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.
Abstract

Sucrose phosphorylase is a bacterial transglucosidase that catalyzes conversion of sucrose and phosphate into α-D-glucose-1-phosphate and D-fructose. The enzyme utilizes a glycoside hydrolase-like double displacement mechanism that involves a catalytically competent β-glucosyl enzyme intermediate. In addition to reaction with phosphate, glucosylated sucrose phosphorylase can undergo hydrolysis to yield α-D-glucose or it can decompose via glucosyl transfer to a hydroxy group in suitable acceptor molecules, giving new α-D-glucosidic products. The glucosyl acceptor specificity of sucrose phosphorylase is reviewed, focusing on applications of the enzyme in glucoside synthesis. Polyhydroxylated compounds such as sugars and sugar alcohols are often glucosylated efficiently. Aryl alcohols and different carboxylic acids also serve as acceptors for enzymatic transglucosylation. The natural osmolyte 2-O-(α-D-glucopyranosyl)-sn-glycerol (GG) was prepared by regioselective glucosylation of glycerol from sucrose using the phosphorylase from Leuconostoc mesenteroides. An industrial process for production of GG as active ingredient of cosmetic formulations has been recently developed. General advantages of sucrose phosphorylase as a transglucosylation catalyst lie in the use of sucrose as a high-energy glucosyl donor and the usually weak hydrolase activity of the enzyme towards substrate and product.  相似文献   

2.
Six α-monoglucosyl derivatives of ginsenoside Rg1 (G-Rg1) were synthesized by transglycosylation reaction of rice seed α-glucosidase in the reaction mixture containing maltose as a glucosyl donor and G-Rg1 as an acceptor. Their chemical structures were identified by spectroscopic analysis, and the effects of reaction time, pH, and glycosyl donors on transglycosylation reaction were investigated. The results showed that rice seed α-glucosidase transfers α-glucosyl group from maltose to G-Rg1 by forming either α-1,3 (α-nigerosyl)-, α-1,4 (α-maltosyl)-, or α-1,6 (α-isomaltosyl)-glucosidic linkages in β-glucose moieties linked at the C6- and C20-position of protopanaxatriol (PPT)-type aglycone. The optimum pH range for the transglycosylation reaction was between 5.0 and 6.0. Rice seed α-glucosidase acted on maltose, soluble starch, and PNP α-D-glucopyranoside as glycosyl donors, but not on glucose, sucrose, or trehalose. These α-monoglucosyl derivatives of G-Rg1 were easily hydrolyzed to G-Rg1 by rat small intestinal and liver α-glucosidase in vitro.  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.
Dextransucrase from Leuconostoc mesenteroides B-512 catalyzes the polymerization of dextran from sucrose. The resulting dextran has 95% α-1 → 6 linkages and 5% α-1 → 3 branch linkages. A purified dextransucrase was insolubilized on Bio-Gel P-2 beads (BGD, Bio-Gel-dextransucrase). The BGD was labeled by incubating it with a very low concentration of [14C]sucrose or it was first charged with nonlabeled sucrose and then labeled with a very low concentration of [14C]sucrose. After extensive washings with buffer, the 14C label remained attached to BGD. This labeled material was previously shown to be [14C]dextran and was postulated to be attached covalently at the reducing end to the active site of the enzyme. When the labeled BGD was incubated with a low molecular weight nonlabeled dextran (acceptor dextran) all of the BGD-bound label was released as [14C]dextran whereas essentially no [14C]dextran was released when the labeled BGD was incubated in buffer alone under comparable conditions. The released [14C]dextran was shown to be a slightly branched dextran by hydrolysis with an exodextranase. Acetolysis of the released dextran gave 7.3% of the radioactivity in nigerose. Reduction with sodium borohydride, followed by acid hydrolysis, gave all of the radioactivity in glucose, indicating that the nigerose was exclusively labeled in the nonreducing glucose unit. These results indicated that [14C]dextran was being released from BGD by virtue of the action of the low molecular weight dextran and that this action gave the formation of a new α-1 → 3 branch linkage. A mehanism for branching is proposed in which a C3-OH on an acceptor dextran acts as a nucleophile on C1 of the reducing end of a dextranosyl-dextransucrase complex, thereby displacing dextran from dextransucrase and forming an α-1 → 3 branch linkage. It is argued that the biosynthesis of branched linkages does not require a separate branching enzyme but can take place by reactions of an acceptor dextran with a dextranosyl-dextransucrase complex.  相似文献   

