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 218 毫秒
1.
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   

2.
Monoclonal antibodies were raised against a (1→3,1→4)-β-glucan-bovine serum albumin (BSA) conjugate. One antibody (BG1) selected for further characterization, was specific for (1→3,1→4)-β-glucan, displaying no binding activity against a (1→3)-β-glucan-BSA conjugate and minimal binding against a cellopentaose-BSA conjugate. A range of oligosaccharides was prepared by enzymatic digestion of (1→3,1→4)-β-glucan, purified by size exclusion chromatography and characterized by 1H-NMR and anion exchange chromatography. These (1→3,1→4)-β-oligoglucosides, together with (1→3)-β- and (1→4)-β-oligoglucosides were used to characterize the binding site of the monoclonal antibody (BG1) by competitive inhibition. The monoclonal antibody showed maximal binding to a heptasaccharide with the structure Glc(1→3) Glc(1→4) Glc(1→4) Glc(1→3) Glc(1→4) Glc(1→4) Glc and was determined to have an affinity constant of 3.8 × 104 M−1 for this oligoglucoside. The monoclonal antibody (BG1) has been used to develop a sensitive sandwich ELISA for the specific quantitation of (1→3,1→4)-β-glucans. The assay operates in the range 1–10 ng ml−1 and shows no significant cross-reaction with tamarind xyloglucan, wheat endosperm arabinoxylan or carboxymethyl-pachyman ((1→3)-β-glucan). When used with a second-stage, rabbit anti-mouse gold conjugate and viewed under the electron microscope, the monoclonal antibody probe was found to bind strongly to the walls of the aleurone in thin sections of immature wheat (Triticum aestivum) cv. Millewa grains but not to the middle lamella region. A previously described specific anti-(1→3)-β-glucan antibody (Meikle et al., 1991) bound to discrete patches on the aleurone walls, believed to be plasmodesmata.  相似文献   

3.
The evolutionary relationships of two classes of plant β-glucan endohydrolases have been examined by comparison of their substrate specificities, their three-dimensional conformations and the structural features of their corresponding genes. These comparative studies provide compelling evidence that the (1→3)-β-glucanases and (1→3,1→4)-β-glucanases from higher plants share a common ancestry and, in all likelihood, that the (1→3,1→4)-β-glucanases diverged from the (1→3)-β-glucanases during the appearance of the graminaceous monocotyledons. The evolution of (1→3,1→4)-β-glucanases from (1→3)-β-glucanases does not appear to have invoked ‘modular’ mechanisms of change, such as those caused by exon shuffling or recombination. Instead, the shift in specificity has been acquired through a limited number of point mutations that have resulted in amino acid substitutions along the substrate-binding cleft. This is consistent with current theories that the evolution of new enzymic activity is often achieved through duplication of the gene encoding an existing enzyme which is capable of performing the required chemistry, in this case the hydrolysis of a glycosidic linkage, followed by the mutational alteration and fine-tuning of substrate specificity. The evolution of a new specificity has enabled a dramatic shift in the functional capabilities of the enzymes. (1→3)-β-Glucanases that play a major role, inter alia, in the protection of the plant against pathogenic microorganisms through their ability to hydrolyse the (1→3)-β-glucans of fungal cell walls, appear to have been recruited to generate (1→3,1→4)-β-glucanases, which quite specifically hydrolyse plant cell wall (1→3,1→4)-β-glucans in the graminaecous monocotyledons during normal wall metabolism. Thus, one class of β-glucan endohydrolase can degrade β-glucans in fungal walls, while the other hydrolyses structurally distinct β-glucans of plant cell walls. Detailed information on the three-dimensional structures of the enzymes and the identification of catalytic amino acids now present opportunities to explore the precise molecular and atomic details of substrate-binding, catalytic mechanisms and the sequence of molecular events that resulted in the evolution of the substrate specificities of the two classes of enzyme.  相似文献   

