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.
Dextran glucosidase from Streptococcus mutans (SMDG) and Bacillus oligo-1,6-glucosidases, members of glycoside hydrolase family 13 enzymes, have the high sequence similarity. Each of them is specific to alpha-1,6-glucosidic linkage at the non-reducing end of substrate to liberate glucose. The activities toward long isomaltooligosaccharides were different in both enzymes, in which SMDG and oligo-1,6-glucosidase showed high and low activities, respectively. We determined the structural elements essential for high activity toward long-chain substrate. From conformational comparison between SMDG and B. cereus oligo-1,6-glucosidase (three-dimensional structure has been solved), Trp238 and short beta-->alpha loop 4 of SMDG were considered to contribute to the high activity to long-chain substrate. W238A had similar kcat/Km value for isomaltotriose to that for isomaltose, suggesting that the affinity of subsite +2 was decreased by Trp238 replacement. Trp238 mutants as well as the chimeric enzyme having longer beta-->alpha loop 4 of B. subtilis oligo-1,6-glucosidase showed lower preference for long-chain substrates, indicating that both Trp238 and short beta-->alpha loop 4 were important for high activity to long-chain substrates.  相似文献   

2.
Three active site residues (Asp199, Glu255, Asp329) and two substrate-binding site residues (His103, His328) of oligo-1,6-glucosidase (EC 3.2.1.10) from Bacillus cereus ATCC7064 were identified by site-directed mutagenesis. These residues were deduced from the X-ray crystallographic analysis and the comparison of the primary structure of the oligo-1,6-glucosidase with those of Saccharomyces carlsbergensis α-glucosidase, Aspergillus oryzae α-amylase and pig pancreatic α-amylase which act on α-1,4-glucosidic linkages. The distances between these putative residues of B. cereus oligo-1,6-glucosidase calculated from the X-ray analysis data closely resemble those of A. oryzae α-amylase and pig pancreatic α-amylase. A single mutation of Asp199→Asn, Glu255→Gln, or Asp329→Asn resulted in drastic reduction in activity, confirming that three residues are crucial for the reaction process of α-1,6-glucosidic bond cleavage. Thus, it is identified that the basic mechanism of oligo-1,6-glucosidase for the hydrolysis of α-1,6-glucosidic linkage is essentially the same as those of other amylolytic enzymes belonging to Family 13 (α-amylase family). On the other hand, mutations of histidine residues His103 and His328 resulted in pronounced dissimilarity in catalytic function. The mutation His328→Asn caused the essential loss in activity, while the mutation His103→Asn yielded a mutant enzyme that retained 59% of the κ0/Km of that for the wild-type enzyme. Since mutants of other α-amylases acting on α-1,4-glucosidic bond linkage lost most of their activity by the site-directed mutagenesis at their equivalent residues to His103 and His328, the retaining of activity by Hisl03→Asn mutation in B. cereus oligo-1,6-glucosidase revealed the distinguished role of His103 for the hydrolysis of α-1,6-glucosidic bond linkage.  相似文献   

3.
Highly branched α-glucan molecules exhibit low digestibility for α-amylase and glucoamylase, and abundant in α-(1→3)-, α-(1→6)-glucosidic linkages and α-(1→6)-linked branch points where another glucosyl chain is initiated through an α-(1→3)-linkage. From a culture supernatant of Paenibacillus sp. PP710, we purified α-glucosidase (AGL) and α-amylase (AMY), which were involved in the production of highly branched α-glucan from maltodextrin. AGL catalyzed the transglucosylation reaction of a glucosyl residue to a nonreducing-end glucosyl residue by α-1,6-, α-1,4-, and α-1,3-linkages. AMY catalyzed the hydrolysis of the α-1,4-linkage and the intermolecular or intramolecular transfer of maltooligosaccharide like cyclodextrin glucanotransferase (CGTase). It also catalyzed the transfer of an α-1,4-glucosyl chain to a C3- or C4-hydroxyl group in the α-1,4- or α-1,6-linked nonreducing-end residue or the α-1,6-linked residue located in the other chains. Hence AMY was regarded as a novel enzyme. We think that the mechanism of formation of highly branched α-glucan from maltodextrin is as follows: α-1,6- and α-1,3-linked residues are generated by the transglucosylation of AGL at the nonreducing ends of glucosyl chains. Then AMY catalyzes the transfer of α-1,4-chains to C3- or C4-hydroxyl groups in the α-1,4- or α-1,6-linked residues generated by AGL. Thus the concerted reactions of both AGL and AMY are necessary to produce the highly branched α-glucan from maltodextrin.  相似文献   

