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.
The isomaltodextranase (EC 3.2.1.94) from Arthrobacter globiformis T6 hydrolysed thirteen dextrans to various extents (11?64% after 13 days) at initially large but gradually decreasing rates. Dextran B-1355 fraction S was, unlike the other dextrans, hydrolysed by the dextranase initially at the lowest rate among the dextrans used, but the rate was maintained for a long period with little decrease, so that the hydrolysis reached as high as 85% after 13 days. Paper chromatography of these dextran digests revealed that this dextranase produces in addition to isomaltose, one or two trisaccharides [isomaltose residues substituted by (1 →2)-, (1→3)-, or (1→4)-α-D-glucopyranosyl groups at the non-reducing D-glucopyranosyl residues] from every dextran used. It is evident that the non-(1→6)-linkages of these trisaccharide products constitute the “anomalous” linkages of the corresponding dextrans. The relative amounts of these trisaccharide products appear to indicate the approxima te relative amounts of a particular linkage among the dextrans, or the relative amounts of two kinds of linkages of each dextran. The kinds and the relative amounts of “anomalous” linkages of some dextrans were established on the basis of the trisaccharides produced by isomaltodextranase.  相似文献   

2.
A structural study of the water-soluble dextran made by Leuconostoc mesenteroides strain C (NRRL B-1298) was conducted by enzymic degradation and subsequent 13C-NMR analysis of the native dextran and its limit dextrins. The α-l,2-debranching enzyme removed almost all of the branched D-glucose residues, and gave a limit dextrin having a much longer sequence of the internal chain length (degree of linearity: n = 24.5 compared with the value of n = 3.3 for the native dextran). The degree of hydrolysis with debranching enzyme corresponded to the content of α-1,2-linkages determined by chemical methods, which suggested that most of the α-l,2-linkages in the dextran B-1298 constituted branch points of a single D-glucose residue. A synergistic increase of susceptibility of the dextran B-1299 was observed by simultaneous use of debranching enzyme and endodex-tranase. 13C-NMR spectral analysis indicated the similarity of structure of dextran B-1298 to that of B-1396, rather than that of B-1299. Occurrence of α-l,3-linkages in the limit dextrin was supported by a newly visualized chemical shift at 83.7 ppm.  相似文献   

3.
An exo-l,6-α-glucosidase (EC 3.2.1.70) (glucodextranase) produced extraceUularly by Arthrobacter globiformis I42 was found to invert the configuration of glucose released from dextran, and to require calcium for protection against warming. Among isomaltodextrins used as substrates for this enzyme, the rate of hydrolysis for isomaltose was the lowest and increased with the degree of polymerization (d. p.) of the saccharides up to d. p. 7. The minor activities accompanying purified glucodextranase preparations (release of glucose from starch, splitting of maltose, nigerose and kojibiose) were ascribed to the glucodextranase itself. Fourteen native dextrans and soluble potato starch were subjected to digestion by this glucodextranase and the rate, process and extent of hydrolysis of these substrates were studied relative to the composition of non-l,6-α-linkages of these polysaccharides.  相似文献   

4.
Dextran fractions from NRRL strains Leuconostoc mesenteroides B-1299 and B-1399, and the native, structurally homogeneous dextrans from L. mesenteroides. B-640, B-1396, B-1422, and B-1424, were examined by 13C-n.m.r. spectroscopy at 34 and at 90°, and by g.l.c.-m.s. The 13C-n.m.r. data indicate that the dextrans of this series branch exclusively through α-d-(1→2)-linkages, and differ from one another only in degree of linearity. Diagnostic, 13C-n.m.r resonances, correlating with 2,6-di-O-substituted α-d-glucosyl residues at branch points, have chemical shifts that are independent of the degree of linearity of the dextran. The intensities of these diagnostic resonances from branching residues, compared to the resonances associated with linear dextran (low degree of branching), are generally proportional to the degree of branching established by methylation-fragmentation analysis. The validity of assignment of the diagnostic, 13C-n. m.r. resonances is substantiated by a critical review of methods previously used to provide structural information on dextrans having α-d-(1→2)-linkages, and by evaluation of the corresponding results on the basis of the ultimate standard-methylation structural analysis.  相似文献   

