-dimensional structures revealed two large differences in the substrate binding clefts. (i) The loop forming acceptor subsite +3 had a different conformation, providing the G2-amylase with more space at acceptor subsite +3, and (ii) the G2-amylase contained a five-residue amino acid insertion that hampers substrate binding at the donor subsites -3/-4 (Biochemistry, 38 (1999) 8385). In an attempt to change CGTase into an enzyme with the reaction and product specificity of the G2-amylase, which is used in the bakery industry, these differences were introduced into Thermoanerobacterium thermosulfurigenes CGTase. The loop forming acceptor subsite +3 was exchanged, which strongly reduced the cyclization activity, however, the product specificity was hardly altered. The five-residue insertion at the donor subsites drastically decreased the cyclization activity of CGTase to the extent that hydrolysis had become the main activity of enzyme. Moreover, this mutant produces linear products of variable sizes with a preference for maltose and had a strongly increased exo-specificity. Thus, CGTase can be changed into a starch hydrolase with a high exo-specificity by hampering substrate binding at the remote donor substrate binding subsites.  相似文献   

11.
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   

12.
Y Chan  P J Braun  D French  J F Robyt 《Biochemistry》1984,23(24):5795-5800
Hydrolysis of partially hydroxyethylated amylose by porcine pancreatic alpha-amylase gives rise to a number of hydroxyethylated di-, tri-, and tetrasaccharides, as well as larger products. No modified monosaccharides were detected. The structures of the products containing two to four D-glucose residues have been analyzed by chromatographic and enzymatic techniques. In no instance were these oligosaccharides modified in the reducing-end residue. The location of hydroxyethylated glucose residues within the oligosaccharides has been interpreted in terms of the ability of that (hydroxyethyl)glucose to bind productively at each of the five subsites of the enzyme active site. Results indicate that subsite 3, the subsite at which catalytic attack occurs, is especially sensitive to changes in the substrate and that unmodified glucose is required for productive binding at this subsite. Other subsites specifically allow binding of some (hydroxyethyl)glucose isomers, but not others. Hydroxyethylation is permitted at C-2, C-3, and C-6 for residues bound at subsite 1 and is permitted at C-6 and possibly at C-2 and C-3 for residues bound at subsite 5. However, substitution is permitted only at C-3 and C-6 for binding at subsite 2 and at C-2 and C-3 for binding at subsite 4.  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.
I Matsui  K Ishikawa  S Miyairi  S Fukui  K Honda 《Biochemistry》1992,31(22):5232-5236
The 210th lysine (K) residue in the Saccharomycopsis alpha-amylase (Sfamy) molecule was replaced by arginine (R) and asparagine (N) residues by site-directed mutagenesis. The influences of the replacements on the bond-cleavage pattern for several substrates were analyzed. Both mutant enzymes, K210R and K210N, cleave mainly the first glycosidic bond from the reducing end of maltotetraose (G4), while the native enzyme hydrolyzes mainly the second bond from the reducing end. We changed successfully the major cleavage point in the hydrolysis reaction of G4. The 8th subsite affinities of the K210R and K210N enzymes are calculated to be +2.52 and -0.01 kcal/mol, respectively, whereas that of the native enzyme is +3.32 kcal/mol as reported in the previous paper. These affinity values suggest that the K210 residue composes the 8th subsite, one of major subsites, and that a positively charged amino residue is necessary for the 8th subsite affinity. The K210N enzyme is found to be less active for short substrates like maltotetraose (G4) than for long substrates like amylose A (approximately G18). The reduced catalytic activity specifically for the short substrates is also attributable to the remarkable decrease in the affinity of the 8th subsite.  相似文献   