4.
The effects of the lysine-reactive chemical modification reagents, uridine 5’ diphospho (UDP)-pyridoxal and formaldehyde (HCHO), on the activity of membrane-bound and solubilized UDP-Glc: (1,3)-β-D-glucan synthase (callose synthase) from red beet (Beta vulgaris L.) storage tissue were compared. Exposure to micromolar levels of UDP-pyridoxal, or millimolar levels of HCHO in the presence of NaCNBH3, resulted in complete enzyme inactivation. Conditions for inhibition of membrane-bound enzyme activity by the two reagents were markedly similar; divalent cations were required for inactivation, and complete protection of activity was obtained with EDTA or EGTA. The substrate, UDP-Glc, protected membrane-bound callose synthase against inactivation by UDP-pyridoxal or HCHO, but protected the solubilized enzyme only against inhibition by UDP-pyridoxal, suggesting that the lysine residue modified by both these reagents is at the enzyme active site, and that the site is more open or has a certain conformational flexibility in the solubilized enzyme. Potential UDP-Glc-binding polypeptides of callose synthase were identified by a two-step labeling procedure. First, nonessential lysine residues were blocked by irreversible modification reaction with HCHO or UDP-pyridoxal in the presence of UDP-Glc to protect lysines at UDP-Glc-binding sites. In the second step, proteins were recovered, reacted with [14C]-HCHO in the absence of UDP-Glc, and polypeptide labeling patterns analyzed by SDS-polyacrylamide gel electrophoresis and fluorography. This procedure reduced incorporation of label by 5- to 8-fold compared to a procedure omitting the preblocking step, and with enzyme partially purified by solubilization in CHAPS followed by product entrapment, labeling was limited to a small set of polypeptides. Taken together with the results of other studies, the data suggest that one or more polypeptides migrating in the 54–57 kDa region are good candidates for the UDP-Glc-binding components of callose synthase.  相似文献   

5.
The action on tamarind seed xyloglucan of the pure, xyloglucan-specific endo-(1→4)-β-D-glucanase from nasturtium (Tropaeolum majus L.) cotyledons has been compared with that of a pure endo-(1→4)-β-D-glucanase (‘cellulase’) of fungal origin. The fungal enzyme hydrolysed the polysaccharide almost completely to a mixture of the four xyloglucan oligosaccharides: Exhaustive digestion with the nasturtium enzyme gave the same four oligosaccharides plus large amounts of higher oligosaccharides and higher-polymeric material. Five of the product oligosaccharides (D,E,F,G,H) were purified and shown to be dimers of oligosaccharides A to C. D (glc8xyl6) had the structure A→A, H (glc8xyl6gal4) was C→C, whereas E (glc8xyl6gal), F (glc8xyl6gal2) and G (glc8xyl6gal3) were mixtures of structural isomers with the appropriate composition. For example, F contained B2→B2 (30%), A→C (30%), C→A (20%), B2B1 (15%) and others (about 5%). At moderate concentration (about 3 mM) oligosaccharides D to H were not further hydrolysed by the nasturtium enzyme, but underwent transglycosylation to give oligosaccharides from the group A, B, C, plus higher oligomeric structures. At lower substrate concentrations, hydrolysis was observed. Similarly, tamarind seed xyloglucan was hydrolysed to a greater extent at lower concentrations. It is concluded that the xyloglucan-specific nasturtium-seed endo-(1→4)-β-D-glucanase has a powerful xyloglucan-xyloglucan endo-transglycosylase activity in addition to its known xyloglucan-specific hydrolytic action. It would be more appropriately classified as a xyloglucan endo-transglycosylase. The action and specificity of the nasturtium enzyme are discussed in the context of xyloglucan metabolism in the cell walls of seeds and in other plant tissues.  相似文献   

6.
Read SM  Delmer DP 《Plant physiology》1987,85(4):1008-1015
UDP-pyridoxal competitively inhibits the Ca2+-, cellobiose-activated (1→3)-β-glucan synthase activity of unfractionated mung bean (Vigna radiata) membranes, with a Ki of 3.8 ± 0.7 micromolar, when added simultaneously with the substrate UDP-glucose in brief (3 minute) assays. Preincubation of membranes with UDP-pyridoxal and no UDP-glucose, however, causes progressive reduction of the Vmax of subsequently assayed enzyme and, after equilibrium is reached, 50% inhibition occurs with 0.84 ± 0.05 micromolar UDP-pyridoxal. This progressive inhibition is reversible provided that the UDP-pyridoxylated membranes are not treated with borohydride, indicating formation of a Schiff's base between the inhibitor and an enzyme amino group. Consistent with this, UDP-pyridoxine is not an inhibitor. The reaction of (1→3)-β-glucan synthase with UDP-pyridoxal is stimulated strongly by Ca2+ and, less effectively, by cellobiose or sucrose, and the enzyme is protected against UDP-pyridoxal by UDP-glucose or by other competitive inhibitors, implying that modification is occurring at the active site. Pyridoxal phosphate is a less potent and less specific inhibitor. Latent (1→3)-β-glucan synthase activity inside membrane vesicles can be unmasked and rendered sensitive to UDP-pyridoxal by the addition of digitonin. Treatment of membrane proteins with UDP-[3H]pyridoxal and borohydride labels a number of polypeptides but labeling of none of these specifically requires Ca2+ and sucrose; however, a polypeptide of molecular weight 42,000 is labeled by UDP-[3H]pyridoxal in the presence of Mg2+ and copurifies with (1→3)-β-glucan synthase activity.  相似文献   