5.
A newly isolated soil-actinomycete, Actinomadura strain R10 (NRRL B-11411), produces an extracellular isomaltodextranase (optinal pH, 5.0) that was purified to homogeneity. It exolytically releases isomaltose and a minor trisaccharide product,α-d-Glcp-(1→3)-α-d-Glcp, from dextran B-512 and, in addition, forms transient transisomaltosylation products. This pattern of products is qualitatively similar to that previously reported for the isomaltodextranase (EC 3.2.1.94, optimal pH, 4-0) of Arthrobacter globiformis T6 (NRRL B-4425). The Arthrobacter isomaltodextranase is most active on the (1→6)-α-d-glucopyranosidic linkage, but the relative activity increases with the degrees of polymerization of isomalto-oligosaccharide substrates. In contrast, the relative activity of Actinomadura isomaltodextranase is almost constant throughout the same series of substrates, and is much higher on 3 O- and 4-O-α-isomaltosyl-oligosaccharides than that exhibited by the Arthrobacter enzyme; the activity of Actinomadura isomaltodextranase on the α-(1→4) linkage is 3-4 times greater than on the α-(1→6). These results indicate that, generically, the bacterial isomaltodextranase is a glycanase, whereas the actinomycetal enzyme is a glycosidase. This difference is reflected in the hydrolysis of dextrans, especially of dextran B-1355 (fraction S), which has a high content of unbranched α-(1→3) linked residues. In the digest of this dextran with Arthrobacter isomaltodextranse, short-chain fragments accumulated that were absent when the Actinomadura enzyme was employed.  相似文献   

6.
Pig-spleen acid α-D-glucosidase, when purified over 2000-fold, has a molecular weight of ≈ 106,000 and is homogeneous by disc-gel electrophoresis. The enzyme splits reducing, α-D-glucosyl disaccharides and almost completely degrades dextrans that contain (1→3)- and (1→6)-linkages. Dextrans containing (1→2)-linkages are only partially hydrolysed. The kinetic parameters for the acid α-D-glucosidase were obtained by using oligo- and poly-saccharide substrates. Variation of pH, temperature, and inhibitors caused changes in the activity of the acid α-D-glucosidase towards oligo- and poly-saccharide substrates. These results support the earlier suggestion that the enzyme has multiple substrate-binding sites.  相似文献   

7.
Summary Mixed culture fermentations containing a dextranase-producing yeast, Lipomyces starkeyi, and a dextransucrase-producing bacterium, Leuconostoc mesenteroides, produced dextrans of selected size using sucrose and starch. Dextran and starch hydrolyzates, produced by a Lipomyces starkeyi dextranase and amylase, were incorporated into the dextran and formed small size dextran (Mr = 40kDa). The yields of total and clinical dextran in mixed culture fermentation using 12% sucrose and 3% starch were higher, 84 and 79% of theoretical yield, than those using 15% sucrose only, 73 and 69% of theoretical yield, respectively.  相似文献   

8.
Investigation into the effect of the reducing sugar of dextran on formation and stability of dextran-coated ultrasmall superparamagnetic iron oxides (USPIO) has demonstrated that reduction of the terminal reducing sugar can have a significant effect on particle size, coating stability, and magnetic properties. Four aspects of polysaccharide-coated USPIO particle synthesis were investigated: (i) the effect reduction of the terminal polysaccharide sugar has upon polysaccharide usage, particle size, stability, and magnetic susceptibility; (ii) the effect an exogenous reducing sugar can have upon particle synthesis; (iii) the effect the molecular weight of the reduced polysaccharide has on particle synthesis; and (iv) the effectiveness of reduced and native dextrans in stabilizing a preformed magnetic sol. For low molecular weight dextrans (MW 20,000 x 10(-6) cgs). Similar results were obtained with a 12 kDa pullulan. The effect of polysaccharide molecular weight on particle size was studied, wherein higher molecular weight reduced dextrans produced larger particles. The effectiveness of the reduced and native dextrans in stabilizing a preformed magnetic sol was compared. Reduced dextrans were found to be superior for stabilizing the magnetic sol. The observed effects of reduction of the terminal sugar in dextran compared with the native dextran were modeled using the Langmuir adsorption isotherm. A good fit of experimental data with this model was found.  相似文献   

9.
Two kinds of αglucosidase which were homogeneous in disc electrophoretic and ultra-centrifugal analysis were isolated from rice seeds by means of ammonium sulfate fractionation and CM-cellulose, Sephadex G–100 and DEAE-cellulose column chromatography and designated as α-glucosidase I and α-glucosidase II.

