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.
Two β-d-glucan endo-hydrolases purified from germinating barley (Hordeum vulgare) hydrolyse (1→4)-β linkages in (1→3),(1→4)-β-d-glucans where the d-glucosyl residue is substituted at O-3, but will not hydrolyse (1→3)-β-d-glucans or (1→4)-β-d-glucans. Methylation analysis of hydrolytic products released from barley (1→3),(1→4)-β-d-glucan indicates that 3-O-β-cellobiosyl-d-glucose and 3-O-β-cellotriosyl-d-glucose are the major oligomers formed. The enzymes exhibit characteristic endo-hydrolase action-patterns on this substrate. Both enzyme can therefore be classified as (1→3),(1→4)-β-d-glucan 4-glucanohydrolases (EC 3.2.1.73). The reduced, pneumococcal polysaccharide RS III, which consists of alternating (1→3)- and (1→4)-linked β-d-glucosyl residues, is hydrolysed by the enzymes to release laminaribiose as a major oligomeric product. Although the kinetic parameters of the two enzymes are similar, one hydrolyses barley (1→3),(1→4)-β-d-glucan at a significantly higher rate than the other and is more stable at elevated temperatures.  相似文献   

2.
Two types of extracellular (1→6)-β-D-glucanases are produced by Bacillus circulans WL-12, and these enzymes are differentiated by their ability to lyse yeast cell-walls. The non-lytic (1→6)-β-D-glucanase was isolated by a combination of Sephadex G-100, Bio-Gel P-100, and DEAE-Bio-Gel A chromatography. The purified enzyme was eloctrophoretically homogeneous and had a molecular weight of 52,000. For the substrate pustulan, the enzyme exhibited the following kinetic properties: pH, 5.0; Km, 0.83 mg of pustulan/ml; Vmax, 104 microequivalents of D-glucose released/min/mg of protein. Pustulan was hydrolysed by an endo-mechanism, producing D-glucose and gentiobiose as preponderant final products. The non-lytic enzyme was specific for the (1→6)-β-D-glucosidic linkage and did not hydrolyse branched, (1→3)-β-D-linked glucans containing (1→6)-interchain linkages. In contrast, the lytic (1→6)-β-D-glucanase produced D-glucose, gentiobiose, and gentiotriose as the final products of pustulan hydrolysis, and exhibited significant activity on branched (1→3)-β-D-glucans having (1→6)-interchain linkages. In these cases, the major products were gentiobiose and D-glucose, suggesting an ability of the lytic enzyme to cleave some (1→3)-linkages surrounding a (1→6)-branch-point. This latter property may explain the ability of this enzyme to weakly lyse yeast cell-walls.  相似文献   

3.
Extraction of the cell walls of Schizosaccharomyces pombe with dilute alkali at 4° yields a mixture of polysaccharides including galactomannan, (1→3)-α-d-glucan, and a branched (1→3)-β-d-glucan. The alkali-insoluble residue contains a lightly branched (1→3)-β-d-glucan, together with smaller amounts of an extremely highly branched (1→6)-β-d-glucan. The properties of the three distinct β-d-glucans are compared with those isolated from other yeasts.  相似文献   

4.
Barley endosperm cell-walls were prepared and analysed. The-carbohydrate portion, which constituted most of the wall material, consisted of 10 % of l-arabinose, 13% Of d-xylose, 74% of d-glucose and 2.5% of d-mannose. Mixed-linkage β-d-glucan represented 70-72% of this material; the remaining 2-4% Of d-glucose maybe present as cellulose and glucomannan. Water and alkali-extracted β-d-glucans contained similar ratios of (1 → 3)- to (1 → 4)-Linkages, namely 3 to 7. The walls, which had a protein content of approximately 5 %, contained unidentfied, alkali-labile linkages. An endo-(1 → 3)β-d-glucanase from malted barley, and a fungal endo-(1 → 4)-β-d-glucanase, caused extensive solubilization of the wall polysaccharides.  相似文献   

5.
The particulate enzyme responsible for the synthesis of β-(1→4)-d-glucans from UDP-[14C-d-glucose has been solubilized and some of its properties have been characterized. Mg2+ markedly enhanced synthesis of β-(1→4)-d-glucans and inhibited synthesis of β-(1→3)-d-glucans. The optimal pH for synthesis of β-(1→4)-d-glucans is near pH 8 and the synthesis was enhanced in these preparations by d-glucose, methyl-β-d-glucopyranoside and cellobiose.  相似文献   

