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 406 毫秒
1.
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   

2.
Leuconostoc mesenteroides NRRL B-1426 dextransucrase synthesized a high molecular mass dextran (>2 × 106 Da) with ~85.5% α-(1→6) linear and ~14.5% α-(1→3) branched linkages. This high molecular mass dextran containing branched α-(1→3) linkages can be readily hydrolyzed for the production of enzyme-resistant isomalto-oligosaccharides. The acceptor specificity of dextransucrase for the transglycosylation reaction was studied using sixteen different acceptors. Among the sixteen acceptors used, isomaltose was found to be the best, having 89% efficiency followed by gentiobiose (64%), glucose (30%), cellobiose (25%), lactose (22.5%), melibiose (17%), and trehalose (2.3%) with reference to maltose, a known best acceptor. The β-linked disaccharide, gentiobiose, showed significant efficiency for oligosaccharide production that can be used as a potential prebiotic.  相似文献   

3.
Wheat bran offers health benefits as a baking ingredient, but is detrimental to bread textural quality. Dextran production by microbial fermentation improves sourdough bread volume and freshness, but extensive acid production during fermentation may negate this effect. Enzymatic production of dextran in wheat bran was tested to determine if dextran-containing bran could be used in baking without disrupting bread texture. The Weissella confusa VTT E-90392 dextransucrase gene was sequenced and His-tagged dextransucrase Wc392-rDSR was produced in Lactococcus lactis. Purified enzyme was characterized using 14C-sucrose radioisotope and reducing value-based assays, the former yielding K m and V max values of 14.7 mM and 8.2 μmol/(mg∙min), respectively, at the pH optimum of 5.4. The structure and size of in vitro dextran product was similar to dextran produced in vivo. Dextran (8.1% dry weight) was produced in wheat bran in 6 h using Wc392-rDSR. Bran with and without dextran was used in wheat baking at 20% supplementation level. Dextran presence improved bread softness and neutralized bran-induced volume loss, clearly demonstrating the potential of using dextransucrases in bran bioprocessing for use in baking.  相似文献   

4.
Exploration of biodiversity lead towards the discovery of novel exopolysaccharide (EPS) producing microbes that have multiple applications. The safety compatibility status of lactic acid bacteria (LAB) makes it an attractive candidate for the production of EPS in industries. Therefore, new bacterial isolates are continuously being identified from different habitats. Current research was conducted to explore indigenous biodiversity for the production of dextransucrase, which is involved in the synthesis of dextran. Dextran is an EPS which is used in different industries. In this study, thirty-nine LAB were isolated from different food samples. The isolates were identified as genus Leuconostoc, Weissella and Streptococcus based on genotypic and phenotypic characteristics. Screening revealed that only eight isolates can produce dextransucrase in high titres. Fermentation conditions of dextran producing LAB was optimized. The results indicated that Weissella confusa exhibited maximum specific activity (1.50?DSU?mg?1) in 8?h at 25?°C with pH 7.5. Dextran produced from Weissella proved to be a useful alternative to commercially used dextran produced by Leuconostoc mesenteroides in industries for various applications.  相似文献   

5.
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   

6.
The dextransucrase gene from Weissella confusa Cab3, having an open reading frame of 4.2?kb coding for 1,402?amino acids, was amplified, cloned, and expressed in Lactococcus lactis. The recombinant dextransucrase, WcCab3-rDSR was expressed as extracellular enzyme in M17 medium with a specific activity of 1.5?U/mg which after purification by PEG-400 fractionation gave 6.1?U/mg resulting in 4-fold purification. WcCab3-rDSR was expressed as soluble and homogeneous protein of molecular mass, approximately, 180?kDa as analyzed by SDS-PAGE. It displayed maximum enzyme activity at 35°C at pH 5.0 in 50?mM sodium acetate buffer. WcCab3-rDSR gave Km of 6.2?mM and Vm of 6.3?µmol/min/mg. The characterization of dextran synthesized by WcCab3-rDSR by Fourier transform infrared and nuclear magnetic resonance spectroscopic analyses revealed the structural similarities with the dextran produced by the native dextransucrase. The modeled structure of WcCab3-rDSR using the crystal structures of dextransucrase from Lactobacillus reuteri (protein data bank, PDB id: 3HZ3) and Streptococcus mutans (PDB id: 3AIB) as templates depicted the presence of different domains such as A, B, C, IV, and V. The domains A and B are circularly permuted in nature having (β/α)8 triose phosphate isomerase-barrel fold making the catalytic core of WcCab3-rDSR. The structure superposition and multiple sequence alignment analyses of WcCab3-rDSR with available structures of enzymes from family 70 GH suggested that the amino acid residue Asp510 acts as a nucleophile, Glu548 acts as a catalytic acid/base, whereas Asp621 acts as a transition-state stabilizer and these residues are found to be conserved within the family.  相似文献   