15.
The nonreducing end of the substrate-binding site of human salivary alpha-amylase contains two residues Trp58 and Trp59, which belong to beta2-alpha2 loop of the catalytic (beta/alpha)(8) barrel. While Trp59 stacks onto the substrate, the exact role of Trp58 is unknown. To investigate its role in enzyme activity the residue Trp58 was mutated to Ala, Leu or Tyr. Kinetic analysis of the wild-type and mutant enzymes was carried out with starch and oligosaccharides as substrates. All three mutants exhibited a reduction in specific activity (150-180-fold lower than the wild type) with starch as substrate. With oligosaccharides as substrates, a reduction in k(cat), an increase in K(m) and distinct differences in the cleavage pattern were observed for the mutants W58A and W58L compared with the wild type. Glucose was the smallest product generated by these two mutants in the hydrolysis oligosaccharides; in contrast, wild-type enzyme generated maltose as the smallest product. The production of glucose by W58L was confirmed from both reducing and nonreducing ends of CNP-labeled oligosaccharide substrates. The mutant W58L exhibited lower binding affinity at subsites -2, -3 and +2 and showed an increase in transglycosylation activity compared with the wild type. The lowered affinity at subsites -2 and -3 due to the mutation was also inferred from the electron density at these subsites in the structure of W58A in complex with acarbose-derived pseudooligosaccharide. Collectively, these results suggest that the residue Trp58 plays a critical role in substrate binding and hydrolytic activity of human salivary alpha-amylase.  相似文献   

16.
The contributions of the -3 subsite and a putative +3 subsite to substrate positioning in ChiA from Serratia marcescens have been investigated by comparing how ChiA and its -3 subsite mutant W167A interact with soluble substrates. The data show that Trp - GlcNAc stacking in the -3 subsite rigidifies the protein backbone supporting the formation of the intermolecular interaction network that is necessary for the recognition and positioning of the N-acetyl groups before the -1 subsite. The +3 subsite exhibits considerable substrate affinity that may promote endo-activity in ChiA and/or assist in expelling dimeric products from the +1 and +2 subsites during processive hydrolysis.  相似文献   

17.
Barley limit dextrinase (HvLD) of glycoside hydrolase family 13 is the sole enzyme hydrolysing α-1,6-glucosidic linkages from starch in the germinating seed. Surprisingly, HvLD shows 150- and 7-fold higher activity towards pullulan and β-limit dextrin, respectively, than amylopectin. This is investigated by mutational analysis of residues in the N-terminal CBM-21-like domain (Ser14Arg, His108Arg, Ser14Arg/His108Arg) and at the outer subsites +2 (Phe553Gly) and +3 (Phe620Ala, Asp621Ala, Phe620Ala/Asp621Ala) of the active site. The Ser14 and His108 mutants mimic natural LD variants from sorghum and rice with elevated enzymatic activity. Although situated about 40 Å from the active site, the single mutants had 15–40% catalytic efficiency compared to wild type for the three polysaccharides and the double mutant retained 27% activity for β-limit dextrin and 64% for pullulan and amylopectin. These three mutants hydrolysed 4,6-O-benzylidene-4-nitrophenyl-63-α-d-maltotriosyl-maltotriose (BPNPG3G3) with 51–109% of wild-type activity. The results highlight that the N-terminal CBM21-like domain plays a role in activity. Phe553 and the highly conserved Trp512 sandwich a substrate main chain glucosyl residue at subsite +2 of the active site, while substrate contacts of Phe620 and Asp621 at subsite +3 are less prominent. Phe553Gly showed 47% and 25% activity on pullulan and BPNPG3G3, respectively having a main role at subsite +2. By contrast at subsite +3, Asp621Ala increased activity on pullulan by 2.4-fold, while Phe620Ala/Asp621Ala retained only 7% activity on pullulan albeit showed 25% activity towards BPNPG3G3. This outcome supports that the outer substrate binding area harbours preference determinants for the branched substrates amylopectin and β-limit dextrin.  相似文献   

