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.
《Carbohydrate research》1985,138(2):207-213
The non-Newtonian behavior and dynamic viscoelasticity of a series of aqueous mixtures of xanthan and guar gum were measured with a rheogoniometer. At a concentration of 0.2% of total gums, gelation did not occur at room temperature but occurred at a low temperature (0°). A much stronger interaction was observed with a mixture of deacetylated xanthan than that with native xanthan. The maximum dynamic modulus was obtained when the ratio of xanthan to guar gum was 2:1. The transition temperatures of dynamic viscoelasticity for mixtures with native and deacetylated xanthan were observed at 25 and 30°, respectively. It was concluded that the side chains of the guar gum molecular prevent an intermolecular interaction with the side chains of the xanthan molecule. An intermolecular interaction between xanthan and guar gum at low temperature might be promoted between the periphery of the side chains of the xanthan molecule and the backbone of the guar gum molecule and dissociation takes place at the transition temperature.  相似文献   

2.
Polarization microscopy has been used to investigate the structure of 50/50 xanthan/galactomannan (guar gum or locust bean gum) mixtures in aqueous solution, the total concentration ranging from 0.5 to 4%. By the use of polarized light microscopy birefringent areas resulting from the formation of cholesteric mesophases in xanthan gum was clearly seen as has previously been reported by several authors. In xanthan/galactomannan mixtures, we also observed birefringent areas. Moreover, these zones in the blend appeared more anisotropic than with xanthan gum alone. This suggests that xanthan molecules organize themselves as liquid crystalline mesophases in definite enriched xanthan areas resulting from a concentration of xanthan inside these birefringent zones. Upon heating, this anisotropy disappears at a temperature well below the helix-coil transition temperature of xanthan molecules. In fact, this loss of order of the mixed system occurs at the same temperature as the melting temperature of the gel, as assessed by the use of rheological measurements. Since the ordered helical structure of the xanthan molecules still exists beyond the melting temperature while anisotropy disappears, this suggests that the xanthan molecules are no longer concentrated in specific areas but more evenly distributed in the medium. Gel melting would, therefore, be the result of the disappearance of these xanthan enriched areas.  相似文献   

3.
《Carbohydrate research》1986,147(2):275-294
A range of galactomannans varying widely in the contents of d-galactose have been compared for self-association and their interaction properties with agarose and xanthan. Whereas, in general, the most interactive galactomannans are those in which the (1→4)-β-d-mannan chain is least substituted by α-d-galactosyl stubs, evidence is presented which indicates that the distribution of d-galactosyl groups along the backbone (fine structure) can have a significant effect on the interaction properties. For galactomannans containing <30% of d-galactose, those which contain a higher frequency of unsubstituted blocks of intermediate length in the β-d-mannan chain are most interactive. For galactomannans containing >40% of d-galactose, those which contain a higher frequency of exactly alternating regions in the β-d-mannan chain are most interactive. This selectivity, on the basis of galactomannan fine-structure, in mixed polysaccharide interactions in vitro could mimic the selectivity of binding of branched plant-cell-wall polysaccharides in biological systems.  相似文献   

4.
《Carbohydrate polymers》1987,7(3):183-224
This paper reviews the extensive and previously unpublished work on the interactions between agarose and 1,4-linked β-d-glycans carried out at Unilever Research, Colworth Laboratory, UK. The effect of the following variables is discussed: (i) galactose content of galactomannans; (ii) substitution patterns in the agarose molecule; (iii) structural variations in the 1,4-β-d-glycan main chain; and (iv) molecular size of the 1,4-β-d-glycans.Double helices of agarose, a non-substituted regular polysaccharide comprising 1,3-linked β-d-galactose and 1,4-linked 3,6-anhydro-α-l-galactose, bind in an ordered cooperative fashion to an extended ribbon ordered conformation of sequences of 1,4-linked β-d-mannopyranose residues in plant galactomannans to give mixed gelling systems. This interaction survives, in a modified form, substitution along the agarose molecule by O-methyl ether and O-sulphate esters at O6 of the d-galactose and O2 of the 3,6-anhydro-l-galactose, and 4,6-linked pyruvic acid ketal groups on the d-galactose. The higher the level of substitution on the agarose, the weaker the interaction with galactomannan.In general, the higher the level of galactose substitution in the galactomannan the lower the extent of interaction with agarose. Evidence is presented, however, which indicates that the fine structural distribution of galactose along the galactomannan molecule is also an important determinant for the co-gelling interaction. Substituted 1,4-linked β-d-glucomannans, β-d-glucans and β-d-xylans which can form closely similar extended ribbon order conformations to the galactomannans also participate in co-gelling interactions with agarose. These β-d-glycans are similar in structure to important skeletal polysaccharides such as hemicelluloses and cellulose. This suggests that the binding between agars and β-d-glycans might mimic biological cohesion between skeleton and gel phases in natural red seaweed cell walls. The sensitivity of the interactions studied to fine details of agar and β-d-glycan structure is what might be expected on biological grounds, since the wide and subtle variations of natural polysaccharide structure are presumed to represent a mechanism for control of their intermolecular interactions.  相似文献   