7.
1,3-β-Glucan synthase activity has been demonstrated in particulate fractions of bark extracts from Mexican lime. With respect to substrate, the enzyme kinetics did not conform to the Michaelis-Menten equation. The value of the Hill coefficient was 1.2 and S0.5 is 1.1 mM. The enzyme had an optimum pH of 7.5. Maltose, sucrose, and especially cellobiose and glucose, were enzyme activators when tested at physiological concentrations. In the presence of 15 mM MgCl2 the enzymic activity was stimulated at 10 μM UDP-glucose but decreased at 1 mM UDP-glucose, suggesting a minor 1,4-β-glucan synthase activity.  相似文献   

8.
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   

9.
(1->3)- and (1->4)-[beta]-glucan synthase activities from higher plants have been physically separated by gel electrophoresis in nondenaturing conditions. The two glucan synthases show different mobilities in native polyacrylamide gels. Further separation by sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed a different polypeptide composition in these synthases. Three polypeptides (64, 54, and 32 kD) seem to be common to both synthase activities, whereas two polypeptides (78 and 38 kD) are associated only with callose synthase activity. Twelve polypeptides (170, 136, 108, 96, 83, 72, 66, 60, 52, 48, 42, and 34 kD) appear to be specifically associated with cellulose synthase activity. The successful separation of (1->3)- and (1->-4)-[beta]-glucan synthase activities was based on the manipulation of digitonin concentrations used in the solubilization of membrane proteins. At low dipitomin concentrations (0.05 and 0.1%), the ratio of the cellulose to callose synthase activity was higher. At higher digitonin (0.5-1%) concentrations, the ratio of the callose to cellulose synthase activity was higher. Rosette-like particles with attached product were observed in samples taken from the top of the stacking gel, where only cellulose was synthesized. Smaller (nonrosette) particles were found in the running gel, where only callose was synthesized. These findings suggest that a higher level of subunit organization is required for in vitro cellulose synthesis in comparison with callose assembly.  相似文献   

10.
The mode of inhibition of UDP, one of the products of the reaction catalyzed by (1→3)-β-d-glucan synthase in sugar beet (Beta vulgaris L.) was investigated. In the absence of added UDP, the enzyme, in the presence of Ca2+, Mg2+, and cellobiose, exhibited Michaelis-Menten kinetics and had an apparent Km of 260 micromolar for UDP-glucose. Complex effects on the kinetics of the (1→3)-β-d-glucan synthase were observed in the presence of UDP. At high UDP-glucose concentrations, i.e. greater than the apparent Km, UDP behaved as a competitive inhibitor with an apparent Ki of 80 micromolar. However, at low UDP-glucose concentrations, reciprocal plots of enzyme activity versus substrate concentration deviated sharply from linearity. This unusual effect of UDP is similar to that reported for fungal (1→3)-β-d-glucan synthase. However, papulacandin B, a potent inhibitor of this fungal enzyme, had no effect on the plant (1→3)-β-d-glucan synthase isolated from sugar beet petioles. The inhibitory effect of UDP was also compared with other known inhibitors of glucan synthases.  相似文献   

11.
The physical dimensions of several (1 → 6) branched (1 → 3) -β-D -glucan samples obtained from different organisms and their derivatives have been studied by electron microscopy, light scattering measurements, viscometry, and gel permeation chromatography. The electron micrographs indicate that in most samples these biopolymers are adequately described as linear worm-like coils. A sample reconstituted from alkaline media appeared as a blend of the linear, circular, and aggregated polymer morphologies. The average mass per unit length, ML = Mw/Lw for the macroscopically linear samples, was estimated to be 2100 ± 200 g mol?1 nm?1. The parameter mL was determined from the contour lengths obtained by electron microscopy and the molecular weight by light scattering measurements. The observed ML was consistent with the triple-helical structure reported from x-ray diffraction studies and observed degree of side-chain substitution. From the molecular snapshots shown in the electron micrographs, the persistence lengths of these β-D -glucans were determined to be 140 ± 30 nm. The experimentally determined intrinsic viscosities were consistent with these estimates of ML and persistence length. Comparison of the molecular weight distributions obtained from gel permeation chromatography and those deduced from the electron micrographs indicates that number and weight average contour lengths are more reliable than z and z + 1 averages. © 1993 John Wiley & Sons, Inc.  相似文献   