4.
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   

5.
Barley limit dextrinase [Hordeum vulgare limit dextrinase (HvLD)] catalyzes the hydrolysis of α-1,6 glucosidic linkages in limit dextrins. This activity plays a role in starch degradation during germination and presumably in starch biosynthesis during grain filling. The crystal structures of HvLD in complex with the competitive inhibitors α-cyclodextrin (CD) and β-CD are solved and refined to 2.5 Å and 2.1 Å, respectively, and are the first structures of a limit dextrinase. HvLD belongs to glycoside hydrolase 13 family and is composed of four domains: an immunoglobulin-like N-terminal eight-stranded β-sandwich domain, a six-stranded β-sandwich domain belonging to the carbohydrate binding module 48 family, a catalytic (β/α)8-like barrel domain that lacks α-helix 5, and a C-terminal eight-stranded β-sandwich domain of unknown function. The CDs are bound at the active site occupying carbohydrate binding subsites + 1 and + 2. A glycerol and three water molecules mimic a glucose residue at subsite − 1, thereby identifying residues involved in catalysis. The bulky Met440, a unique residue at its position among α-1,6 acting enzymes, obstructs subsite − 4. The steric hindrance observed is proposed to affect substrate specificity and to cause a low activity of HvLD towards amylopectin. An extended loop (Asp513-Asn520) between β5 and β6 of the catalytic domain also seems to influence substrate specificity and to give HvLD a higher affinity for α-CD than pullulanases. The crystal structures additionally provide new insight into cation sites and the concerted action of the battery of hydrolytic enzymes in starch degradation.  相似文献   

6.
Dextran glucosidase from Streptococcus mutans (SMDG), an exo-type glucosidase of glycoside hydrolase (GH) family 13, specifically hydrolyzes an α-1,6-glucosidic linkage at the non-reducing ends of isomaltooligosaccharides and dextran. SMDG shows the highest sequence similarity to oligo-1,6-glucosidases (O16Gs) among GH family 13 enzymes, but these enzymes are obviously different in terms of substrate chain length specificity. SMDG efficiently hydrolyzes both short-and long-chain substrates, while O16G acts on only short-chain substrates. We focused on this difference in substrate specificity between SMDG and O16G, and elucidated the structure-function relationship of substrate chain length specificity in SMDG. Crystal structure analysis revealed that SMDG consists of three domains, A, B, and C, which are commonly found in other GH family 13 enzymes. The structural comparison between SMDG and O16G from Bacillus cereus indicated that Trp238, spanning subsites +1 and +2, and short βα loop 4, are characteristic of SMDG, and these structural elements are predicted to be important for high activity toward long-chain substrates. The substrate size preference of SMDG was kinetically analyzed using two mutants: (i) Trp238 was replaced by a smaller amino acid, alanine, asparagine or proline; and (ii) short βα loop 4 was exchanged with the corresponding loop of O16G. Mutant enzymes showed lower preference for long-chain substrates than wild-type enzyme, indicating that these structural elements are essential for the high activity toward long-chain substrates, as implied by structural analysis.  相似文献   