5.
The influence of xanthan conformation on the physicochemical behaviour of their mixtures with galactomannan from Schizolobium parahybae mannose:galactose ratio (M/G=3), was studied by viscoelastic measurements, differential scanning calorimetry (DSC) and chiroptical (circular dichroism) methods. The results suggested a more effective interaction of the galactomannan with disordered xanthan segments, which are more abundant in low salt concentrations but are still present in lower proportion at temperatures lower than the temperature of xanthan conformational transition (Tm). The dependence of ellipticity with temperature in a circular dichroism (CD) spectra suggested an ordering of the xanthan chains induced by galactomannan at the temperature of gel formation (Tg≈25°C), under conditions where xanthan alone exhibits a disordered conformation. The lower Tg value found (≈25°C) compared with that (60°C) usually described in the literature is certainly related to the M/G ratio and the galactosyl unit distribution along the mannan main chain.  相似文献   

6.
The influence of the galactomannan characteristic ratios (M/G) on the temperature of gelation (Tg) and the gel strength of mixtures of galactomannan with xanthan is reported. Two galactomannans were investigated: one highly substituted from the seeds of Mimosa scabrella (M/G = 11), and the other, less substituted, from the endosperm of Schizolobium parahybae, with (M/G = 30) [Ganter JLMS, Zawadzki-Baggio SF, Leitner SC, Sierakowski MR, Reicher F. J Carbohydr Chem 1993;12:753]. The xanthan:galactomannan systems (4:2 g l(-1), in 5 mM NaCl) showed a temperature of gel formation (Tg) of 24 degrees C for that of S. parahybae [Bresolin TMB, Milas M, Rinaudo M and Ganter JLMS. Int J Biol Macromol 1998;23:263] and 20 degrees C for the galactomannan of M. scabrella, determined by viscoelastic measurements and microcalorimetry. A Tg of 40-50 degrees C was found by Shatwell et al. [Shatwell KP, Sutherland IW, Ross-Murphy SB, Dea ICM. Carbohydr Polym 1991;14:29] for locust bean gum-LBG (M/G = 43). Lundin and Hermansson [Lundin L, Hermansson AM. Carbohydr Polym 1995;26:129] reported a difference of 13 degrees C for Tg of two LBG samples with M/G = 3 (40 degrees C) and 5 (53 degrees C), in mixtures with xanthan. It appears that the more substituted galactomannans have lower temperatures of gelation in the presence of xanthan. The mechanism of gelation depends also on the M/G ratio. For the lower values it involves only disordered xanthan chains in contrast to M/G ratios higher than 3. In addition, the presence of the galactomannan from M. scabrella increased slightly the temperature of the conformational change (Tm) of xanthan probably due to the ionic strength contribution of proteins (3.9%) present in the galactomannan. On the other hand, the galactomannans from S. parahybae, with 1.5% of proteins and M. scabrella, with 2.4% of protein, did not show this effect, the Tm of xanthan alone or in a mixture being practically unchanged.  相似文献   

