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.
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   

2.
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   

3.
The X-ray crystal structure of a catalytic site mutant of beta-amylase, E172A (Glu172 --> Ala), from Bacillus cereus var. mycoides complexed with a substrate, maltopentaose (G5), and the wild-type enzyme complexed with maltose were determined at 2.1 and 2.0 A resolution, respectively. Clear and continuous density corresponding to G5 was observed in the active site of E172A, and thus, the substrate, G5, was not hydrolyzed. All glucose residues adopted a relaxed (4)C(1) conformation, and the conformation of the maltose unit for Glc2 and Glc3 was much different from those of other maltose units, where each glucose residue of G5 is named Glc1-Glc5 (Glc1 is at the nonreducing end). A water molecule was observed 3.3 A from the C1 atom of Glc2, and 3.0 A apart from the OE1 atom of Glu367 which acts as a general base. In the wild-type enzyme-maltose complex, two maltose molecules bind at subsites -2 and -1 and at subsites +1 and +2 in tandem. The conformation of the maltose molecules was similar to that of the condensation product of soybean beta-amylase, but differed from that of G5 in E172A. When the substrate flips between Glc2 and Glc3, the conformational energy of the maltose unit was calculated to be 20 kcal/mol higher than that of the cis conformation by MM3. We suggest that beta-amylase destabilizes the bond that is to be broken in the ES complex, decreasing the activation energy, DeltaG(++), which is the difference in free energy between this state and the transition state.  相似文献   

4.
5.
We investigated the effects of guanidine hydrochloride (GuHCl) and high pressure on the conformational flexibility of the active site of sweet potato beta-amylase by monitoring the sulfhydryl reaction and the enzymatic activity. The reactivity of Cys345 at the active site, one of six inert half cystine residues of this enzyme, was enhanced by GuHCl at concentrations below 0.5 M. A GuHCl-induced change of the active site was also observed through an intensity change in the near-UV circular dichroism (CD) spectrum. On the other hand, the native conformation of sweet potato beta-amylase observed through fluorescence polarization, far-UV CD spectrum and intrinsic fluorescence was not influenced by GuHCl at concentrations below 0.5 M. Therefore, Cys345 reaction caused by GuHCl was due to an alteration of the local conformation of the active site. GuHCl-induced reaction of Cys345, located in the vicinity of subsites 3 and 4, is attributed to enhanced subsite flexibility, which is responsible for substrate slipping in a single-chain attack mechanism. Due to the flexible conformation, the local region of the subsite is more susceptible to GuHCl perturbation than the molecule overall. The enzymatic activity of sweet potato beta-amylase was reversibly inhibited by GuHCl at concentrations below 0.5 M, and kinetic analysis of the enzymatic mechanism showed that GuHCl decreases the kcat value. High pressure below 400 MPa also inactivated sweet potato beta-amylase with an increase in Cys345 reactivity. These findings indicated that excessively enhanced subsite flexibility reduced the enzymatic activity of sweet potato beta-amylase.  相似文献   

6.
7.
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   

8.
Maltosyltransferase (MTase) from the hyperthermophile Thermotoga maritima represents a novel maltodextrin glycosyltransferase acting on starch and malto-oligosaccharides. It catalyzes the transfer of maltosyl units from alpha-1,4-linked glucans or malto-oligosaccharides to other alpha-1,4-linked glucans, malto-oligosaccharides or glucose. It belongs to the glycoside hydrolase family 13, which represents a large group of (beta/alpha)(8) barrel proteins sharing a similar active site structure. The crystal structures of MTase and its complex with maltose have been determined at 2.4 A and 2.1 A resolution, respectively. MTase is a homodimer, each subunit of which consists of four domains, two of which are structurally homologous to those of other family 13 enzymes. The catalytic core domain has the (beta/alpha)(8) barrel fold with the active-site cleft formed at the C-terminal end of the barrel. Substrate binding experiments have led to the location of two distinct maltose-binding sites; one lies in the active-site cleft, covering subsites -2 and -1; the other is located in a pocket adjacent to the active-site cleft. The structure of MTase, together with the conservation of active-site residues among family 13 glycoside hydrolases, are consistent with a common double-displacement catalytic mechanism for this enzyme. Analysis of maltose binding in the active site reveals that the transfer of dextrinyl residues longer than a maltosyl unit is prevented by termination of the active-site cleft after the -2 subsite by the side-chain of Lys151 and the stretch of residues 314-317, providing an explanation for the strict transfer specificity of MTase.  相似文献   