7.
The microbial community of artisanal corn fermentation called Chicha were isolated, purified and then identified using protein profile by Matrix Assisted Laser Desorption Ionization Time-of-Flight Mass Spectrometry (MALDI-TOF) and confirmed by partial ribosomal gene sequencing. Samples from Chicha beverage were chemically characterized by gas and liquid chromatography (HPLC and GC-MS). Aerobic mesophilic bacteria (AMB) (35.8% of total of isolated microorganisms), lactic acid bacteria (LAB) (21.6%) and yeast (42.6%) were identified. Species of the genera Klebsiella, Bacillus, Staphylococcus, Micrococcus, Enterobacter, and Weissella were identified. Rhodotorula mucilaginosa, Lodderomyces elongisporus, Candida metapsilosis, and C. bohicensis were the yeasts found. The LAB isolates detected were responsible for the high concentrations of lactic acid found during the fermentation process (1.2 g L??1), which is directly related to the decrease in pH values (from 6.95 to 3.70). Maltose was the main carbohydrate detected during corn fermentation (7.02 g L??1 with 36 h of fermentation). Ethanol was found in low concentrations (average 0.181 g L??1), making it possible to characterize the beverage as non-alcoholic. Twelve volatile compounds were identified by gas chromatography; belonging to the groups acids, alcohols aldehydes, acetate and others. MALDI-TOF was successfully used for identification of microbiota. Weissella confusa and W. cibaria were detected in the final product (after 36 h of fermentation), W. confusa is often classified as probiotic and deserve further application studies.  相似文献   

8.
The open reading frame of dsrE563, a dextransucrase gene obtained from a constitutive mutant (CB4-BF563) of Leuconostoc mesenteroides B-1299, consists of 8,511 bp encoding 2,836 amino acid residues. DsrE563 contains two catalytic domains (CD1 and CD2). Two truncated derivative mutants DsrE563ΔCD2ΔGBD (DsrE563-1) and DsrE563ΔCD2ΔVR (DsrE563-2) of DsrE563 were constructed and expressed using the pRSETC vector in Escherichia coli. The derivatives DsrE563-1 (deletion of 1,620 amino acids from the C-terminus) and DsrE563-2 (deletion of 1,258 amino acids from the C-terminus and 349 amino acids from the N-terminus) were expressed as active enzymes. Both enzymes synthesized less-soluble dextran, mainly containing α-1,6 glucosidic linkage. The synthesized less-soluble dextran also had a branched α-1,3 linkage. DsrE563-2 showed 4.5-fold higher dextransucrase activity than that of DsrE563-1 and showed higher acceptor reaction efficiency than that of dextransucrase from L. mesenteroides 512 FMCM when various mono or disaccharides were used as acceptors. Thus, the glucan-binding domain was important for both enzyme expression and dextransucrase activity.  相似文献   

9.
A water-soluble dextran was produced by purified dextransucrase from Leuconostoc mesenteroides NRRL B-640. The dextran was purified by alcohol precipitation. The structure of dextran was determined by FT-IR, 1H NMR, 13C NMR and 2-dimensional NMR spectroscopic techniques. NMR techniques (1D 1H, 13C and 2D HMQC) were used to fully assign the 1H and 13C spectra. All the spectral data showed that the dextran contains d-glucose residues in a linear chain with consecutive α(1  6) linkages. No branching was observed in the dextran structure. The viscosity of dextran solution decreased with the increase in shear rate exhibiting a typical non-Newtonian pseudoplastic behavior. The surface morphology of dried and powdered dextran studied using Scanning electron microscopy revealed the cubical porous structure.  相似文献   