7.
Highly branched α-glucan molecules exhibit low digestibility for α-amylase and glucoamylase, and abundant in α-(1→3)-, α-(1→6)-glucosidic linkages and α-(1→6)-linked branch points where another glucosyl chain is initiated through an α-(1→3)-linkage. From a culture supernatant of Paenibacillus sp. PP710, we purified α-glucosidase (AGL) and α-amylase (AMY), which were involved in the production of highly branched α-glucan from maltodextrin. AGL catalyzed the transglucosylation reaction of a glucosyl residue to a nonreducing-end glucosyl residue by α-1,6-, α-1,4-, and α-1,3-linkages. AMY catalyzed the hydrolysis of the α-1,4-linkage and the intermolecular or intramolecular transfer of maltooligosaccharide like cyclodextrin glucanotransferase (CGTase). It also catalyzed the transfer of an α-1,4-glucosyl chain to a C3- or C4-hydroxyl group in the α-1,4- or α-1,6-linked nonreducing-end residue or the α-1,6-linked residue located in the other chains. Hence AMY was regarded as a novel enzyme. We think that the mechanism of formation of highly branched α-glucan from maltodextrin is as follows: α-1,6- and α-1,3-linked residues are generated by the transglucosylation of AGL at the nonreducing ends of glucosyl chains. Then AMY catalyzes the transfer of α-1,4-chains to C3- or C4-hydroxyl groups in the α-1,4- or α-1,6-linked residues generated by AGL. Thus the concerted reactions of both AGL and AMY are necessary to produce the highly branched α-glucan from maltodextrin.  相似文献   

8.
To understand the role of His and Glu in the catalytic activity of Bacillus licheniformis α-amylase (BLA), His235 was replaced with Glu. The mutant enzyme, H235E, was characterized in terms of its mode of action using labeled and unlabeled maltooctaose (Glc8). H235E predominantly produced maltotridecaose (Glc13) from Glc8, exhibiting high substrate transglycosylation activity, with Km = 0.38 mM and kcat/Km = 20.58 mM−1 s−1 for hydrolysis, and Km2 = 18.38 mM and kcat2/Km2 = 2.57 mM−1 s−1 for transglycosylation, while the wild-type BLA exhibited high hydrolysis activity exclusively. Glu235—located on a wide open groove near subsite +1—is likely involved in transglycosylation via formation of an α-1,4-glycosidic linkage and may recognize and stabilize the non-reducing end glucose of the acceptor molecule.  相似文献   

9.
The dextransucrase produced by Streptococcus bovis 148 was purified about 20.8-fold to electrophoretic homogeneity (specific activity 246.3 units/mg) from the culture supernatant. Molecular weight of the native enzyme was estimated to be 600 kDa, whereas the molecular weight per subunit was 150 kDa. The dextran synthesized consisted chiefly of α-1,6-glucosidic linkage, containing small numbers of branches that had the α-1,3-glucosidic linkage.  相似文献   

10.
A p-nitrophenyl-alpha-D-glucopyranoside-hydrolyzing alpha-glucosidase of a thermophile, Bacillus thermoglucosidius KP 1006, was purified to an electrophoretically-homogeneous state. Its molecular weight was estimated as 60 000 by gel electrophoresis. The molecular activity (ko) and the Km value at 60 degrees C and pH 6.8 for p-nitrophenyl-alpha-D-glucopyranoside were 233 s-1 and 0.24 mM, respectively. The enzyme cleft the non-reducing terminal alpha-1,6-glucosidic bonds of isomaltose, panose, isomaltotriose, isomaltotetraose, and isomaltopentaose. The ko values were 72.4, 194, 208, 233 and 167 s-1, and the Km values were 3.3, 9.5, 11, 13 and 21 mM, respectively. Each isomaltosaccharide was hydrolyzed to glucose by the cleavage of single glucose units from its nonreducing end. The present study suggests that the enzyme is an oligo-1,6-glucosidase (dextrin 6-alpha-glucanohydrolase, EC 3.2.1.10) and an exo-glucosidase.  相似文献   