9.
Fibrobacter succinogenes 1,3-1,4-beta-D-glucanase (Fsbeta-glucanase) catalyzes the specific hydrolysis of beta-1,4 glycosidic bonds adjacent to beta-1,3 linkages in beta-D-glucans or lichenan. This is the first report to elucidate the crystal structure of a truncated Fsbeta-glucanase (TFsbeta-glucanase) in complex with beta-1,3-1,4-cellotriose, a major product of the enzyme reaction. The crystal structures, at a resolution of 2.3 angstroms, reveal that the overall fold of TFsbeta-glucanase remains virtually unchanged upon sugar binding. The enzyme accommodates five glucose residues, forming a concave active cleft. The beta-1,3-1,4-cellotriose with subsites -3 to -1 bound to the active cleft of TFsbeta-glucanase with its reducing end subsite -1 close to the key catalytic residues Glu56 and Glu60. All three subsites of the beta-1,3-1,4-cellotriose adopted a relaxed C(1)4 conformation, with a beta-1,3 glycosidic linkage between subsites -2 and -1, and a beta-1,4 glycosidic linkage between subsites -3 and -2. On the basis of the enzyme-product complex structure observed in this study, a catalytic mechanism and substrate binding conformation of the active site of TFsbeta-glucanase is proposed.  相似文献   

10.
Methyl 2,4-dinitrophenyl disulfide (MDPS) is shown to be an effective methanethiolating reagent for sulfhydryl groups in proteins via thiol-disulfide exchange reaction. It reacts with the two reactive sulfhydryl groups (SH1 and SH2) in soybean beta-amylase. A decrease of the enzymatic activity accompanies the methanethiolation of SH2. After complete methanethiolation of SH2, the modified enzyme still has 9% of the initial activity. Modification of SH2 with cyanide and iodoacetamide reduces the enzymatic activity to 65 and 2% of the initial activity, respectively. Apparently, the residual activity depends upon the size of the substituent at SH2. The modified enzymes still have the almost same Km values for amylopectin and Kd values for enzyme-maltose and enzyme-cyclohexaamylose complexes as the native enzyme. In contrast to maltose and cyclohexaamylose, the Kd value of the enzyme-glucose complex increases in the order of cyanide-, MDPS-, and iodoacetamide-modified enzymes, indicating that SH2 is located near the binding site of glucose. It is proposed from the subsite structure of soybean beta-amylase that the position of SH2 and the glucose binding site is around subsite 1, where the nonreducing ends of the substrate bind productively.  相似文献   

11.
Interaction of soybean beta-amylase with glucose   总被引:1,自引:0,他引:1  
The interaction of soybean beta-amylase with glucose was investigated by inhibition kinetics studies and spectroscopic measurements. The inhibition type, inhibitor constant (Ki) and dissociation constant (Kd) of beta-amylase-glucose complex were dependent on pH. At pH 8.0, glucose behaved as a competitive inhibitor (Ki = 34 mM). Binding of glucose produced a characteristic difference spectrum and a change of circular dichroism (CD) at pH 8.1. By using difference absorbance at 292 nm and difference ellipticity at 290 nm, Kd values for beta-amylase-glucose complex were determined to be 45 and 46 mM, respectively. In contrast to pH 8.0, glucose behaved as a mixed-type inhibitor (Ki = 320 mM) at pH 5.4. The Kd values obtained from the difference spectrum were increased by lowering the pH from 8. The pH dependence of the Ki and Kd values suggested that one ionizable group of pK = 8.0, which is shifted to 6.9 by the binding of glucose, controls the binding affinity of glucose. The binding of glucose competed with the binding of cyclohexaamylose and maltose at pH 8.0. The modification of SH groups of the enzyme affected the binding of glucose but did not affect the binding of maltose or cyclohexaamylose at pH 8.0. It was concluded from these results that the binding site of glucose is different from that of maltose and cyclohexaamylose. Presumably, glucose may bind to the subsite 1 of soybean beta-amylase.  相似文献   