5.
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   

6.
Kinetic and molecular docking studies were performed to characterize the binding of α-d-glucose 1-phosphate (αGlc 1-P) at the catalytic subsite of a family GH-13 sucrose phosphorylase (from L. mesenteroides) in wild-type and mutated form. The best-fit binding mode of αGlc 1-P dianion had the phosphate group placed anti relative to the glucosyl moiety (adopting a relaxed 4C1 chair conformation) and was stabilized mainly by hydrogen bonds from residues of the enzyme?s catalytic triad (Asp196, Glu237 and Asp295) and from Arg137. Additional feature of the αGlc 1-P docking pose was an intramolecular hydrogen bond (2.7 Å) between the glucosyl C2-hydroxyl and the phosphate oxygen. An inactive phosphonate analog of αGlc 1-P did not show binding to sucrose phosphorylase in different experimental assays (saturation transfer difference NMR, steady-state reversible inhibition), consistent with evidence from molecular docking study that also suggested a completely different and strongly disfavored binding mode of the analog as compared to αGlc 1-P. Molecular docking results also support kinetic data in showing that mutation of Phe52, a key residue at the catalytic subsite involved in transition state stabilization, had little effect on the ground-state binding of αGlc 1-P by the phosphorylase. However, when combined with a second mutation involving one of the catalytic triad residues, the mutation of Phe52 by Ala caused complete (F52A_D196A; F52A_E237A) or very large (F52A_D295A) disruption of the proposed productive binding mode of αGlc 1-P with consequent effects on the enzyme activity. Effects of positioning of αGlc 1-P for efficient glucosyl transfer from phosphate to the catalytic nucleophile of the enzyme (Asp196) are suggested. High similarity between the αGlc 1-P conformers bound to sucrose phosphorylase (modeled) and the structurally and mechanistically unrelated maltodextrin phosphorylase (experimental) is revealed.  相似文献   

7.
Rice BGlu1 β-glucosidase nucleophile mutant E386G is a glycosynthase that can synthesize p-nitrophenyl (pNP)-cellooligosaccharides of up to 11 residues. The X-ray crystal structures of the E386G glycosynthase with and without α-glucosyl fluoride were solved and the α-glucosyl fluoride complex was found to contain an ordered water molecule near the position of the nucleophile of the BGlu1 native structure, which is likely to stabilize the departing fluoride. The structures of E386G glycosynthase in complexes with cellotetraose and cellopentaose confirmed that the side chains of N245, S334, and Y341 interact with glucosyl residues in cellooligosaccharide binding subsites +2, +3, and +4. Mutants in which these residues were replaced in BGlu1 β-glucosidase hydrolyzed cellotetraose and cellopentaose with k(cat) /K(m) values similar to those of the wild type enzyme. However, the Y341A, Y341L, and N245V mutants of the E386G glycosynthase synthesize shorter pNP-cellooligosaccharides than do the E386G glycosynthase and its S334A mutant, suggesting that Y341 and N245 play important roles in the synthesis of long oligosaccharides. X-ray structural studies revealed that cellotetraose binds to the Y341A mutant of the glycosynthase in a very different, alternative mode not seen in complexes with the E386G glycosynthase, possibly explaining the similar hydrolysis, but poorer synthesis of longer oligosaccharides by Y341 mutants.  相似文献   