7.
This study describes the effects of mixtures of xanthan gum and galactomannan, guar gum, or locust bean gum, on the lipids in plasma and liver in non-diabetic and diabetic rats. Non-diabetic rats were fed cholesterol-free diets with 3% guar gum, locust bean gum, or xanthan gum (3G, 3L, and 3X), or a mixture of xanthan gum and guar gum or locust bean gum (1:2, w/w) (2G1X, 2L1X) for 2 weeks. Rats fed diets not containing these polysaccharides were used as controls. The total cholesterol in plasma and the triacylglycerol in liver were significantly lowered in rats fed the 2G1X diet. The 3G, 3X, 3L, and 2L1X diets showed no significant effect on the total cholesterol and triacylglycerol in plasma and liver. In the streptozotocin-induced (STZ) diabetic rats, the total cholesterol in plasma was lowered in rats fed the 3G, 3X or 2G1X diet for 4 weeks, and the 2G1X diet was more effective than the 3G and 3X diets. The triacylglycerol in plasma in STZ diabetic rats was also significantly lowered by the 2G1X diet. These results showed that a mixture of xanthan gum and guar gum has an improved hypolipidemic effect on non-diabetic and STZ diabetic rats. The effects of the 2G1X diet on the diabetic symptoms in STZ diabetic rats, suppression of food and water intakes, decrease in glucose in urine, and lowering of plasma glucose, were also observed.  相似文献   

8.
This study describes the effects of mixtures of xanthan gum and galactomannan, guar gum, or locust bean gum, on the lipids in plasma and liver in non-diabetic and diabetic rats. Non-diabetic rats were fed cholesterol-free diets with 3% guar gum, locust bean gum, or xanthan gum (3G, 3L, and 3X), or a mixture of xanthan gum and guar gum or locust bean gum (1:2, w/w) (2G1X, 2L1X) for 2 weeks. Rats fed diets not containing these polysaccharides were used as controls. The total cholesterol in plasma and the triacylglycerol in liver were significantly lowered in rats fed the 2G1X diet. The 3G, 3X, 3L, and 2L1X diets showed no significant effect on the total cholesterol and triacylglycerol in plasma and liver. In the streptozotocin-induced (STZ) diabetic rats, the total cholesterol in plasma was lowered in rats fed the 3G, 3X or 2G1X diet for 4 weeks, and the 2G1X diet was more effective than the 3G and 3X diets. The triacylglycerol in plasma in STZ diabetic rats was also significantly lowered by the 2G1X diet. These results showed that a mixture of xanthan gum and guar gum has an improved hypolipidemic effect on non-diabetic and STZ diabetic rats. The effects of the 2G1X diet on the diabetic symptoms in STZ diabetic rats, suppression of food and water intakes, decrease in glucose in urine, and lowering of plasma glucose, were also observed.  相似文献   

9.
Galactomannan (G) from Gleditsia sinensis Lam. and xanthan gum (X), as sustained release materials for controlled delivery of theophylline, were mixed in different ratios of 7:3, 5:5, and 3:7 to yield enhanced release-controlling performance. The polysaccharides content of tablets was 10% (w/w), either alone or in mixtures. From in vitro dissolution test, G10% and X10% matrices released 91.4 and 87.7% of drug within 24 h, respectively. The synergistic interactions between galactomannan and xanthan effectively retarded the drug diffusion, and the most sustained drug release (75.5% at 24 h) was found in formulation GX7:3. The drug release data fitted to the kinetic model indicated the anomalous transport mechanism (0.5 < n < 1.0). Additionally, the swelling behavior and morphological changes of the tablets were investigated. The results illustrated the potential of binary mixtures of G. sinensis galactomannan and xanthan as novel sustained release materials for controlled drug delivery.  相似文献   

10.
Food Biophysics - Molecular dynamics for pastes of two normal and one waxy potato starches and their binary mixtures with either arabic, guar or xanthan gum was determined. Spin–lattice and...  相似文献   