Both α-glucosidases hydrolyzed maltose and soluble starch to glucose and showed same optimal pH (4.0) on the both substrates. In addition, both enzymes acted on various α-linked gluco-oligosaccharides and soluble starch but little or not on α-linked hetero-glucosides and α-l,6-glucan (dextran).

Activity of the enzymes on maltose and soluble starch was inhibited by Tris and erythritol. α-Glucosidase II was more sensitive to the inhibitors than α-glucosidase I.

Km value for maltose was 1.1 mM for α-glucosidase I and 2.0 mM for α-glucosidase II.  相似文献   

10.
Certain strains of Gluconobacter oxydans have been known since the 1940s to produce the enzyme dextran dextrinase (DDase; EC2.4.1.2)—a transglucosidase converting maltodextrins into (oligo)dextran. The enzyme catalyses the transfer of an α1,4 linked glucosyl unit from a donor to an acceptor molecule, forming an α1,6 linkage: consecutive glucosyl transfers result in the formation of high molecular weight dextran from maltodextrins. In the early 1990s, the group of K. Yamamoto in Japan revived research on DDase, focussing on the purification and characterisation of the intracellular DDase produced by G. oxydans ATCC 11894. More recently, this was taken further by Y. Suzuki and coworkers, who investigated the properties and kinetics of the extracellular DDase formed by the same strain. Our group further elaborated on fermentation processes to optimise DDase production and dextran formation, DDase characterisation and its use as a biocatalyst, and the physiological link between intracellular and extracellular DDase. Here, we present a condensed overview of the current scientific status and the application potential of G. oxydans DDase and its products, (oligo)dextrans. The production of DDase as well as of dextran is first described via optimised fermentation processes. Specific assays for measuring DDase activity are also outlined. The general characteristics, substrate specificity, and mode of action of DDase as a transglucosidase are described in detail. Two forms of DDase are produced by G. oxydans depending on nutritional fermentation conditions: an intracellular and an extracellular form. The relationship between the two enzyme forms is also discussed. Furthermore, applications of DDase, e.g. production of (oligo)dextran, transglucosylated products and speciality oligosaccharides, are summarized.  相似文献   

11.
Dextran synthesis has been studied since the Second World War, when it was used as blood plasma expander. This polysaccharide composed of glucose units is linked by an α-1,6-glucosidic bond. Dextransucrase is a bacterial extra cellular enzyme, which promotes the dextran synthesis from sucrose. When, besides sucrose, another substrate (acceptor) is also present in the reactor, oligosaccharides are produced and part of the glucosyl moieties from glucose is consumed to form these acceptor products, decreasing the dextran yield. Although dextran enzymatic synthesis has been extensively studied, there are few published studies regarding its molecular weight distribution. In this work, the effect of maltose on yield and dextran molecular weight synthesized using dextransucrase from Leuconostoc mesenteroides B512F, was investigated. According to the obtained results, maltose is not able to control and reduce dextran molecular weight distribution and synthesis carried out with or without maltose presented the same molecular weight distribution profile.  相似文献   

12.
Dextrans are the main exopolysaccharides produced by Leuconostoc species. Other dextran-producing lactic acid bacteria include Streptococci, Lactobacilli, and Weissella species. Commercial production and structural analysis has focused mainly on dextrans from Leuconostoc species, particularly on Leuconostoc mesenteroides strains. In this study, we used NMR spectroscopy techniques to analyze the structures of dextrans produced by Leuconostoc citreum E497 and Weissella confusa E392. The dextrans were compared to that of L. mesenteroides B512F produced under the same conditions. Generally, W. confusa E392 showed better growth and produced more EPS than did L. citreum E497 and L. mesenteroides B512F. Both L. citreum E497 and W. confusa E392 produced a class 1 dextran. Dextran from L. citreum E497 contained about 11% alpha-(1-->2) and about 3.5% alpha-(1-->3)-linked branches whereas dextran from W. confusa E392 was linear with only a few (2.7%) alpha-(1-->3)-linked branches. Dextran from W. confusa E392 was found to be more linear than that of L. mesenteroides B512F, which, according to the present study, contained about 4.1% alpha-(1-->3)-linked branches. Functionality, whether physiological or technological, depends on the structure of the polysaccharide. Dextran from L. citreum E497 may be useful as a source of prebiotic gluco-oligosaccharides with alpha-(1-->2)-linked branches, whereas W. confusa E392 could be a suitable alternative to widely used L. mesenteroides B512F in the production of linear dextran.  相似文献   