8.
Inverting cellobiose phosphorylase (CtCBP) and cellodextrin phosphorylase (CtCDP) from Clostridium thermocellum ATCC27405 of glycoside hydrolase family 94 catalysed reverse phosphorolysis to produce cellobiose and cellodextrins in 57% and 48% yield from α-d-glucose 1-phosphate as donor with glucose and cellobiose as acceptor, respectively. Use of α-d-glucosyl 1-fluoride as donor increased product yields to 98% for CtCBP and 68% for CtCDP. CtCBP showed broad acceptor specificity forming β-glucosyl disaccharides with β-(1→4)- regioselectivity from five monosaccharides as well as branched β-glucosyl trisaccharides with β-(1→4)-regioselectivity from three (1→6)-linked disaccharides. CtCDP showed strict β-(1→4)-regioselectivity and catalysed linear chain extension of the three β-linked glucosyl disaccharides, cellobiose, sophorose, and laminaribiose, whereas 12 tested monosaccharides were not acceptors. Structure analysis by NMR and ESI-MS confirmed two β-glucosyl oligosaccharide product series to represent novel compounds, i.e. β-d-glucopyranosyl-[(1→4)-β-d-glucopyranosyl]n-(1→2)-d-glucopyranose, and β-d-glucopyranosyl-[(1→4)-β-d-glucopyranosyl]n-(1→3)-d-glucopyranose (n = 1–7). Multiple sequence alignment together with a modelled CtCBP structure, obtained using the crystal structure of Cellvibrio gilvus CBP in complex with glucose as a template, indicated differences in the subsite +1 region that elicit the distinct acceptor specificities of CtCBP and CtCDP. Thus Glu636 of CtCBP recognized the C1 hydroxyl of β-glucose at subsite +1, while in CtCDP the presence of Ala800 conferred more space, which allowed accommodation of C1 substituted disaccharide acceptors at the corresponding subsites +1 and +2. Furthermore, CtCBP has a short Glu496-Thr500 loop that permitted the C6 hydroxyl of glucose at subsite +1 to be exposed to solvent, whereas the corresponding longer loop Thr637–Lys648 in CtCDP blocks binding of C6-linked disaccharides as acceptors at subsite +1. High yields in chemoenzymatic synthesis, a novel regioselectivity, and novel oligosaccharides including products of CtCDP catalysed oligosaccharide oligomerisation using α-d-glucosyl 1-fluoride, all together contribute to the formation of an excellent basis for rational engineering of CBP and CDP to produce desired oligosaccharides.  相似文献   

9.
Previous studies have indicated that α-d-1-fluoroglucose is a glycosyl donor for glucosyl transferases (5, 6) including dextransucrases formed by Leuconostoc and Streptococcus mutans. The present report confirms these observations with dextransucrase isolated from S. sanguis and conclusively establishes the details of this reaction as well as proving that mechanism of fluoroglucose transfer is comparable to that glucosyl transfer from sucrose. A new procedure for monitoring the reaction is reported, and is based on the measurement of proton formation using the pH indicator, bromcresol purple. Production of F? was found to be stoichiometric with proton production. Rate studies with the substrate indicate that α-1-fluoroglucose undergoes spontaneous hydrolysis, which is greatly increased in the presence of nucleophilic buffers. When [14C]maltose and α-1-fluoroglucose or [14C]α-1-fluoroglucose and maltose were incubated with dextransucrase, a series of oligosaccharide products was observed. The results indicate that the glucosyl moiety of α-1-fluoroglucose transferred to the acceptor. The nature of formation of the products are consistent with a series of precursor-product reactions. Product analysis of the saccharides by borohydride reduction analysis demonstrated that the glucosyl unit was added to the nonreducing end of maltose. When either [14C]fructose or [14C]-α-1-fluoroglucose were incubated with enzyme, a reaction was observed which was analogous to the isotopic-exchange reaction catalyzed by the enzyme in the presence of [14C]fructose and sucrose.  相似文献   