12.
Hydrolysis of small substrates (maltose, maltotriose and o-nitrophenylmaltoside) catalysed by porcine pancreatic alpha-amylase was studied from a kinetic viewpoint over a wide range of substrate concentrations. Non-linear double-reciprocal plots are obtained at high maltose, maltotriose and o-nitrophenylmaltoside concentrations indicating typical substrate inhibition. These results are consistent with the successive binding of two molecules of substrate per enzyme molecule with dissociation constants Ks1 and Ks2. The Hill plot, log [v/(V-v)] versus log [S], is clearly biphasic and allows the dissociation constants of the ES1 and ES2 complexes to be calculated. Maltose and maltotriose are inhibitors of the amylase-catalysed amylose and o-nitrophenylmaltoside hydrolysis. The inhibition is of the competitive type. The (apparent) inhibition constant Kiapp varies with the inhibitor concentration. These results are also consistent with the successive binding of at least two molecules of maltose or maltotriose per amylase molecule with the dissociation constants Ki1 and Ki2. These inhibition studies show that small substrates and large polymeric ones are hydrolysed at the same catalytic site(s). The values of the dissociation constants Ks1 and Ki1 of the maltose-amylase complexes are identical. According to the five-subsite energy profile previously determined, at low concentration, maltose (as substrate and as inhibitor) binds to the same two sites (4,5) or (3,4), maltotriose (as substrate and as inhibitor) and o-nitrophenyl-maltoside (as substrate) bind to the same three subsites (3,4,5). The dissociation constants Ks2 and Ki2 determined at high substrate and inhibitor concentration are consistent with the binding of the second ligand molecule at a single subsite. The binding mode of the second molecule of maltose (substrate) and o-nitrophenylmaltoside remains uncertain, very likely because of the inaccuracy due to simplifications in the calculations of the subsite binding energies. No binding site(s) outside the catalytic one has been taken into account in this model.  相似文献   

13.
The crystal structure of beta-amylase from Bacillus cereus var. mycoides was determined by the multiple isomorphous replacement method. The structure was refined to a final R-factor of 0.186 for 102,807 independent reflections with F/sigma(F) > or = 2.0 at 2.2 A resolution with root-mean-square deviations from ideality in bond lengths, and bond angles of 0.014 A and 3.00 degrees, respectively. The asymmetric unit comprises four molecules exhibiting a dimer-of-dimers structure. The enzyme, however, acts as a monomer in solution. The beta-amylase molecule folds into three domains; the first one is the N-terminal catalytic domain with a (beta/alpha)8 barrel, the second one is the excursion part from the first one, and the third one is the C-terminal domain with two almost anti-parallel beta-sheets. The active site cleft, including two putative catalytic residues (Glu172 and Glu367), is located on the carboxyl side of the central beta-sheet in the (beta/alpha)8 barrel, as in most amylases. The active site structure of the enzyme resembles that of soybean beta-amylase with slight differences. One calcium ion is bound per molecule far from the active site. The C-terminal domain has a fold similar to the raw starch binding domains of cyclodextrin glycosyltransferase and glucoamylase.  相似文献   