11.
Guar gum, a beta-D-(1-->4)-linked D-mannan with alpha-D-galactopyranosyl units attached as side groups, was treated with alpha-galactosidase, an enzyme that splits off the alpha-D-galactosyl units to obtain a galactomannan with a low galactose content. The galactose-depleted polysaccharide was then selectively oxidized in C(6) position and epimerized using mannuronan C(5)-epimerases, namely AlgE1, AlgE4, AlgE6, and their mixtures, obtaining new pseudo-alginates. In this paper, we report a full high field 1D and 2D NMR study of guar gum as such and of the galactose-depleted, oxidized and epimerized compounds, respectively. From the 1H NMR spectra, the degree of epimerization, the distribution of mannuronic acid (M) and guluronic acid (G) residues and the average G-block length, N(G>1), were obtained. By means of NMR diffusion experiments, it was also shown that no significant degradation of the polysaccharide occurs as a consequence of the epimerization reactions.  相似文献   

12.
The gelation of agarose is investigated by rheological methods and electron microscopy, as well as the thickening properties of xanthan. The gelling and thickening agents have been investigated in pure water to compare the results with theoretical models. The gelation of agarose was shown to follow two steps upon cooling, which could be addressed to the formation of helices and their aggregation. In addition to the rheology, transmission electron micrographs of freeze-dried samples have been taken to underline the date by corresponding structures at different stages of the gelling process. The xanthan molecules, which have been approximated by rigid highly charged rodlike molecules, undergo a jamming transition at a critical concentration. This concentration shows a strong dependence on the length of the molecules, which supports the high thickening effect of xanthan. When both, agarose and xanthan are mixed, the gel structure becomes very different. The gelling process is now determined by one step only. It is proposed that the jamming xanthan molecules prevent the formation of the aggregates of the agarose gel. The gels themselves appear then less elastic, and should yield a better mouth feeling.  相似文献   

13.
Galactose and mannose, released on hydrolysis of galactomannan in the endosperm of germinating seeds of carob, guar, honey locust and lucerne were absorbed by the cotyledons and further metabolized. In guar, the distribution of 14C from [U-14C]-d-glucose, d-mannose and D-galactose into various cotyledon fractions did not provide evidence for preferential channelling of d-galactose into cell wall fractions and d-mannose into glycolysis. Phosphomannoisomerase, which has previously been reported in animals and microorganisms was detected in a number of legume seeds. In honey locust it was located in the cotyledons and its level declined after galactomannan was depleted. This enzyme from lucerne was purified until free of phosphoglucoisomerase and some of its properties are described.  相似文献   

14.
The effect of acidification on a typical commercial xanthan and on pyruvate-free xanthan (PFX), alone and in gelling mixtures with konjac glucomannan (KGM), has been studied by differential scanning calorimetry (DSC) and small-deformation oscillatory measurements of storage modulus (G′) and loss modulus (G″). For both xanthan samples, progressive reduction in pH caused a progressive increase in temperature of the disorder–order transition in DSC, and a progressive reduction in gelation temperature with KGM. This inverse correlation is interpreted as showing that synergistic gelation involves disruption of the xanthan 5-fold helix, probably by attachment of KGM to the cellulosic backbone of the xanthan molecule (as proposed previously by a research group in the Institute of Food Research, Norwich, UK). Higher transition temperature accompanied by lower gelation temperature for PFX in comparison with commercial xanthan at neutral pH is explained in the same way. However, an additional postulate from the Norwich group, that attachment of KGM (or galactomannans) can occur only when the xanthan molecule is disordered, is inconsistent with the observation that gelation of acidified mixtures of KGM with PFX can occur at temperatures more than 60 °C below completion of conformational ordering of the PFX component (as characterised by DSC). Increase in G′ on cooling for mixtures of commercial xanthan with KGM at pH values of 4.5 and 4.25 occurred in two discrete steps, the first following the temperature-course observed for the same mixtures at neutral pH and the second occurring over the lower temperatures observed for mixtures of KGM with PFX at the same values of pH. These two “waves” of gel formation are attributed to interaction of KGM with, respectively, xanthan sequences that had retained a high content of pyruvate substituents, and sequences depleted in pyruvate by acid hydrolysis. At pH values of 4.0 and lower, gelation of mixtures of KGM with commercial xanthan followed essentially the same temperature-course as for mixtures with PFX, indicating extensive loss of pyruvate under these more strongly acidic conditions. Mixtures prepared at pH values in the range 4.0–3.5 gave comparable moduli at room temperature (20 °C) to those obtained at neutral pH, but showed substantial softening on heating to body temperature, suggesting possible applications in replacement of gelatin in products where “melt-in-the-mouth” characteristics are important for acceptability to the consumer.  相似文献   