10.
In a study undertaken to illustrate the inadequacy of the familiar concept of carbohydrases as hydrolases, crystalline α-amylases from six different sources, as well as crude salivary amylase, were examined and found to catalyze the synthesis of maltose and maltosaccharides from α-d-glucopyranosyl fluoride, a stereoanalog of α-d-glucopyranose. These syntheses apparently involve initial formation of maltosyl fluoride and higher maltosaccharide 1-fluorides, traces of which were found in digests with certain α-amylases. That the reactions are due to the α-amylases themselves and not to some accompanying enzyme(s) appears certain from the purity and diversity of the preparations; their failure (with one exception) to attack α- or β-maltose; the correspondence of the synthesized products with the known specificity of α-amylases for α-1,4-d-glucosidic linkages (and capacity of different α-amylases to hydrolyze saccharides of different sizes). The “saccharifying” α-amylase of B. sublilis var amylosacchariticus was unique in producing maltosaccharides from both α- and β-maltose (i.e., by α-d-glucosyl transfer). However, the entire group of α-amylases had the capacity to promote α-d-glucosyl transfer from α-d-glucosyl fluoride to C4-carbinol sites, demonstrating for the first time that the catalytic range of α-amylase extends beyond hydrolysis and its reversal. Indeed, all transferred the glucosyl group of α-d-glycosyl fluoride preferentially to C4-carbinols rather than water—a finding neither anticipated nor explained by the representation of α-amylases as hydrolases.  相似文献   

11.
With all glucobioses (eleven types) as acceptors, Leuconostoc mesenteroides (NRRL B-512) was grown on a sucrose medium. The trisaccharides produced were analyzed for their yields, and the trisaccharide structures were determined after separation on a column.

The yield of the tri- and higher-saccharides indicated that isomaltose (28%) was the most efficient acceptor and maltose (24%) was next. With the other eight glucobioses, oligosaccharides were obtained in 3~15% yield. Among these sugars, α,β-trehalose (15%) and β,β-trehalose (11%) were efficient acceptors next to maltose, but α,α-trehalose was inert.

In every case, except for cellobiose, α-glucosyl transfer occurred to the position 6 of non-reducing moiety of glucobiose.

The sugars produced contained five new trisaccharides which were isolated as pure compounds.  相似文献   

12.
Abstract

Sucrose phosphorylase is a bacterial α-transglucosidase that catalyses glucosyl transfer from sucrose to phosphate, releasing d-fructose and α-d-glucose 1-phosphate as the product of the first (enzyme glucosylation) and second (enzyme deglucosylation) step of the enzymatic reaction, respectively. The transferred glucosyl moiety of sucrose is accommodated at the catalytic subsite of the phosphorylase through a network of charged hydrogen bonds whereby a highly conserved residue pair of Asp and Arg points towards the equatorial hydroxyl at C4. To examine the role of this ‘hyperpolar’ binding site for the substrate 4-OH, we have mutated Asp49 and Arg395 of Leuconostoc mesenteroides sucrose phosphorylase individually to Ala (D49A) and Leu (R395L), respectively, and also prepared an ‘uncharged’ double mutant harbouring both site-directed substitutions. The efficiency for enzyme glucosylation from sucrose was massively decreased in purified preparations of D49A (107-fold) and R395L (105-fold) as compared to wild-type enzyme. The double mutant was not active above the detection limit. Enzyme deglucosylation to phosphate proceeded relatively efficient in D49A as well as R395L, about 500-fold less than in the wild-type phosphorylase. Substrate inhibition by phosphate and a loss in selectivity for reaction with phosphate as compared to water were new features in the two mutants. Asp49 and Arg395 are both essential in the catalytic reaction of L. mesenteroides sucrose phosphorylase.  相似文献   