18.
In the three-dimensional structure of a rice class I chitinase (OsChia1b) determined recently, a loop structure (loop II) is located at the end of the substrate-binding cleft, and is thus suggested to be involved in substrate binding. In order to test this assumption, deletion of the loop II region from the catalytic domain of OsChia1b and replacement of Trp159 in loop II with Ala were carried out. The loop II deletion and the W159A mutation increased hydrolytic activity not only towards (GlcNAc)6 but also towards polysaccharide substrates. Similar results were obtained for kcat/Km values determined for substrate reduced-(GlcNAc)5. The two mutations shifted the splitting positions in (GlcNAc)6 to the reducing end side, but the shift was less intensive in the Trp mutant. Theoretical analysis of the reaction time course indicated that sugar residue affinity at the +3 subsite was reduced from -2 kcal/mol to +0.5 kcal/mol by loop II deletion. Reduced affinity at the +3 subsite might enhance the release of product fragments, resulting in higher turnover and higher enzymatic activities. Thus, we concluded that loop II is involved in sugar residue binding at the +3 subsite, but that Trp159 itself appears to contribute only partly to sugar residue interaction at the subsite.  相似文献   

19.
Depolymerization of polysaccharides is catalyzed by highly specific enzymes that promote hydrolysis of the scissile glycosidic bond by an activated water molecule. 1,3-1,4-beta-Glucanases selectively cleave beta-1,4 glycosidic bonds in 3-O-substituted glucopyranosyl units within polysaccharides with mixed linkage. The reaction follows a double-displacement mechanism by which the configuration of the anomeric C(1)-atom of the glucosyl unit in subsite -I is retained. Here we report the high-resolution crystal structure of the hybrid 1,3-1,4-beta-glucanase H(A16-M)(E105Q/E109Q) in complex with a beta-glucan tetrasaccharide. The structure shows four beta-d-glucosyl moieties bound to the substrate-binding cleft covering subsites -IV to -I, thus corresponding to the reaction product. The ten active-site residues Asn26, Glu63, Arg65, Phe92, Tyr94, Glu105, Asp107, Glu109, Asn182 and Trp184 form a network of hydrogen bonds and hydrophobic stacking interactions with the substrate. These residues were previously identified by mutational analysis as significant for stabilization of the enzyme-carbohydrate complex, with Glu105 and Glu109 being the catalytic residues. Compared to the Michaelis complex model, the tetrasaccharide moiety is slightly shifted toward that part of the cleft binding the non-reducing end of the substrate, but shows previously unanticipated strong stacking interactions with Phe92 in subsite -I. A number of specific hydrogen-bond contacts between the enzyme and the equatorial O(2), O(3) and O(6) hydroxyl groups of the glucosyl residues in subsites -I, -II and -III are the structural basis for the observed substrate specificity of 1,3-1,4-beta-glucanases. Kinetic analysis of enzyme variants with the all beta-1,3 linked polysaccharide laminarin identified key residues mediating substrate specificity in good agreement with the structural data. The comparison with structures of the apo-enzyme H(A16-M) and a covalent enzyme-inhibitor (E.I) complex, together with kinetic and mutagenesis data, yields new insights into the structural requirements for substrate binding and catalysis. A detailed view of enzyme-carbohydrate interactions is presented and mechanistic implications are discussed.  相似文献   

20.
The action pattern of human salivary amylase (HSA) was examined by utilising as model substrates 2-chloro-4-nitrophenyl (CNP) beta-glycosides of maltooligosaccharides of dp 4-8 and some 4-nitrophenyl (NP) derivatives modified at the nonreducing end with a 4,6-O-benzylidene (Bnl) group. The product pattern and cleavage frequency were investigated by product analysis using HPLC. The results revealed that the binding region in HSA is longer than five subsites usually considered in the literature and suggested the presence of at least six subsites; four glycone binding sites (-4, -3, -2, -1) and two aglycone binding sites (+1, +2). In the ideal arrangement, the six subsites are filled by a glucosyl unit and the release of maltotetraose (G4) from the nonreducing end is dominant. The benzylidene group was also recognisable by subsites (-3) and (-4). The binding modes of the benzylidene derivatives indicated a favourable interaction between the Bnl group and subsite (-3) and an unfavourable one with subsite (-4). Thus, subsite (-4) must be more hydrophylic than hydrophobic. As compared with the action of porcine pancreatic alpha-amylase (PPA) on the same substrates, the results showed differences in the three-dimensional structure of active sites of HSA and PPA.  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