6.
Water-soluble (1→3),(1→4)-β-d-glucans isolated from barleys grown in Australia and the UK were depolymerised using a purified (1→3),(1→4)-β-d-glucan 4-glucanohydrolase (EC 3.2.1.73). Oligomeric products were quantitatively separated by high resolution gel filtration chromatography and their structures defined by methylation analysis. Approximately 90% (w/w) of each polysaccharide consists of cellotriosyl and cellotetraosyl residues separated by single (1→3)-linkages but blocks of 5–11 (1→4)-linked glucosyl residues are also present in significant proportions. Periodate oxidation followed by Smith degradation suggested that contiguous (1→3)-linked β-glucosyl residues are either absent, or present in very low frequency. The potential for misinterpretation of data due to incomplete Smith degradation was noted.The irregularly-spaced (1→3)-linkages interrupt the relatively rigid, ribbon-like (1→4)-β-glucan conformation and confer a flexibility and ‘irregular’ shape on the barley (1→3),(1→4)-β-d-glucan, consistent with its solubility in water. Molecular models incorporating the major structural features confirm that the polysaccharide is likely to assume a worm-like conformation in solution. Non-covalent interactions between long blocks of (1→4)-linkages in (1→3),(1→4)-β-d-glucans, or between these blocks and other polysaccharides, offer a possible explanation for the organisation of polysaccharides in the framework of the cell wall.  相似文献   

7.
Antitumor activities of two (1 → 6)-branched (1 → 3)-β-d-glucans, isolated from the fruiting body of Auricularia auricula-judae (“kikurage”, an edible mushroom), and other branched polysaccharides containing a backbone chain of (1 → 3)-α-d-glucosidic or (1 → 3)-α-d-mannosidic linkages [and their corresponding (1 → 3)-d-glycans, derived by mild, Smith degradation] were compared. Among these polysaccharides, a water-soluble, branched (1 → 3)-β-d-glucan (glucan I) of A. auriculajudae exhibited potent, inhibitory activity against implanted Sarcoma 180 solid tumor in mice. The alkali-insoluble, branched (1 → 3)-β-d-glucan (glucan II), a major constituent of the fruiting body, showed essentially no inhibitory activity. When the latter glucan, having numerous branches attached, was modified by controlled, periodate oxidation, borohydride reduction, and mild, acid hydrolysis, the resulting, water-soluble, degraded glucan, having covalently linked polyhydroxy groups attached at O-6 of the (1 → 3)-linked d-glucosyl residues, exhibited potent antitumor activity. Further investigations using the glucan-polyalcohol indicated that the attachment of the polyhydroxy groups to the (1 → 3)-β-d-glucan backbone may enhance the antitumor potency of the glucan. On the other hand, partial introduction  相似文献   

8.
《Carbohydrate research》1986,147(1):69-85
The insoluble material that remains after extraction of Zea shoots with cold buffer was treated successively with 3m LiCl and hot water. The polysaccharides solubilized by these treatments were mostly (1→3),(1→4)-β-d-glucans. The β-d-glucan from the hot-water-soluble fraction was hydrolyzed by Bacillus subtilis (1→3),(1→4)-β-d-glucan 4-glucanohydrolase. The oligosaccharides were characterized by methylation analysis of the enzymic fragments and by methylation analysis of secondary fragments generated by treatment of the isolated oligosaccharides with Streptomyces QM B814 cellulase. The results demonstrate that the native polysaccharide consists mainly of cellotriosyl and cellotetraosyl residues joined by single (1→3) linkages. Evidence is presented to show that certain other glucosyl sequences are also present in the native polysaccharide including (a) two, three, or four contiguous (1→3)-linkages; (b) blocks of more than four (1→4)-linked glucose residues; (c) regions having alternating (1→3)- and (1→4)-linkages.  相似文献   

9.
《Carbohydrate research》1986,147(2):265-274
Syntheses, based on silver trifluoromethanesulfonate-promoted Koenigs-Knorr type condensations, are described of the d-glucotrioses, β-d-Glcp-(1→3)-β-d-Glcp-(1→4)-d-Glcp and β-d-Glcp-(1→4)-β-d-Glcp-(1→3)-d-Glcp, and the d-Glucotetraoses, β-d-Glcp-(1→3)-β-d-Glcp-(1→4)-β-d-Glcp-(1→4)-d-Glcp, β-d-Glcp-(1→4)-β-d-Glcp-(1→3)-β-d-Glcp-(1→4)-d-Glcp, and β-d-Glcp-(1→4)-β-d-Glcp-(1→4)-β-d-Glcp-(1→3)-d-Glcp, corresponding to the tri- and tetra-saccharide units in the linear chains of (1→4)- and (1→3)-linked β-d-glucopyranosyl residues of lichenan, and of oat and barley β-d-glucans.  相似文献   