12.
13C-nmr has been employed to probe the molecular conformation and crystal structure of (1 → 6)-β-D -glucan (pustulan) in the solution, gel, and solid states. CP/MAS 13C-nmr spectra recorded for partially crystalline solid pustulan display a resonance near 82 ppm that is absent in solution spectra. The intensity and peak width of this resonance were found to depend on relative crystallinity as determined by x-ray diffraction. CP/MAS spectra of aqueous pustulan gels also exhibit the 82-ppm resonance, suggesting that the gelation mechanism may involve microcrystalline junction zones. Since the 82-ppm resonance is absent in the CP/MAS spectrum of the (1 → 6)-β-linked dimer gentiobiose, we tentatively conclude the crystal structure of this dimer does not adequately model the yet undetermined structure of pustulan.  相似文献   

13.
We describe a approach for the synthesis of (1→3)-β-D-oligosaccharide derivatives 10–18. 1–9 were synthesized by treating peracetylated (1→3)-β-D-oligosaccharides with the corresponding alkenyl alcohols and Lewis acid (SnCl4) catalyst. Epoxidation of the corresponding alkenyl oligoglucosides took place by m-CPBA. NaOMe in dry methanol was used for the deacetylation of the blocked derivatives, to give 10–18 in an overall yields of 25–32%. In subsequent glucan-binding protein of soybean assays, we found that 16 was most active, with an IC50 value of 9 mM. However, the activities of 17, 18, 13, 14, 15, 10, 11, and 12 were gradually decreased. At the same time, we found 16 was most active as compared to the other (1→3)-β-D- oligoglucoside derivatives in eliciting phytoalexin accumulation in soybean cotyledon tissue, and 16 was kept longer time than (1→3)-β-D-glucohexaose, which indicated 16 is much more stable than (1→3)-β-D-glucohexaose. Published in 2004.  相似文献   

14.
Sinorhizobium meliloti synthesizes a linear mixed-linkage (1 → 3)(1 → 4)-β-d -glucan (ML β-glucan, MLG) in response to high levels of cyclic diguanylate (c-di-GMP). Two proteins BgsA and BgsB are required for MLG synthesis, BgsA being the glucan synthase which is activated upon c-di-GMP binding to its C-terminal domain. Here we report that the product of bgrR (SMb20447) is a diguanylate cyclase (DGC) that provides c-di-GMP for the synthesis of MLG by BgsA. bgrR is the first gene of a hexacistronic bgrRSTUWV operon, likely encoding a partner-switching regulatory network where BgrR is the final target. Using different approaches, we have determined that the products of genes bgrU (containing a putative PP2C serine phosphatase domain) and bgrW (with predicted kinase effector domain), modulate the phosphorylation status and the activity of the STAS domain protein BgrV. We propose that unphosphorylated BgrV inhibits BgrR DGC activity, perhaps through direct protein–protein interactions as established for other partner switchers. A bgrRSTUWV operon coexists with MLG structural bgsBA genes in many rhizobial genomes but is also present in some MLG non-producers, suggesting a role of this partner-switching system in other processes besides MLG biosynthesis.  相似文献   

15.
A (1→3)-β-glucan synthase has been isolated from petiole tissue of sugar beet (Beta vulgaris L.). Enzyme activity is associated with a membrane fraction with a density of 1.03 grams per cubic centimeter when subjected to isopycnic density gradient centrifugation in Percoll. The reaction product was determined to be a linear (1→3)-β-glucan by methylation analysis and by glucanase digestion. (1→3)-β-Glucan synthase activity is markedly stimulated by Ca2+; activation is half-maximal at about 50 micromolar Ca2+ and is nearly saturated at 100 micromolar. Other divalent cations tested, Mg2+, Mn2+, and Sr2+, also stimulate enzyme activity but are less effective. Enzyme activity was also stimulated up to 12-fold by β-glucosides. Sirofluor, the fluorochrome from aniline blue, inhibited enzyme activity 95% when included at 1 millimolar. The enzyme was solubilized in Zwittergent 3-14; 85% of total enzyme activity was solubilized in 0.03% detergent and the optimal detergent-to-protein ratio was 0.3 at 3 milligrams per milliliter protein.  相似文献   

16.
Fibrinogen showed essentially no binding (KD>1 mM ) to platelet αIIbβ3 integrin in solution in the presence of Triton or octylglucoside above critical micellar concentrations. Under these conditions the integrin was an αβ monomer. After removal of the detergent from the Triton containing buffer (25 mM Tris/HCl;, 150 mM NaCl, 1 mM CaCl2, 1 mM MgCl2, pH 7.4) the integrin formed aggregates with hexamers as the most prominent species, as demonstrated by analytical ultracentrifugation and electron microscopy. Tracer sedimentation equilibrium experiments indicate that fibrinogen binds to the integrin aggregates, but with a surprisingly large KD (at least 3 μM ). This value is 10- to 100-fold higher than values determined by solid phase assays or with integrins reconstituted onto lipid bilayers.  相似文献   