13.
A debranching enzyme purified from germinating rice endosperm hydrolyzed oligosaccharides having maltosyl or maltotriosyl branches (B4-B6) moderately. Hydrolysis of maltosylmaltose by a “pullulanase” of higher plant origin has been scarcely reported, while our enzyme debranched maltosylmaltose like microbial pullulanase. Additionally, the enzyme slowly hydrolyzed isopanose to glucose and maltose.

Gel-filtration analyses of hydrolysis products of polysaccharides with the enzyme suggested that while it hydrolyzed α-1,6-linkages of pullulan at random, it hydrolyzed amylopectin and glycogen at the outer α-1,6-linkages preferentially In the hydrolysis products of glycogen with the enzyme for a longer incubation time, large molecular-weight glucans still remained. This indicated that the enzyme was able to hydrolyze a few of the α-1,6-linkages of glycogen.  相似文献   

14.
An isomaltotriose-producing dextranase II, detected in the culture supernatant of Flavobacterium sp. M-73, was purified to an electrophoretically pure state. Successive chromatography on hydrophobic columns of Amberlite CG-50 and aminooctyl-Sepharose was very effective as the first step of purification. Further purification of the enzyme was performed by affinity column chromatography on isomaltotriose-Sepharose and preparative polyacrylamide gel electrophoresis.

The purified enzyme was shown to be a monomer and had a molecular weight of 114,000. Dextranase II was most active at pH 7.0 and 35°C. It was stable at 4°C for 24 hr over a pH range of 6.5~12.0 and up to 35°C on heating for 10 min. This enzyme had a strict specificity for consecutive α-l,6-glucosidic linkages and readily hydrolyzed clinical dextran and Sephadex gels. The degree of hydrolysis of clinical dextran was 31% expressed as apparent conversion into D-glucose. The amount of isomaltotriose in the hydrolyzate was determined to be 63%.  相似文献   

15.
Dextran is widely exploited in medical products and as a component of drug-delivering nanoparticles (NPs). Here, we tested whether dextran can serve as the main substrate of NPs and form a stable backbone. We tested dextrans with several molecular masses under several synthesis conditions to optimize NP stability. The analysis of the obtained nanoparticles showed that dextran NPs that were synthesized from 70 kDa dextran with a 5% degree of oxidation of the polysaccharide chain and 50% substitution with dodecylamine formed a NP backbone composed of modified dextran subunits, the mean diameter of which in an aqueous environment was around 100 nm. Dextran NPs could be stored in a dry state and reassembled in water. Moreover, we found that different chemical moieties (e.g., drugs such as doxorubicin) can be attached to the dextran NPs via a pH-dependent bond that allows release of the drug with lowering pH. We conclude that dextran NPs are a promising nano drug carrier.  相似文献   