10.
β-d-Mannosidase (β-d-mannoside mannohydrolase EC 3.2.1.25) was purified 160-fold from crude gut-solution of Helix pomatia by three chromatographic steps and then gave a single protein band (mol. wt. 94,000) on SDS-gel electrophoresis, and three protein bands (of almost identical isoelectric points) on thin-layer iso-electric focusing. Each of these protein bands had enzyme activity. The specific activity of the purified enzyme on p-nitrophenyl β-d-mannopyranoside was 1694 nkat/mg at 40° and it was devoid of α-d-mannosidase, β-d-galactosidase, 2-acet-amido-2-deoxy-d-glucosidase, (1→4)-β-d-mannanase, and (1→4)-β-d-glucanase activities, almost devoid of α-d-galactosidase activity, and contaminated with <0.02% of β-d-glucosidase activity. The purified enzyme had the same Km for borohydride-reduced β-d-manno-oligosaccharides of d.p. 3–5 (12.5mm). The initial rate of hydrolysis of (1→4)-linked β-d-manno-oligosaccharides of d.p. 2–5 and of reduced β-d-manno-oligosaccharides of d.p. 3–5 was the same, and o-nitrophenyl, methylumbelliferyl, and naphthyl β-d-mannopyranosides were readily hydrolysed. β-d-Mannobiose was hydrolysed at a rate ~25 times that of 61-α-d-galactosyl-β-d-mannobiose and 63-α-d-galactosyl-β-d-mannotetraose, and at ~90 times the rate for β-d-mannobi-itol.  相似文献   

11.
The effects of polysaccharide structure and environment on the formation of fluorescent complexes between the polysaccharide and a fluorochrome (4,4′ - [carbonylbis (benzene-4,1-diyl) bis(imino)] bisbenzenesulfonic acid (Sirofluor) isolated from the triarylmethane dye, aniline blue, have been studied. Amongst the wide range of water-soluble polysaccharides tested, fluorescent complexes are formed only with glucans, the strongest fluorescence being obtained with linear (1 → 3)-β-d-glucans and with linear (1 → 3)-β-d-glucans bearing single glucose residues attached at the 6-position. The fluorescence of complexes formed with water-insoluble polysaccharides depends on the ionic environment as well as the polysaccharide structure. (1 → 3)-β-d-Glucans form strongly fluorescent complexes in the dry state and in the presence of water or phosphate buffer. Various cellulose ((1 → 4)-β-d-glucan) samples form strongly fluorescent complexes in the dry state and in the presence of phosphate buffer, but are significantly reduced in the presence of water alone.  相似文献   

12.
An exo-β-glucan hydrolase, present in the digestive juice of the snail, Helix pomatia, has been purified to homogeneity by chromatography on Bio-Gel P-60, Sephadex G-200, DEAE-cellulose, and DEAE-Sephadex. The enzyme degrades β-(1 → 3)-linked oligosaccharides and polysaccharides, rapidly and to completion, or near completion, yielding glucose as the major product of enzyme action. Mixed linkage (1→3; 1→4)-β-glucans are also extensively degraded and β-(1→6)- and β-(1→4)-linked glucose polymers are slowly degraded by the enzyme. This enzyme differs from other exo-β-glucanases, reported previously, in the broadness of its substrate specificity. The Km values for action on laminarin and lichenin are respectively 1.22 and 2.22 mg/ml; the maximum velocity of action on laminarin is approximately twice that on lichenin. The enzyme has a molecular weight of 82,000 as determined by polyacrylamide gel electrophoresis. Maximum activity is exhibited at pH 4.3 and at temperatures of 50–55 °C.  相似文献   

13.
The specificity of the sialidase activity present in rat kidney cortex (12 000 × g pellet) was studied with various tritiated oligosaccharidic substrates: (i) αNeuAc2 → 3βGall → 4Glc-itol[3H], αNeuAc2 → 6βGall → 4Glc-itol[3H] and αNeuAc2 → 8αNeuAc2 → 3βGall → 4Glc-itol[3H] from bovine colostrum; (ii) α-NeuAc2 → 6βGall → 4βGlcNAc-itol[3H], αNeuAc2 → 3βGal1 → 4βGlcNAcl → 2αManl → 3βMan1 → 4GlcNAc-itol[3H]. αNeuAc2 → 6βGall → 4βGlcNAcl → 2αManl α 3(βGall → 4GlcNAcl → 2αManl → 6)βManl → 4GlcNAc-itol [3H]et αNeuAc2 → 6βGall → 4βGlcNAcl → 2αManl-3(αNeuAc2 → 6βGall → 4βGlcNAcl → 2αManl → 6)βManl 4GlNAc-itol[3H] isolated from the urine of a patient with mucolipidosis I. The enzyme cleaves α2 → 3 and α2 → 8 linkages at a greater rate than the α2 → 6 bonds. Its activity decreases with the length of the oligosaccharidic chain. Substitution of a glucose moiety by Nacetylglucosamine results in diminished activity. The specificity of rat kidney sialidase differs from that reported for other mammalian of viral sialidases.  相似文献   