14.
Sierks MR  Svensson B 《Biochemistry》2000,39(29):8585-8592
Molecular recognition using a series of deoxygenated maltose analogues was used to determine the substrate transition-state binding energy profiles of 10 single-residue mutants at the active site of glucoamylase from Aspergillus niger. The individual contribution of each substrate hydroxyl group to transition-state stabilization with the wild type and each mutant GA was determined from the relation Delta(DeltaG()) = -RT ln[(k(cat)/K(M))(x)/(k(cat)/K(M))(y)], where x represents either a mutant enzyme or substrate analogue and y the wild-type enzyme or parent substrate. The resulting binding energy profiles indicate that disrupting an active site hydrogen bond between enzyme and substrate, as identified in crystal structures, not only sharply reduces or eliminates the energy contributed from that particular hydrogen bond but also perturbs binding contributions from other substrate hydroxyl groups. Replacing the active site acidic groups, Asp55, Glu180, or Asp309, with the corresponding amides, and the neutral Trp178 with the basic Arg, all substantially reduced the binding energy contribution of the 4'- and 6'-OH groups of maltose at subsite -1, even though both Glu180 and Asp309 are localized at subsite 1. In contrast, the substitution, Asp176 --> Asn, located near subsites -1 and 1, did not substantially perturb any of the individual hydroxyl group binding energies. Similarly, the substitutions Tyr116 --> Ala, Ser119 --> Tyr, or Trp120 --> Phe also did not substantially alter the energy profiles even though Trp120 has a critical role in directing conformational changes necessary for activity. Since the mutations at Trp120 and Asp176 reduced k(cat) values by 50- and 12-fold, respectively, a large effect on k(cat) is not necessarily accompanied by changes in hydroxyl group binding energy contributions. Two substitutions, Asn182 --> Ala and Tyr306 --> Phe, had significant though small effects on interactions with 3- and 4'-OH, respectively. Binding interactions between the enzyme and the glucosyl group in subsite -1, particularly with the 4'- and 6'-OH groups, play an important role in substrate binding, while subsite 1 interactions may play a more important role in product release.  相似文献   

15.
The crystal structures of Klebsiella pneumoniae pullulanase and its complex with glucose (G1), maltose (G2), isomaltose (isoG2), maltotriose (G3), or maltotetraose (G4), have been refined at around 1.7-1.9A resolution by using a synchrotron radiation source at SPring-8. The refined models contained 920-1052 amino acid residues, 942-1212 water molecules, four or five calcium ions, and the bound sugar moieties. The enzyme is composed of five domains (N1, N2, N3, A, and C). The N1 domain was clearly visible only in the structure of the complex with G3 or G4. The N1 and N2 domains are characteristic of pullulanase, while the N3, A, and C domains have weak similarity with those of Pseudomonas isoamylase. The N1 domain was found to be a new type of carbohydrate-binding domain with one calcium site (CBM41). One G1 bound at subsite -2, while two G2 bound at -1 approximately -2 and +2 approximately +1, two G3, -1 approximately -3 and +2 approximately 0', and two G4, -1 approximately -4 and +2 approximately -1'. The two bound G3 and G4 molecules in the active cleft are almost parallel and interact with each other. The subsites -1 approximately -4 and +1 approximately +2, including catalytic residues Glu706 and Asp677, are conserved between pullulanase and alpha-amylase, indicating that pullulanase strongly recognizes branched point and branched sugar residues, while subsites 0' and -1', which recognize the non-reducing end of main-chain alpha-1,4 glucan, are specific to pullulanase and isoamylase. The comparison suggested that the conformational difference around the active cleft, together with the domain organization, determines the different substrate specificities between pullulanase and isoamylase.  相似文献   

16.
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   

17.
Cyclohexadextrin and maltose bound to soybean beta-amylase and affected the environments of tryptophan and tyrosine residues, producing characteristic difference spectra in the ultraviolet region. The difference spectrum produced by cyclohexadextrin, a competitive inhibitor, had peaks at 285, 292, and 299 nm, while that by maltose, a reaction product, had peaks at 285 and 292 nm and a small trough at around 300 nm. By using the peaks at 292 and 299 nm, the dissociation constants of enzyme-cyclohexadextrin and enzyme-maltose complexes were calculated to be 0.35 mM and 8.1 mM, respectively. The effects of modification of SH groups of beta-amylase on the interaction of the enzyme with these sugars were examined by using beta-amylase carboxymethylated at the SH1 site and the enzyme modified at SH1 and SH2 sites with iodoacetamide or with 5,5'-dithiobis-(2-nitrobenzoic acid) (DTNB). The dissociation constants of the enzyme-cyclohexadextrin and enzyme-maltose complexes were not changed by the modification of these SH groups, but the modification of SH2, the so-called essential SH group of soybean beta-amylase, strongly affected the difference spectra produced by maltose. The spectrophotometric titration of beta-amylase by cyclohexadextrin in the presence of maltose showed that cyclohexadextrin and maltose bind to the enzyme competitively, regardless of the modification of SH2. These results indicated that SH2 is located near the binding site of cyclohexadextrin and maltose, but is not involved in the binding of these sugars.  相似文献   