17.
The bacterial strain PP710, isolated from soil and identified as Paenibacillus species, produced a low-digestibility α-glucan containing a large amylase-resistant portion. This α-glucan was obtained in high yields from maltodextrin (dextrose equivalent 3) by using the condensed culture supernatant of the strain as the enzyme preparation. The water-soluble dietary fiber content of the low-digestibility α-glucan was 80.2%, and showed resistance to a rat intestinal enzyme preparation. The α-glucan was found to be a novel highly branched α-glucan by acid hydrolysis, NMR analysis, gel permeation chromatography, methylation analysis, and enzymatic digestion.  相似文献   

18.
The evidence accumulated to date indicates that 1,3-β-glucan synthase (EC 2.3.1.12) and 1,4-β-glucan synthase (EC 2.4.1.12) are regulated by different effectors. Further that the same synthase has different effectors, depending upon its presence in green plants, fungi, and bacteria. Synthases from plants require divalent cations and β-linked glucosides whereas fungal enzymes require neither cations nor β-glucosides, but most require nucleoside triphosphates for activation. Two endogenous effectors have been characterized and shown to produce activation in vitro. One is 3',5'-cyclic diguanylic acid that is the activator of cellulose synthase in bacteria. The other is a β-linked glucosyl dioleoyl diglyceride from mung bean, capable of activating synthases that produce both β-(1–3) and β-(1–4) products. The results of product analysis of the β-linked glucoside activated reaction suggest that the synthesis of (1–3) and (1–4) glucosyl linkages may share a common enzyme in plants. All synthases utilize uridine 5'-diphosphoglucose (UDPG) and are associated with the plasma membrane. Efforts to solubilize the synthases from cellular fractions enriched in plasma membranes have been generally successful. The purification of the soluble enzymes, however, remains a major obstacle to the full understanding of their regulation.  相似文献   

19.
The effect of salinity on the non-enzymic and enzymic antioxidant activity, shoot proliferation and nutrient accumulation was studied in in vitro cultures of the rootstock CAB-6P (Prunus cerasus L.). Three concentrations (0, 30 and 60 mM) of NaCl or CaCl2 were added to a modified MS medium. Between the two salt treatments used, only the explants treated with CaCl2 presented significant decrease in growth parameters. The concentrations of Na+ and Cl in the explants treated with NaCl were increased, as NaCl in the culture medium increased. Furthermore, in the explants treated with CaCl2 the concentrations of Ca2+ and Cl were increased while that of K+ decreased, as CaCl2 concentration increased. The activity of peroxidase in leaves as well as the number of its anionic isoforms was increased under 30 mM CaCl2 as well as 60 mM NaCl or CaCl2. On the contrary, increasing salinity, from 0 to 60 mM CaCl2, resulted in a reduction of the catalase activity in leaves followed by disappearance of the only one catalase isoform that was detected in leaves (60 mM CaCl2). In the stems of the explants treated with NaCl the peroxidase activity was reduced. In the stems and leaves of the explants grown in saline substrate the non-enzymic antioxidant activity was significantly increased. The results suggest that the stems and leaves of CAB-6P explants presented variable antioxidant responses that were depended on the salt form used. The contribution of enzymic and non-enzymic protection mechanisms to the adaptation of CAB-6P explants under salinity stress is discussed.  相似文献   

20.
A small quantity of (1→3)-β-d-glucan was extracted with a (1→3),(1→4)-β-d-glucan by hot water after treatment of the insoluble fraction of a buffer homogenate of Zea shoots with 3 molar LiCl. An ammonium sulfate precipitation procedure effected a separation of the (1→3)-β-d-glucan from the more prevalent (1→3),(1→4)-β-d-glucan. The minor component polysaccharide precipitated at a concentration of 20% ammonium sulfate (w/v) and was, as a consequence of precipitation, rendered insoluble in water. The insoluble products were dissolved in 1 normal NaOH followed by neutralization with CH3COOH. The purified polysaccharide accounted for approximately 0.3% of total hot water extract. It consisted mostly of glucose and its average mol wt was estimated to be about 7.0 × 104, based on elution from a calibrated Sepharose CL-4B column. Methylation analysis and enzymic hydrolysis or partial acid-hydrolysis of the polysaccharide followed by analysis of the hydrolysate showed that the polysaccharide consisted of (1→3)-β-linked glucose residues.  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