14.
We characterized a glycoside hydrolase family 112 protein from Opitutus terrae (Oter_1377 protein). The enzyme phosphorolyzed d-galactosyl-β1→4-l-rhamnose (GalRha) and also showed phosphorolytic activity on d-galactosyl-β1→3-d-glucose as a minor substrate. In the reverse reaction, the enzyme showed higher activity on l-rhamnose derivatives than on d-glucose derivatives. The enzyme was stable up to 45 °C and at pH 6.0–7.0. The values of kcat and Km of the phosphorolytic activity of the enzyme on GalRha were 60 s?1 and 2.1 mM, respectively. Thus, Oter_1377 protein was identified as d-galactosyl-β1→4-l-rhamnose phosphorylase (GalRhaP). The presence of GalRhaP in O. terrae suggests that genes encoding GalRhaP are widely distributed in different organisms.  相似文献   

15.
Uridine diphosphate glucose pyrophosphorylase (UDP-Glc PPase, EC2.7.7.9) was purified 65 fold from immature grains of wheat (Triticum aestivum L. cv, WH-147) by ammonium sulphate fractionation, DEAE-cellulose anion exchange chromatography and Sephadex G-100 permeation chromatography. The partially purified enzyme, having molecular weight of 72 kD, exhibited broad pH optimum between 8 and 9 and was stable at 4°C for 15 days. At pH 8.5, the enzyme followed typical hyperbolic kinetics with respect to UDP-glucose and inorganic pyrophosphate (Km 0.22 mM and 0.66 mM respectively). The enzyme showed absolute requirement for Mg2+ and did not appear to require sulfhydryl groups for its activity. Initial velocity and product inhibition studies indicated sequential addition of substrates and sequential release of products.  相似文献   

16.
Changes in (1→3,1→4)-β-D-glucan endohydrolase (EC 3.2.1.73) protein levels were investigated in segments from second leaves of wheat (Triticum aestivum L.). The abundance of the enzyme protein markedly increased when leaf segments were incubated in the dark whereas the enzyme rapidly disappeared when dark-incubated segments were illuminated or fed with sucrose. Addition of cycloheximide (CHI) to the incubation medium led to the disappearance of previously synthesized (1→3,1→4)-β-glucanase and suppressed the dark-induced accumulation indicating that the enzyme was rather unstable. The degradation of (1→3,1→4)-β-glucanase was analyzed without the interference of de-novo synthesis in intercellular washing fluid (IWF). The loss of the enzyme protein during incubation of IWF (containing naturally present peptide hydrolases) indicated that the stability increased from pH 4 to pH 7 and that an increase in the temperature from 25 to 35 °C considerably decreased the stability. Chelating divalent cations in the IWF with o-phenanthroline also resulted in a lowered stability of the enzyme. A strong temperature effect in the range from 25 to 35 °C was also observed in wheat leaf segments. Diurnal changes in (1→3,1→4)-β-glucanase activity were followed in intact second leaves from young wheat plants. At the end of the dark period, the activity was high but constantly decreased during the light phase and remained low if the light period was extended. Activity returned to the initial level during a 10-h dark phase. During a diurnal cycle, changes in (1→3,1→4)-β-glucanase activity were associated with reciprocal changes in soluble carbohydrates. The results suggest that the synthesis and the proteolytic degradation of an apoplastic enzyme may rapidly respond to changing environmental conditions.  相似文献   

17.
《Carbohydrate research》1987,163(1):115-122
Acidic cyclosophorans [cyclic (1→2)-β-d-glucans] containing methylmalonic acid, or succinic acid, or both, were isolated by DEAE-cellulose chromatography from culture filtrates and cells of some strains of Agrobacterium radiobacter, Rhizobium phaseoli, and R. trifolii. The evidence suggests that one carboxyl group of the dicarboxylic acid is in ester linkage with an hydroxyl group of a sugar unit.  相似文献   