13.
The Aspergillus awamori K4 β-xylosidase gene (Xaw1) sequence was deduced by sequencing RT-PCR and PCR products. The ORF was 2,412 bp and the predicted peptide was 804 amino acids long, corresponding to a molecular weight of 87,156 Da. The mature protein was 778 amino acids long with a molecular weight of 84,632 Da. A homology search of the amino acid sequence revealed that it was very similar to the Aspergillus niger β-xylosidase gene with only five amino acid differences. K4 β-xylosidase had the same catalytic mechanism as family 3 β-glucosidases, involving Asp in region A. At an early stage in the reaction with xylobiose and xylotriose, the hydrolysis rate was much lower than the transxylosylation rate, decreasing gradually as the substrate concentration increased, whereas the transxylosylation rate increased greatly. Aspergillus awamori K4 β-xylosidase had broad acceptor specificity toward alcohols, hydroxybenzenealcohols, sugar alcohols and disaccharides. A consensus portion involving the hydroxymethyl group of the acceptor was confirmed in the major transfer products 1(4)-O-β-d-xylosyl erythritol, (2-hydroxyl)-phenyl-methyl-β-d-xylopyranoside, 6S-O-β-d-xylosyl maltitol (S: sorbitol residue) and 6G-O-β-d-xylosyl palatinose (G: glucosyl residue). This might suggest that the methylene in the hydroxymethyl group facilitates base-catalyzed hydroxyl group attack of the anomeric center of the xylosyl–enzyme intermediate.  相似文献   

14.
The hyperthermophilic endoglucanase Cel5A from Thermotoga maritima can find applications in lignocellulosic biofuel production, because it catalyzes the hydrolysis of glucan- and mannan-based polysaccharides. Here, we report the crystal structures in apo-form and in complex with three ligands, cellotetraose, cellobiose and mannotriose, at 1.29? to 2.40? resolution. The open carbohydrate-binding cavity which can accommodate oligosaccharide substrates with extensively branched chains explained the dual specificity of the enzyme. Combining our structural information and the previous kinetic data, it is suggested that this enzyme prefers β-glucosyl and β-mannosyl moieties at the reducing end and uses two conserved catalytic residues, E253 (nucleophile) and E136 (general acid/base), to hydrolyze the glycosidic bonds. Moreover, our results also suggest that the wide spectrum of Tm_Cel5A substrates might be due to the lack of steric hindrance around the C2-hydroxyl group of the glucose or mannose unit from active-site residues.  相似文献   

15.
The capacity of CGTase for using βCD (beta cyclodextrin) as glucosyl donor and transferring it to sucrose molecules was investigated. We showed that this enzyme was able to produce polyglucosyl-fructosides (GnF) by a coupling reaction between βCD and sucrose. Maltooligosaccharides were also synthesised but in lesser amounts. The degree of polymerisation (DP) of the different products was limited to a value of 8 and this allowed us to purify all of them by size exclusion chromatography. Mass spectrometry and NMR analysis of the unknown products revealed that they consisted of linear maltooligosaccharides of various DP bound to the glucose moiety of a sucrose molecule by a α(1→4) linkage.  相似文献   

16.
Transglucosyl-amylase was purified 96-fold and partially characterized. The Km value with dextrin as substrate was 9.1 mg/ml. Glycerol, an acceptor of d-glucose, appeared to inhibit dextrin hydrolysis noncompetitively. The energy of activation of the enzyme was 7,920 cal/mole. Indirect determinations showed that synthesis of d-glucosyl glycerol was significantly affected by the nature of the amylaceous substrate. Glucosyl-glycerol synthesis did not increase as incubation temperature was raised from 50 to 60 C. Direct determinations by gas-liquid chromatography indicated that the synthesis of glucosyl glycerol, as a function of the concentration of either enzyme, substrate, or glycerol, traced a curvilinear path approaching 15 mg/ml as the maximum. When enzyme, substrate, and glycerol at high concentrations were varied in all possible combinations, however, conditions for producing as much as 47.5 mg/ml of glucosyl glycerol were established.  相似文献   