18.
A modified amylose containing 10% of tritiated D-allose residues has been hydrolyzed by porcine pancreatic alpha amylase (PPA). This reaction produced a number of radioactive oligosaccharides of low molecular weight, including modified mono-, di-, and tri-saccharides, as well as larger products. Analysis of these products by chemical and enzymic methods identified D-allose, two isomers of modified maltose, and isomers of modified maltotriose. These results may be interpreted in terms of current PPA models to indicate that D-allose residues may be productively bound at all five subsites of the active site of the enzyme. The distribution of modified residues in these products, however, further suggests that productive binding of D-allose at the subsite where catalytic attack occurs (subsite 3) is less favorable than binding of D-glucose. These results are compared with results of a series of PPA substrates having modifications at C-3 and at other positions. Trends observed in enzyme hydrolysis of these modified substrates reflect factors that contribute to PPA catalysis, with respect to steric, electronic, and hydrogen-bonding interactions between enzyme and substrate.  相似文献   

19.
Glycoside hydrolase family (GH) 11 xylanase A from Bacillus subtilis (BsXynA) was subjected to site-directed mutagenesis to probe the role of aglycon active site residues with regard to activity, binding of decorated substrates and hydrolysis product profile. Targets were those amino acids identified to be important by 3D structure analysis of BsXynA in complex with substrate bound in the glycon subsites and the + 1 aglycon subsite. Several aromatic residues in the aglycon subsites that make strong substrate–protein interactions and that are indispensable for enzyme activity, were also important for the specificity of the xylanase. In the + 2 subsite of BsXynA, Tyr65 and Trp129 were identified as residues that are involved in the binding of decorated substrates. Most interestingly, replacement of Tyr88 by Ala in the + 3 subsite created an enzyme able to produce a wider variety of hydrolysis products than wild type BsXynA. The contribution of the + 3 subsite to the substrate specificity of BsXynA was established more in detail by mapping the enzyme binding site of the wild type xylanase and mutant Y88A with labelled xylo-oligosaccharides. Also, the length of the cord – a long loop flanking the aglycon subsites of GH11 xylanases – proved to impact the hydrolytic action of BsXynA. The aglycon side of the active site cleft of BsXynA, therefore, offers great potential for engineering and design of xylanases with a desired specificity.  相似文献   

20.
Glycoside hydrolase family 77 (GH77) belongs to the alpha-amylase superfamily (Clan H) together with GH13 and GH70. GH77 enzymes are amylomaltases or 4-alpha-glucanotransferases, involved in maltose metabolism in microorganisms and in starch biosynthesis in plants. Here we characterized the amylomaltase from the hyperthermophilic bacterium Thermus thermophilus HB8 (Tt AMase). Site-directed mutagenesis of the active site residues (Asp293, nucleophile; Glu340, general acid/base catalyst; Asp395, transition state stabilizer) shows that GH77 Tt AMase and GH13 enzymes share the same catalytic machinery. Quantification of the enzyme's transglycosylation and hydrolytic activities revealed that Tt AMase is among the most efficient 4-alpha-glucanotransferases in the alpha-amylase superfamily. The active site contains at least seven substrate binding sites, subsites -2 and +3 favoring substrate binding and subsites -3 and +2 not, in contrast to several GH13 enzymes in which subsite +2 contributes to oligosaccharide binding. A model of a maltoheptaose (G7) substrate bound to the enzyme was used to probe the details of the interactions of the substrate with the protein at acceptor subsites +2 and +3 by site-directed mutagenesis. Substitution of the fully conserved Asp249 with a Ser in subsite +2 reduced the activity 23-fold (for G7 as a substrate) to 385-fold (for maltotriose). Similar mutations reduced the activity of alpha-amylases only up to 10-fold. Thus, the characteristics of acceptor subsite +2 represent a main difference between GH13 amylases and GH77 amylomaltases.  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