15.
The purpose of this work was to investigate the viscoelastic properties of aqueous suspensions of crude collagen powder extracted from bovine hides and nonsubmitted to the hydrolysis reaction that leads to gelatin. The studied variables included the collagen concentration and the addition of xanthan gum or maltodextrin at varied concentrations during heating/cooling of the mixtures. Differential scanning calorimetry thermograms showed that the addition of polysaccharides decreased the endothermic peak areas observed at the denaturation temperature of collagen. The rheological properties of the pure collagen suspensions were highly dependent on concentration: 4% and 6% collagen suspensions presented a great increase in the storage modulus after heating/cooling, whereas for concentrations of 8% and 10% G′ decreased during heating and did not recover its original value after heating/cooling. The frequency sweeps showed that the thermal treatment was responsible by the strengthening of the interactions that formed the polymer network. Addition of 0.1% xanthan gum to collagen suspensions increased the gel strength, especially after heating/cooling of the system, whereas increasing gum concentration to 0.3% resulted in a weaker gel, which could indicate thermodynamic incompatibility between the biopolymers. Mixtures of collagen and maltodextrin resulted in more fluid structures than those obtained with pure collagen at the same collagen concentration and the range of temperatures in which these mixtures behaved as a gel decreased with increasing concentrations of both collagen and maltodextrin, suggesting incompatibilities between the biopolymers.  相似文献   

16.
The conformational properties of xanthans with partially hydrolyzed side chains were in vestigated by optical rotation, CD, and differential scanning calorimetry (DSC). All variants displayed the well-known temperature-driven, cooperative order–disorder transition, and both optical rotation and DSC showed that the transition temperature was essentially independent of the content of terminal β-mannose. It was found that up to 80% of the changes in the specific optical rotation accompanying the transition reflects conformational changes linked to the terminal β-mannose in the side chains. Modification of the sidechains also affected the CD when xanthan was in the ordered state, but in this case the data suggest that the glucuronic acid is the major component determining the magnitude of the CD signal. DSC measurements showed that the transition enthalpy (ΔHcal) increased linearly with the fraction of β-mannose, again indicating that a significant part (up to 80%) of ΔHcal reflects conformational changes in the side chains. The conformational transition of the xanthan variants generally showed a higher degree of cooperativity (sharper transition) than unmodified, pyruvated xanthan. Calculation of the cooperativity parameter σ by means of the Zimm–Bragg theory (OR data) or from the ratio between ΔHcal and the van't Hoff enthalpy (ΔHvH) using DSC data showed a correlation between σ and the content of β-mannose, but the two methods gave different results when the content of β-mannose approached 100%. The ionic strength dependence of the transition temperature, expressed as d (log I)/d(T?1m), was nearly identical for intact xanthan and a sample containing only 6% of the terminal β-mannose. Application of the Manning polyelectrolyte theory does not readily account for the observed ΔHcal values, neither does it provide new information on the nature of the ordered and disordered conformations in xanthan. © 1993 John Wiley & Sons, Inc.  相似文献   

17.
This study concerns the efficacy of partial agar substitution by galactomannans as support in plant regeneration media for Nicotiana tabacum. The production of multiple shoots from leaf-derived callus and their rooting were evaluated. The galactomannans applied were obtained from Cassia fastuosa (cassia) and Cyamopsis tetragonolobus (guar gum — a commercial galactomannan) seeds. The results obtained on media solidified with mixtures of agar/galactomannan (3 g dm−3 each) gels were compared with those on media gelled with a standard concentration of agar (6 g dm−3). The in vitro performance allowed to conclude that the use of galactomannans raised the number of shoots and improved their quality. Furthermore, the length of roots and the size of leaves were significantly higher in the media solidified with agar/guar galactomannan mixtures.  相似文献   