10.
Pinene is a monoterpenes (C10) that is produced in a genetically-engineered microbial host for its industrial applications in fragrances, flavoring agents, pharmaceuticals, and biofuels. Herein, we have metabolically-engineered Corynebacterium glutamicum, to produce pinene and studied its toxicity in C. glutamicum. Geranyl diphosphate synthases (GPPS) and pinene synthases (PS), obtained from Pinus taeda and Abies grandis, were co-expressed with over-expressed native 1-deoxy-d-xylulose-5-phosphate synthase (Dxs) and isopentenyl diphosphate isomerase (Idi) from C. glutamicum using CoryneBrick vector. Most strains expressing PS-GPPSs produced detectable amounts of pinene, but co-expression of DXS and IDI with PS (P. taeda) and GPPS (A. grandis) resulted in 27 μg ± 7 α-pinene g?1 cell dry weight, which is the first report in C. glutamicum. Further engineering of PS and GPPS in the C. glutamicum strain may increase pinene production.  相似文献   

11.
The hydrolytic activity of a recombinant β-glycosidase from Dictyoglomus turgidum that specifically hydrolyzed the xylose at the C-6 position and the glucose in protopanaxatriol (PPT)-type ginsenosides followed the order Rf > Rg1 > Re > R1 > Rh1 > R2. The production of aglycone protopanaxatriol (APPT) from ginsenoside Rf was optimal at pH 6.0, 80 °C, 1 mg ml?1 Rf, and 10.6 U ml?1 enzyme. Under these conditions, D. turgidum β-glycosidase converted ginsenoside R1 to APPT with a molar conversion yield of 75.6 % and a productivity of 15 mg l?1 h?1 after 24 h by the transformation pathway of R1 → R2 → Rh1 → APPT, whereas the complete conversion of ginsenosides Rf and Rg1 to APPT was achieved with a productivity of 1,515 mg l?1 h?1 after 6.6 h by the pathways of Rf → Rh1 → APPT and Rg1 → Rh1 → APPT, respectively. In addition, D. turgidum β-glycosidase produced 0.54 mg ml?1 APPT from 2.29 mg ml?1 PPT-type ginsenosides of Panax ginseng root extract after 24 h, with a molar conversion yield of 43.2 % and a productivity of 23 mg l?1 h?1, and 0.62 mg ml?1 APPT from 1.35 mg ml?1 PPT-type ginsenosides of Panax notoginseng root extract after 20 h, with a molar conversion yield of 81.2 % and a productivity of 31 mg l?1 h?1. This is the first report on the APPT production from ginseng root extract. Moreover, the concentrations, yields, and productivities of APPT achieved in the present study are the highest reported to date.  相似文献   

12.
A Gram-negative, rod-shaped, slightly halophilic and facultatively anaerobic bacterium, designated strain D15-8WT, was isolated from the sediment of the South China Sea. Growth was found to occur optimally at 25 °C, between pH 7.0 and 8.0 and with 1–5 % (w/v) NaCl. The strain was observed to utilize a variety of organic substrates and polycyclic aromatic hydrocarbons as sole carbon sources. The G+C content of the genomic DNA was determined to be 58.7 %. The predominant respiratory quinone was found to be Q-9. The significant fatty acids were determined to be C16:0, C16:1 ω9c, C18:1 ω9c, C12:0 and C14:0 3OH. Analysis of 16S rRNA gene sequences showed that strain D15-8WT fits within the phylogenetic cluster of the genus Marinobacter and is most closely related to Marinobacter segnicrescens CGMCC 1.6489T, Marinobacter bryozoorum DSM 15401T, Marinobacter lacisalsi CECT 7297T and Marinobacter daqiaonensis CGMCC1.9167T. The DNA–DNA hybridization values between strain D15-8WT and the type strains of the most closely related species were 42.3 % (CGMCC 1.6489T), 39.8 % (DSM 15401T), 37.3 % (CECT 7297T) and 35.2 % (CGMCC1.9167T). The results of this polyphasic study indicate that strain D15-8WT represents a novel species of the genus Marinobacter, for which the name Marinobacter nanhaiticus sp. nov. is proposed. The type strain is D15-8WT (=CGMCC 1.11019T=KCTC 23749T).  相似文献   