16.
Dextran fractions from NRRL strain Streptococcus sp. B-1526 and the native, structurally homogeneous dextrans from Acetobacter capsulatum B-1225, Leuconostoc mesenteroides B-1307, and L. dextranicum B-1420 were examined by 13C-n.m.r. spectroscopy at 90°. Dextran B-1526 fraction I and dextran B-1420 were also examined by g.l.c:-m.s., methylation-structural analysis. All of these dextrans and dextran fractions branch, either primarily or exclusively, through α-d-(1→4)-glucopyranosyl linkages; however, their degrees of branching differ. Several 13C-n.m.r. resonances that are diagnostic for 4,6-di-O-substituted α-d-glucopyranosyl residues have been identified. Comparison was made with dextrans from L. mesenteroides B-742 fraction L and Streptobacterium dextranicum B-1254 fraction S[L], for which previously published, methylation-structural analyses had established the presence of 4,6-di-O-substituted α-d-glucopyranosyl residues at the branch points. These fermentation culture, and in a sedimented gum-phase (fraction I). The product from the soluble phase is designated here as fraction S in order to simplify the terminology. Originally7, this product was not designated a fraction, because it was, by definition8, the main dextran product. The same distinction also applies to the pairs of products from strains B-1380, B-1420, and b-1394 (see ref. 7). The attempts thus made to establish the significance of the phase separation were indeterminant.Methods.— Methods previously described were used for the mythylation9 of the dextrans and for structural analysis6.38 by combined g.l.c-electron-impact mass spectrometry of the aldononitriles. For each permethylation, three successive Hakomori39 methylations were employed on an initial, 40-mg sample, with ~80% (final weight) recovery of each permethylated dextran. Successive formolysis and acetic acid hydrolysis were employed, and, after each step, the resulting solutions were clear, colorless, and free from suspended material. All mass spectra were recorded with a Hewlett-Packard 5980A GC/MS integrated g.l.c.-m.s.-computer system. The g.l.c. peak-integrals reported in Table II were obtained with a Barber-Coleman Series 5000 g.l.c. instrument equipped with hydrogen-flame detectors. On-column injection with glass columns (2mmi.d. x 1.23m) was employed for all chromatograhy.The 13C-n.m.r. conditions and the methods for the preparation of dextran samples have been described4. In general, a Varian XL-100-15 spectrometer equipped with a Nicolet TT-100 system was employed in the Fourier-transform mode. The dextran samples, ~0.3g/4 mL of deuterium oxide, were maintained at 90°. Chemical shifts are expressed in p.p.m. relative to external tetramethylsilane, but were actually calculated by reference to the solvent lock-signal. The convolution-difference resolution-enhancement (c.d.r.e.) technique has been described40.  相似文献   

17.
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   

18.
The general properties and specificity of a dextran α-(1→2)-debranching enzyme from Flavobacterium have been examined in order to apply this enzyme to the structural analysis of highly branched dextrans. The optimum pH range and temperature were pH 5.5–6.5, and 45°, respectively. The enzyme was stable up to 40° on heating for 10 min, and over a pH range of 6.5–9.0 on incubation at 4° for 24 h. The effects of various metal ions and chemical reagents have also been examined. The debranching enzyme has a strict specificity for the (1→2)-α-d-glucosidic linkage at branch points of dextrans and related branched oligosaccharides, and produces d-glucose as the only reducing sugar. The degree of hydrolysis of the dextrans by this enzyme and the Km value (mg/mL) were as follows: B-1298 soluble, 25.2%, 0.21; B-1299 soluble, 31.5%, 0.27; and B-1397, 11.8%, 0.91. The debranching enzyme thus has a novel type of specificity as a dextranhydrolase. We have termed this enzyme as dextran α-(1→2)-debranching enzyme, and its systematic name is also discussed.  相似文献   

19.
Fragmentation of a dextran from Leuconostoc mesenteroides B* by controlled acetolysis gave five trisaccharides; isomaltotriose (0.12%), 3-α-isomaltosylglucose (0.80%), 6-α-nigero-sylglucose (0.43%), 3,6-di-α-glucosylglucose (0.18%) and nigerotriose (0.12%). Among these trisaccharides the last three were first isolated from dextran. Isolation of nigerotriose indicated that some of the 1,3-linkages in the dextran molecule were contiguous.  相似文献   

20.
Dextransucrase from Leuconostoc mesenteroides was produced in a semicontinuous culture with slow addition of a concentrated sucrose solution. The resulting high activity of the fermentation broth allowed a one-step purification method, by gel permeation chromatography (GPC) in 96.4% yield. This procedure resulted in 140-fold purification, with specific activity of 122 U/mg. The enzyme was immobilized onto an amino-Spherosil support activated with glutaraldehyde. Preparations with dextransucrase activities as high as 40.5 U/g of support were obtained, when low specific area supports were used and maltose was added during the enzyme coupling. Diffusional limitations were found during enzyme reaction, as shown by a kinetic study. As a consequence of immobilization, the average molecular weight of dextrans seems to increase. Immobilized dextransucrase looks promising for low-molecular-weight dextran production. Clinical dextran was synthesized when the polysaccharides produced in the presence of maltose were used as acceptor of a second synthesis reaction. The molecular weight distribution of the resulting production was less disperse than when clinical dextran was produced by acid hydrolysis of high-molecular-weight dextran.  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