18.
This study investigated the behavior of mannan-degrading enzymes, specifically focusing on differences with respect to their substrate specificities and their synergistic associations with enzymes from different glycoside hydrolase (GH) families. Galactosidases from Cyamopsis tetragonolobus seeds (Aga27A, GH27) and Aspergillus niger (AglC, GH36) were evaluated for their abilities to synergistically interact with mannanases from Clostridium cellulovorans (ManA, GH5) and A. niger (Man26A, GH26) in hydrolysis of guar gum and locust bean gum. Among the mannanases, Man26A was more efficient at hydrolyzing both galactomannan substrates, while among the galactosidases; Aga27A was the most effective at removing galactose substituents on both galactomannan substrates and galactose-containing oligosaccharides. An optimal protein mass ratio of glycoside hydrolases required to maximize the release of both reducing sugar and galactose residues was determined. Clear synergistic enhancement of locust bean gum hydrolysis with respect to reducing sugar release was observed when both mannanases at 75% enzyme dosage were supplemented with 25% enzyme protein dosage of Aga27A. At a protein ratio of 75% Man26A to 25% Aga27A, the presence of Man26A significantly enhanced galactose release by 25% Aga27A (2.36 fold) with locust bean gum, compared to when Aga27A was used alone at 100% enzyme protein dosage. A dosage of Aga27A at 75% and ManA at 25% protein content liberated the highest reducing sugar release on guar gum hydrolysis. A dosage of Man26A and Aga27A at 75–25% protein content, respectively, liberated reducing sugar release equivalent to that when Man26A was used alone at 100% protein content. From the findings obtained in this study, it was observed that the GH family classification of an enzyme affects its substrate specificity and synergistic interactions with other glycoside hydrolases from different families (more so than its EC classification). The GH26 Man26A and GH27 Aga27A enzymes appeared to be more promising for applications that involve the hydrolysis of galactomannan containing biomass. This method of screening for maximal compatibility between various GH families can ultimately lead to a more rational development of tailored enzyme cocktails for lignocellulose hydrolysis.  相似文献   

19.
The effects on amylose conformation of percentage water in dimethyl sulfoxide (DMSO)/water mixtures were measured by following changes in specific optical rotation, limiting viscosity number, and 13C-NMR chemical shifts. The temperature dependence of specific optical rotation showed differences in amylose conformation at four chosen ratios of dimethyl sulfoxide/water. An amylose conformational change was also deduced from 13C-NMR chemical shift data. Changes in limiting viscosity of amylose in different proportions of DMSO/water, and the effect of tetramethylurea on the specific rotation of amylose, indicate that intramolecular hydrogen bonding decreases with increased water content. 66.6% DMSO appears to be a crossover concentration, below which the helical conformation is progressively lost as water is added. When water content is over 60%, transition to a conformation which allows iodine complexation to take place is complete. A transition of amylose conformation from helix to loose helix to random coil with increasing water content was deduced from the experimental results.  相似文献   

20.
The objectives were to characterize propranolol hydrochloride-loaded matrix tablets using guar gum, xanthan gum, and hydroxypropylmethylcellulose (HPMC) as rate-retarding polymers. Tablets were prepared by wet granulation using these polymers alone and in combination, and physical properties of the granules and tablets were studied. Drug release was evaluated in simulated gastric and intestinal media. Rugged tablets with appropriate physical properties were obtained. Empirical and semi-empirical models were fit to release data to elucidate release mechanisms. Guar gum alone was unable to control drug release until a 1:3 drug/gum ratio, where the release pattern matched a Higuchi profile. Matrix tablets incorporating HPMC provided near zero-order release over 12 h and erosion was a contributing mechanism. Combinations of HPMC with guar or xanthan gum resulted in a Higuchi release profile, revealing the dominance of the high viscosity gel formed by HPMC. As the single rate-retarding polymer, xanthan gum retarded release over 24 h and the Higuchi model best fit the data. When mixed with guar gum, at 10% or 20% xanthan levels, xanthan gum was unable to control release. However, tablets containing 30% guar gum and 30% xanthan gum behaved as if xanthan gum was the sole rate-retarding gum and drug was released by Fickian diffusion. Release profiles from certain tablets match 12-h literature profiles and the 24-h profile of Inderal® LA. The results confirm that guar gum, xanthan gum, and HPMC can be used for the successful preparation of sustained release oral propranolol hydrochoride tablets.  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