13.
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   

14.
A novel bacterial strain, designated WS-MY5T, capable of degrading a variety of polysaccharides was isolated from a brown algae (Undaria pinnatifida) reservoir at Wando in the South Sea, South Korea. Strain WS-MY5T was found to grow optimally at 30 °C, at pH 7.0–7.5 and in the presence of 2 % (w/v) NaCl. A neighbour-joining phylogenetic tree based on 16S rRNA gene sequences showed that strain WS-MY5T falls within the clade comprising Winogradskyella species, clustering with the type strains of Winogradskyella pacifica, Winogradskyella arenosi, Winogradskyella rapida and Winogradskyella thalassocola, with which it exhibited 16S rRNA gene sequence similarity values of 97.3–98.8 %. It exhibited sequence similarity values of 93.0–96.2 % to the type strains of the other recognized Winogradskyella species. Strain WS-MY5T was found to contain MK-6 as the predominant menaquinone and anteiso-C15:0, iso-C15:0, iso-C15:0 3-OH, iso-C17:0 3-OH and iso-C15:1 G as the major fatty acids. The major polar lipids of strain WS-MY5T were identified as phosphatidylethanolamine, two unidentified lipids and two unidentified aminolipids. The DNA G+C content of strain WS-MY5T was determined to be 33.2 mol% and its DNA–DNA relatedness values with the type strains of W. pacifica, W. arenosi, W. rapida and W. thalassocola were in the range 16–28 %. Differential phenotypic properties, together with its phylogenetic and genetic distinctiveness, enabled strain WS-MY5T to be differentiated from the recognized Winogradskyella species. On the basis of the data presented here, strain WS-MY5T is considered to represent a novel species of the genus Winogradskyella, for which the name Winogradskyella undariae sp. nov. is proposed. The type strain is WS-MY5T (=KCTC 32261T=CCUG 63832T).  相似文献   

15.
Weissella confusa VTT E-90392 is an efficient producer of a dextran that is mainly composed of α-(1→6)-linked D-glucosyl units and very few α-(1→3) branch linkages. A mixture of the Chaetomium erraticum endodextranase and the Aspergillus niger α-glucosidase was used to hydrolyze W. confusa dextran to glucose and a set of enzyme-resistant isomaltooligosaccharides. Two of the oligosaccharides (tetra- and hexasaccharide) were isolated in pure form and their structures elucidated. The tetrasaccharide had a nonreducing end terminal α-(1→3)-linked glucosyl unit (α-D-Glcp-(1→3)-α-D-Glcp-(1→6)-α-D-Glcp-(1→6)-α-D-Glc), whereas the hexasaccharide had an α-(1→3)-linked isomaltosyl side group (α-D-Glcp-(1→6)[α-D-Glcp-(1→6)-α-D-Glcp-(1→3)]-α-D-Glcp-(1→6)-α-D-Glcp-(1→6)-α-D-Glc). A mixture of two isomeric oligosaccharides was also obtained in the pentasaccharide fraction, which were identified as (α-D-Glcp-(1→6)-α-D-Glcp-(1→3)-α-D-Glcp-(1→6)-α-D-Glcp-(1→6)-α-D-Glc) and (α-D-Glcp-(1→6)[α-D-Glcp-(1→3)]-α-D-Glcp-(1→6)-α-D-Glcp-(1→6)-α-D-Glc). The structures of the oligosaccharides indicated that W. confusa dextran contains both terminal and elongated α-(1→3)-branches. This is the first report evidencing the presence of elongated branches in W. confusa dextran. The (1)H and (13)C NMR spectroscopic data on the enzyme-resistant isomaltooligosaccharides with α-(1→3)-linked glucosyl and isomaltosyl groups are published here for the first time.  相似文献   

16.
A novel gene (amyZ) encoding a cold-active and salt-tolerant α-amylase (AmyZ) was cloned from marine bacterium Zunongwangia profunda (MCCC 1A01486) and the protein was expressed in Escherichia coli. The gene has a length of 1785 bp and encodes an α-amylase of 594 amino acids with an estimated molecular mass of 66 kDa by SDS-PAGE. The enzyme belongs to glycoside hydrolase family 13 and shows the highest identity (25 %) to the characterized α-amylase TVA II from thermoactinomyces vulgaris R-47. The recombinant α-amylase showed the maximum activity at 35 °C and pH 7.0, and retained about 39 % activity at 0 °C. AmyZ displayed extreme salt tolerance, with the highest activity at 1.5 M NaCl and 93 % activity even at 4 M NaCl. The catalytic efficiency (k cat/K m) of AmyZ increased from 115.51 (with 0 M NaCl) to 143.30 ml mg?1 s?1 (with 1.5 M NaCl) at 35 °C and pH 7.0, using soluble starch as substrate. Besides, the thermostability of the enzyme was significantly improved in the presence of 1.5 M NaCl or 1 mM CaCl2. AmyZ is one of the very few α-amylases that tolerate both high salinity and low temperatures, making it a potential candidate for research in basic and applied biology.  相似文献   