11.
The action of α-1,6-glucan glucohydrolase on α-(1→6)-D-glucosidic linkages in oligosaccharides that also contain an α-(1→2)-, α-(1→3)-, or α-(1→4)-D-glucosidic linkage has been investigated. The enzyme could hydrolyse α-(1→6)-D-glucosidic linkages from the non-reducing end, including those adjacent to an anomalous linkage. α-(1→6)-D-Glucosidic linkages at branch points were not hydrolysed, and the enzyme could neither hydrolyse nor by-pass the anomalous linkages. These properties of α-1,6-glucan glucohydrolase explain the limited hydrolysis of dextrans by the exo-enzyme. Hydrolysis of the main chain of α-(1→6)-D-glucans will always stop one D-glucose residue away from a branch point. The extent of hydrolysis by α-1,6-glucan glucohydrolase of some oligosaccharide products of the action on dextran of Penicillium funiculosum and P. lilacinum dextranase, respectively, has been compared. Differences in the specificity of the two endo-dextranases were revealed. The Penicillium enzymes may hydrolyse dextran B-512 to produce branched oligosaccharides that retain the same 1-unit and 2-unit side-chains that occur in dextran.  相似文献   

12.
α,α-Trehalose is of interest because of its cryoprotective and antidessicant properties, and because it possesses various technical anomalies such as 13C NMR spectra that give misleading indications of intramolecular structural symmetry. It is a non-reducing disaccharide, with the glycosidic oxygen atom shared by the anomeric carbon atoms of the two glucose rings, and is therefore subject to a proposed ‘overlapping’ exo-anomeric effect. We report here a study of the electron density of trehalose with X-ray diffraction and quantum mechanics calculations, similar to a recent study of sucrose, also a non-reducing molecule. In particular we studied the electron density around the glycosidic linkage and the hydrogen bonding with both deformation density and Atoms in Molecules (AIM) analyses. A total of 129,952 single crystal X-ray intensity measurements were collected on α,α-trehalose dihydrate to a resolution of sin θ/λ = 1.18 Å−1 at 100 K and refined with an aspherical multipole model to a final agreement factor of R1 = 0.0160. Wavefunctions were calculated at three levels of theory. Redistribution of electron density due to anomeric effects was reduced in trehalose, compared to sucrose. Five new C-H?O hydrogen bonds were confirmed with bond critical points and bond paths from AIM analyses, as were the previously proposed O-H?O hydrogen bonds.  相似文献   

13.
Di-O-α-maltosyl-β-cyclodextrin ((G2)2-β-CD) was synthesized from 6-O-α-maltosyl-β-cyclodextrin (G2-β-CD) via a transglycosylation reaction catalyzed by TreX, a debranching enzyme from Sulfolobus solfataricus P2. TreX showed no activity toward glucosyl-β-CD, but a transfer product (1) was detected when the enzyme was incubated with maltosyl-β-CD, indicating specificity for a branched glucosyl chain bigger than DP2. Analysis of the structure of the transfer product (1) using MALDI-TOF/MS and isoamylase or glucoamylase treatment revealed it to be dimaltosyl-β-CD, suggesting that TreX transferred the maltosyl residue of a G2-β-CD to another molecule of G2-β-CD by forming an α-1,6-glucosidic linkage. When [14C]-maltose and maltosyl-β-CD were reacted with the enzyme, the radiogram showed no labeled dimaltosyl-β-CD; no condensation product between the two substrates was detected, indicating that the synthesis of dimaltosyl-β-CD occurred exclusively via transglycosylation of an α-1,6-glucosidic linkage. Based on the HPLC elution profile, the transfer product (1) was identified to be isomers of 61,63- and 61,64-dimaltosyl-β-CD. Inhibition studies with β-CD on the transglycosylation activity revealed that β-CD was a mixed-type inhibitor, with a Ki value of 55.6 μmol/mL. Thus, dimaltosyl-β-CD can be more efficiently synthesized by a transglycosylation reaction with TreX in the absence of β-CD. Our findings suggest that the high yield of (G2)2-β-CD from G2-β-CD was based on both the transglycosylation action mode and elimination of the inhibitory effect of β-CD.  相似文献   