17.
《Carbohydrate research》1986,146(1):39-42
Trehalase was previously shown (see ref. 5) to hydrolyze α-d-glucosyl fluoride, forming β-d-glucose, and to synthesize α,α-trehalose from β-d-glucosyl fluoride plus α-d-glucose. Present observations further define the enzyme's separate cosubstrate requirements in utilizing these nonglycosidic substrates. α-d-Glucopyranose and α-d-xylopyranose were found to be uniquely effective in enabling Trichoderma reesei trehalase to catalyze reactions with β-d-glucosyl fluoride. As little as 0.2mm added α-d-glucose (0.4mm α-d-xylose) substantially increased the rate of enzymically catalyzed release of fluoride from 25mm β-d-glucosyl fluoride at 0°. Digest of β-d-glucosyl fluoride plus α-d-xylose yielded the α,α-trehalose analog, α-d-glucopyranosyl α-d-xylopyranoside, as a transient (i.e., subsequently hydrolyzed) transfer-product. The need for an aldopyranose acceptor having an axial 1-OH group when β-d-glucosyl fluoride is the donor, and for water when α-d-glucosyl fluoride is the substrate, indicates that the catalytic groups of trehalose have the flexibility to catalyze different stereochemical reactions.  相似文献   

18.
Nigerose and nigerooligosaccharides served as acceptors for a glucosyltransferase GTF-I from cariogenic Streptococcus sobrinus to give a series of homologous acceptor products. The soluble oligosaccharides (dp 5-9) strongly activated the acceptor reaction, resulting in the accumulation of water-insoluble (1-->3)-alpha-D-glucan. The enzyme transferred the labeled glucosyl residue from D-[U-13C]sucrose to the 3-hydroxyl group at the non-reducing end of the (1-->3)-alpha-D-oligosaccharides, as unequivocally shown by NMR 13C-13C coupling patterns. The values of the 13C-13C one-bond coupling constant (1J) are also presented for the C-1-C-6 of the 13C-labeled alpha-(1-->3)-linked glucosyl residue and of the non-reducing-end residue.  相似文献   

19.
Dodecyl glucooligosides, a class of interesting non ionic surfactant molecules were synthesized by cyclodextrin glucanotransferase from Bacillus macerans using either α-cyclodextrin (α-CD) or soluble starch as glycosyl donor and dodecyl β-d-glucoside (C12G1) or dodecyl β-d-maltoside (C12G2) as acceptor substrates. The primary coupling products obtained in the respective reactions were identified as dodecyl glucoheptaoside and dodecyl maltooctaoside by mass spectrometry. Higher yields of coupling products were obtained using α-CD as donor, while more dispoportionation occurred with starch. Nearly 78% conversion of the acceptor substrate C12G1 into dodecyl glucooligosides could be achieved at 132 μg/ml of CGTase in 20 min, while 93% of C12G2 could be transformed into products at 17.6 μg/ml of enzyme in 120 min using soluble starch as donor substrate. For applications requiring pure compounds like C12G7, synthesis using α-CD is advantageous. However, for applications in which a mixture of elongated alkyl glycosides is needed, reactions employing starch are clearly competitive.  相似文献   

20.
Abstract

Glucansucrases (GTFs) catalyzes the synthesis of α-glucans from sucrose and oligosaccharides in the presence of an acceptor sugar by transferring glucosyl units to the acceptor molecule with different linkages. The acceptor reactions can be affected by several parameters and this study aimed to determine the optimal reaction parameters for the production of glucansucrase-based oligosaccharides using sucrose and maltose as the donor and acceptor sugars, respectively via a hybrid technique of Response Surface Method (RSM) and Particle Swarm Optimization (PSO). The experimental design was performed using Central Composite Design and the tested parameters were enzyme concentration, acceptor:donor ratio and the reaction period. The optimization studies showed that enzyme concentration was the most effective parameter for the final oligosaccharides yields. The optimal values of the significant parameters determined for enzyme concentration and acceptor:donor ratio were 3.45?U and 0.62, respectively. Even the response surface plots for input parameters verified the PSO results, an experimental validation study was performed for the reverification. The experimental verification results obtained were also consistent with the PSO results. These findings will help our understanding in the role of different parameters for the production of oligosaccharides in the acceptor reactions of GTFs.  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