17.
A novel cold-adapted lipolytic enzyme gene, est97, was identified from a high Arctic intertidal zone sediment metagenomic library. The deduced amino acid sequence of Est97 showed low similarity with other lipolytic enzymes, the maximum being 30 % identity with a putative lipase from Vibrio caribbenthicus. Common features of lipolytic enzymes, such as the GXSXG sequence motif, were detected. The gene product was over-expressed in Escherichia coli and purified. The recombinant Est97 (rEst97) hydrolysed various ρ-nitrophenyl esters with the best substrate being ρ-nitrophenyl hexanoate (K m and k cat of 39 μM and 25.8 s?1, respectively). This esterase activity of rEst97 was optimal at 35 °C and pH 7.5 and the enzyme was unstable at temperatures above 25 °C. The apparent melting temperature, as determined by differential scanning calorimetry was 39 °C, substantiating Est97 as a cold-adapted esterase. The crystal structure of rEst97 was determined by the single wavelength anomalous dispersion method to 1.6 Å resolution. The protein was found to have a typical α/β-hydrolase fold with Ser144-His226-Asp197 as the catalytic triad. A suggested, relatively short lid domain of rEst97 is composed of residues 80–114, which form an α-helix and a disordered loop. The cold adaptation features seem primarily related to a high number of methionine and glycine residues and flexible loops in the high-resolution structures.  相似文献   

18.
Thirty samples of Italian durum wheat semolina and whole durum wheat semolina, generally used for the production of Southern Italy's traditional breads, were subjected to microbiological analysis in order to explore their lactic acid bacteria (LAB) diversity and to find strains with antifungal activity. A total of 125 presumptive LAB isolates (Gram-positive and catalase-negative) were characterized by repetitive extragenic palindromic-PCR (REP-PCR) and sequence analysis of the 16S rRNA gene, leading to the identification of the following species: Weissella confusa, Weissella cibaria, Leuconostoc citreum, Leuconostoc mesenteroides, Lactococcus lactis, Lactobacillus rossiae and Lactobacillus plantarum. The REP-PCR results delineated 17 different patterns whose cluster analysis clearly differentiated W. cibaria from W. confusa isolates. Seventeen strains, each characterized by a different REP-PCR pattern, were screened for their antifungal properties. They were grown in a flour-based medium, comparable to a real food system, and the resulting fermentation products (FPs) were tested against fungal species generally contaminating bakery products, Aspergillus niger, Penicillium roqueforti and Endomyces fibuliger. The results of the study indicated a strong inhibitory activity – comparable to that obtained with the common preservative calcium propionate (0.3% w/v) – of ten LAB strains against the most widespread contaminant of bakery products, P. roqueforti. The screening also highlighted the unexplored antifungal activity of L. citreum, L. rossiae and W. cibaria (1 strain), which inhibited all fungal strains to the same or a higher extent compared with calcium propionate. The fermentation products of these three strains were characterized by low pH values, and a high content of lactic and acetic acids.  相似文献   

19.
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   

20.
The regioselectivity of alternansucrase (EC 2.4.1.140) differs from dextransucrase (EC 2.4.1.5) in ways that can be useful for the synthesis of novel oligosaccharide structures. For example, it has been recently shown that the major oligosaccharides produced when maltose is the acceptor include one trisaccharide structure, two tetrasaccharides, one pentasaccharide, two hexasaccharides, one heptasaccharide, and at least two octasaccharides, containing no adjacent α-(1→3) linkages and no more than two consecutive α-(1→6) linkages. This may shed some light on how the enzyme works to produce the alternating structure. Another characteristic of alternansucrase that distinguishes it from dextransucrase is its greater ability to use leucrose as an acceptor. Leucrose, produced by glucosyl transfer to fructose released from the initial sucrose substrate, represents a very poor substrate for Leuconostoc mesenteroides NRRL B-512F dextransucrase. Alternansucrase, however, continues to transfer glucosyl units to leucrose, resulting in some unusual glucosyl-fructose oligosaccharides.  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