14.
The recombinant catalytic α-subunit of N-glycan processing glucosidase II from Schizosaccharomyces pombe (SpGIIα) was produced in Escherichia coli. The recombinant SpGIIα exhibited quite low stability, with a reduction in activity to <40% after 2-days preservation at 4 °C, but the presence of 10% (v/v) glycerol prevented this loss of activity. SpGIIα, a member of the glycoside hydrolase family 31 (GH31), displayed the typical substrate specificity of GH31 α-glucosidases. The enzyme hydrolyzed not only α-(1→3)- but also α-(1→2)-, α-(1→4)-, and α-(1→6)-glucosidic linkages, and p-nitrophenyl α-glucoside. SpGIIα displayed most catalytic properties of glucosidase II. Hydrolytic activity of the terminal α-glucosidic residue of Glc2Man3-Dansyl was faster than that of Glc1Man3-Dansyl. This catalytic α-subunit also removed terminal glucose residues from native N-glycans (Glc2Man9GlcNAc2 and Glc1Man9GlcNAc2) although the activity was low.  相似文献   

15.
Proteins belonging to the glycoside hydrolase family 63 (GH63) are found in bacteria, archaea, and eukaryotes. Eukaryotic GH63 proteins are processing α-glucosidase I enzymes that hydrolyze an oligosaccharide precursor of eukaryotic N-linked glycoproteins. In contrast, the functions of the bacterial and archaeal GH63 proteins are unclear. Here we determined the crystal structure of a bacterial GH63 enzyme, Escherichia coli K12 YgjK, at 1.78 Å resolution and investigated some properties of the enzyme. YgjK consists of the N-domain and the A-domain, joined by a linker region. The N-domain is composed of 18 antiparallel β-strands and is classified as a super-β-sandwich. The A-domain contains 16 α-helices, 12 of which form an (α/α)6-barrel; the remaining 4 α-helices are found in an extra structural unit that we designated as the A′-region. YgjK, a member of the glycoside hydrolase clan GH-G, shares structural similarity with glucoamylase (GH15) and chitobiose phosphorylase (GH65), both of which belong to clan GH-L. In crystal structures of YgjK in complex with glucose, mannose, and galactose, all of the glucose, mannose, and galactose units were located in the catalytic cleft. YgjK showed the highest activity for the α-1,3-glucosidic linkage of nigerose, but also hydrolyzed trehalose, kojibiose, and maltooligosaccharides from maltose to maltoheptaose, although the activities were low. These findings suggest that YgjK is a glucosidase with relaxed specificity for sugars.  相似文献   

16.
We estimated the subsite structure of α-amylase II (TVA II) from Thermoactinomyces vulgaris R-47 expressed in Escherichia coli. TVA II has eight subsites, and the catalytic site is between the 5th and 6th subsite from the non-reducing end side. The subsite affinities, A-5, A-4, A-3, A-2, (A-1+A+1), A+2, and A+3, were calculated to be -0.35, 0.93, 0.55, 2.56, 1.18, 1.71, and 0.01 kcal mol-1, respectively.  相似文献   