18.
Mutational experiments were performed to decrease the protease productivity of Humicola grisea var. thermoidea YH-78 using UV light and N-methyl-N′-nitro-N-nitrosoguanidine. A protease-negative mutant, no. 140, exhibited higher endoglucanase activity than the parent strain in mold bran culture at 50°C for 4 days. The culture extract rapidly disintegrated filter paper but produced a small amount of reducing sugar. About 30% of total endoglucanase activity in the extract was adsorbed onto Avicel. The electrophoretically homogeneous preparation of Avicel-adsorbable endoglucanase (molecular weight, 128,000) showed intensive filter-paper-disintegrating activity but did not release reducing sugar. The preparation also exhibited a highly synergistic effect with the cellulase preparation from Trichoderma reesei in the hydrolysis of microcrystalline cellulose. This endoglucanase was observed via scanning electron microscopy to disintegrate Avicel fibrils layer by layer from the surface, yielding thin sections with exposed chain ends. A mutant, no. 191, producing higher protease activity and an Avicel-unadsorbable, Avicel-nondisintegrating endoglucanase was isolated. The purified enzyme (molecular weight, 63,000) showed no disintegrating activity on filter paper and Avicel and a less synergistic effect with the T. reesei cellulase in hydrolyzing microcrystalline cellulose than did the former enzyme. Endoglucanase was therefore divided into two types, Avicel disintegrating and Avicel nondisintegrating.  相似文献   

19.
Structural features of noncellulosic β-d-glucans of Zea mays, Hordeum vulgare, Triticum vulgare, Secale cereale, and Sorghum bicolor were compared. Treatment of cell walls derived from these species with specific Bacillus subtilis or Rhizopus glucanases yields virtually identical profiles upon Bio-Gel P-2 fractionation of the liberated oligosaccharides. The two predominant reaction products, a trisaccharide and tetrasaccharide, were identified as 3-O-β-cellobiosyl-d-glucose and 3-O-β-cellotriosyl-d-glucose respectively by virtue of the specificity of these enzymes and by paper chromatography and electrophoresis. The similarity of the reaction product profiles indicates a rather regular repeating sequence in all β-d-glucans examined. The ratios of 3-O-β-cellobiosyl-d-glucose to 3-O-β-cellotriosyl-d-glucose indicates that 30.4–30.9% of the β-glucosyl linkages in the intact molecule are 1 → 3. The yields of wall glucan as estimated from the quantity of oligosaccharides released, range from 41 μg/mg wall in Hordeum to 97 μg/mg wall from Sorghum.  相似文献   

20.
A (1→3),(1→4)-β-glucan synthase catalysing the synthesis of (1→3),(1→4)-β-glucan (mixed-linkage glucan) was investigated using microsomal membranes prepared from developing barley (Hordeum vulgare L. cv. Shikokuhadaka 97) endosperms harvested 21 days after flowering. The microsomal fraction produced (1→3),(1→4)-β-glucan by incorporation of [14C]Glc from UDP-[14C]Glc. The production of (1→3),(1→4)-β-glucan was ascertained by specific enzymatic digestion with endo-(1→3),(1→4)-β-glucanase (lichenase; EC 3.2.1.73) from Bacillus amyloliquefaciens, which released a radiolabelled trisaccharide (3-O-β-cellobiosyl-glucose) and a tetrasaccharide (3-O-β-cellotriosyl-glucose), the diagnostic oligosaccharides for the identification of (1→3),(1→4)-β-glucan. Digestion of the products with exo-(1→3)-β-glucanase (EC 3.2.1.58) from Basidiomycete QM806 released radiolabelled Glc, indicating that not only (1→3),(1→4)-β-glucans but also (1→3)-β-glucans (callose) had been formed due to the presence of (1→3)-β-glucan (callose) synthase (EC 2.4.1.34) in the microsomal fraction. The activity of (1→3),(1→4)-β-glucan synthase was maximal at pH 9.0 and at 25°C and in the presence of at least 2 mM Mg2+. The apparent Km and Vmax values for UDP-Glc were 0.33 mM and 480 pmol min−1 mg protein−1, respectively. Investigating the dependence of enzyme activity on developmental stage (7–35 days after flowering) of the endosperms, we found an increase of activity during the initial development reaching a maximum at 19 days, followed by a gradual decrease as the endosperms matured. The amount of (1→3),(1→4)-β-glucan in the cell walls of the endosperms, however, increased gradually towards maturation, even after 19 days. Analysing the relationship between enzyme activity and (1→3),(1→4)-β-glucan deposition in cell walls of endosperms prepared from 12 different barley varieties harvested 11–22 days after flowering showed that some varieties had both low activity and low glucan content, and in some both were high. But for several other varieties, the availability of donor substrate and other factors seem to influence the production of (1→3),(1→4)-β-glucan as well.  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