17.
A glucodextranase (iGDase) from Arthrobacter globiformis I42 hydrolyzes alpha-1,6-glucosidic linkages of dextran from the non-reducing end to produce beta-D-glucose via an inverting reaction mechanism and classified into the glycoside hydrolase family 15 (GH15). Here we cloned the iGDase gene and determined the crystal structures of iGDase of the unliganded form and the complex with acarbose at 2.42-A resolution. The structure of iGDase is composed of four domains N, A, B, and C. Domain A forms an (alpha/alpha)(6)-barrel structure and domain N consists of 17 antiparallel beta-strands, and both domains are conserved in bacterial glucoamylases (GAs) and appear to be mainly concerned with catalytic activity. The structure of iGDase complexed with acarbose revealed that the positions and orientations of the residues at subsites -1 and +1 are nearly identical between iGDase and GA; however, the residues corresponding to subsite 3, which form the entrance of the substrate binding pocket, and the position of the open space and constriction of iGDase are different from those of GAs. On the other hand, domains B and C are not found in the bacterial GAs. The primary structure of domain C is homologous with a surface layer homology domain of pullulanases, and the three-dimensional structure of domain C resembles the carbohydrate-binding domain of some glycohydrolases.  相似文献   

18.
Bacteria that reside in the mammalian intestinal tract efficiently hydrolyze dietary carbohydrates, including starch, that escape digestion in the small intestine. Starch is an abundant dietary carbohydrate comprised of α1,4 and α1,6 linked glucose, yet mammalian intestinal glucoamylases cannot effectively hydrolyze starch that has frequent α1,6 branching as these structures hinder recognition and processing by α1,4‐specific amylases. Here we present the structure of the cell surface amylase SusG from Bacteroides thetaiotaomicron complexed with a mixed linkage amylosaccharide generated from transglycosylation during crystallization. Although SusG is specific for α1,4 glucosidic bonds, binding of this new oligosaccharide at the active site demonstrates that SusG can accommodate α1,6 branch points at subsite ?3 to ?2, and also at subsite+1 adjacent to the site of hydrolysis, explaining how this enzyme may be able to process a wide range of limit dextrins in the intestinal environment. These data suggest that B. thetaiotaomicron and related organisms may have a selective advantage for amylosaccharide scavenging in the gut.  相似文献   

19.
The structure of cauloside D, one of the main saponins isolated from Caulophyllum robustum roots, was shown to be 3-O-α-l-arabinopyranosyl hederagenin-28-O-α-l-rhamnopyranosyl-(1→4)-β-d-glucopyranosyl(1→6)-β-d-glucopyranoside with the aid of methylation and enzymatic hydrolysis by the digestive juice of the Eulota maackii. Cauloside A was shown to be identical with saponin A, isolated from C. robustum Maxim. previously. The composition of the digestive juice of E. maakii was shown to include enzymes that catalyse the cleavage of α-arabinosidic, β-1,6-glucosidic and acyl-O-β-glucosidic linkages.  相似文献   

20.
The gene for a membrane-bound, halophilic, and thermostable α-amylase, AmyB, from Halothermothrix orenii was cloned and sequenced. The crystal structure shows that, in addition to the typical domain organization of family 13 glycoside hydrolases, AmyB carries an additional N-terminal domain (N domain) that forms a large groove—the N-C groove—some 30 Å away from the active site. The structure of AmyB with the inhibitor acarbose at 1.35 Å resolution shows that a nonasaccharide has been synthesized through successive transglycosylation reactions of acarbose. Unexpectedly, in a complex of wild-type AmyB with α-cyclodextrin and maltoheptaose at 2.2 Å resolution, a maltotetraose molecule is bound in subsites − 1 to + 3, spanning the cleavage point at − 1/+ 1, with the − 1 glucosyl residue present as a 2So skew boat. This wild-type AmyB complex was obtained in the presence of a large excess of substrate, a condition under which it is possible to capture Michaelis complexes, which may explain the observed binding across − 1/+ 1 and ring distortion. We observe three methionine side chains that serve as “binding platforms” for glucosyl rings in AmyB, a seemingly rare occurrence in carbohydrate-binding proteins. The structures and results from the biochemical characterization of AmyB and AmyB lacking the N domain show that the N domain increases binding of the enzyme to raw starch. Furthermore, theoretical modeling suggests that the N-C groove can accommodate, spatially and chemically, large substrates such as A-starch.  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