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 31 毫秒
1.
Feruloylated arabinoxylans isolated from wheat flour and wheat bran were compared in their cross-linking behaviour with respect to viscosity properties and cross-linking products formed when various oxidative agents were applied to dilute solutions. Optimal conditions for each oxidative agent were investigated. In case of hydrogen peroxide and peroxidase, similar conditions were found for both types of arabinoxylans but wheat bran arabinoxylans gave a larger viscosity increase upon cross-linking than those of wheat flour.

When glucose, glucoseoxidase and peroxidase or ammonium persulphate were used as oxidative agents, differences in the concentration of reagent needed to induce cross-linking and in viscosity increase were observed. The distribution of coupling products for both types of arabinoxylans and the different oxidative treatments was approximately 5 : 3 : 1 : 1 for 8-5, 8-O-4, 8-8 and 5-5, respectively. The low ferulate recovery after oxidative treatment was assumed to be caused by formation of unknown compounds, such as higher oligomers and lignin-linked products.

A 1 : 1 mixture of flour arabinoxylan and feruloylated pectin showed a maximum synergistic effect on viscosity upon oxidative treatment using hydrogen peroxide and peroxidase. Both polysaccharides were shown to participate in cross-linking.  相似文献   


2.
Creeping fig (Ficus pumila Linn.) seeds can be manually rubbed and squeezed to produce a water extract (WE) that can gel at room temperature without any additives. Its gelling material, extraction behavior, and mechanism of spontaneous gel-formation were investigated. Gelling material was extracted from seeds using water, ammonium oxalate and hydrochloric acid, respectively. Results showed the main component of three extracts is low methoxyl pectin. The pectin locates in a transparent layer on the surface of seeds, as revealed by an inverted microscope. Hence, explained the feasibility of the squeezing and rubbing method in traditional handcraft. Comparing with the other methods, water-extracted pectin has high galacturonic acid content, viscosity-average molecular weight and intrinsic viscosity but low degree of methoxylation. This pectin forms the major component of WE, together with the high amount of calcium ions present in WE, it suggests the spontaneous gelation may be based on the ‘egg-box’ formation.  相似文献   

3.
A theoretical discussion of the extraction process and especially of the factors influencing extractant penetration in pectin-containing raw materials is presented. The necessity of adding surfactant in the pectin extraction is proved. Experiments were carried out with two types of apple pressings differing in anhydrouronic acid content and quantity of extractable pectin.

The addition of low molecular alcohols in concentrations from 1% to 3% to the acid extragent resulted in an acceleration of extraction and increase in the pectin yield by 55–90%. Ethylene glycol, glycerol and diethylene glycol had a better effect than monohydric alcohols. The effect of the process duration on the pectin yield during acid extraction was studied. A 25-min extraction was sufficient for taking out the extractable pectin. It was shown that the addition of alcohols resulted in a measurable increase in the pectin gel strength.  相似文献   


4.
Pectins extracted from Krueo Ma Noy (Cissampelos pareira) leaves mainly consisted of galacturonic acid with trace amount of neutral sugars. The dominant structure of Krueo Ma Noy pectin was established as a 1,4-linked -D-galacturonan by a combination of carboxyl reduction and methylation analysis, and confirmed by FT-IR spectroscopy. The degree of esterification of Krueo Ma Noy pectins was 41.7 and 33.7% for crude and dialyzed pectins, respectively. Krueo Ma Noy pectin has an average molecular weight of 55 kDa, radius of gyration of 15.2 nm and intrinsic viscosity of 2.3 dl/g. Krueo Ma Noy pectin exhibited gelling properties in aqueous solutions at 0.5% (w/v) at 5 °C. Gels were formed at concentrations of 1.0% (w/v) and above even at room temperature. The gel strength, melting point, and melting enthalpy of Krueo Ma Noy pectin increased with polysaccharide concentration.  相似文献   

5.
Pectins were extracted from roots and petioles of sugar beet, and treated with alpha-arabinosidase, 1,4-beta-galactanase or polygalacturonase. They were then cross-linked using hydrogen peroxide and peroxidase. The effects on pectin molecular size were monitored by size-exclusion chromatography and viscometry. A decrease in apparent molecular size was observed after alpha-arabinosidase and polygalacturonase treatment, and all three enzymes caused a decrease in viscosity. The pectins were then cross-linked using hydrogen peroxide and peroxidase, and the effects on dehydrodiferulate formation were monitored by HPLC. Pretreatment with polygalacturonase caused no significant change in subsequent dehydrodiferulate cross-linking, while pretreatment with alpha-arabinosidase caused a slight change in the ratios of the different dehydrodiferulates formed. Pretreatment with 1,4-beta-d-galactanase caused a more significant change in the ratios of the different dehydrodiferulates formed, and also greatly increased the overall recovery of total ferulates (monomers plus dehydrodiferulates), both in root pectin and petiole pectin. The possible effects of polysaccharide microstructure on the dimerisation and further polymerisation of pectin-linked ferulates are discussed.  相似文献   

6.
Currently, oligo[poly(ethylene glycol) fumarate] (OPF) hydrogels are being investigated as an injectable and biodegradable system for tissue engineering applications. In this study, cytotoxicity of each component of the OPF hydrogel formulation and the resulting cross-linked network was examined. Specifically, OPF synthesized with poly(ethylene glycol) (PEG) of different molecular weights (MW), the cross-linking agent [PEG-diacrylate (PEG-DA)], and the redox initiator pair [ammonium persulfate (APS) and ascorbic acid (AA)] were evaluated for cytotoxicity at 2 and 24 h using marrow stromal cells (MSCs) as model cells. The effect of leachable byproducts of OPF hydrogels on cytotoxicity was also investigated. Upon exposure to various concentrations of OPF for 2 h, greater than 50% of the MSCs were viable, regardless of OPF molecular weight or concentration in the media. After 24 h, the MSCs maintained more than 75% viability except for OPF concentrations higher than 25% (w/v). When examining the cross-linking agent, PEG-DA of higher MW (3400) demonstrated significantly higher viability compared to PEG-DA with MW 575 at all concentrations tested. Considering initiators, when MSCs were exposed to AA and APS, as well as the combination of AA and APS, higher viability was observed at lower concentrations. Once cross-linked, the leachable products from the OPF hydrogels had minimal adverse effects on the viability of MSCs (percentage of live cells was higher than 90% regardless of hydrogel types). The results suggest that, after optimization of cross-linking parameters, OPF-based hydrogels hold promise as novel injectable scaffolds or cell carriers in tissue engineering.  相似文献   

7.
The effects of temperature and concentration on the viscosity of orange peel pectin solutions were examined at five different temperatures between 20 and 60°C and five concentration levels between 2.5–20 kg/m3. The effects of temperature was described by an Arrhenius-type equation. The activation energy for viscous flow was in the range 19.53–27.16 kJ/mol, depending on the concentration. The effect of concentration was described by two types of equation, power-law and exponential. Equations were derived which describes the combined effects of temperature and concentration on the viscosity for two different models in the range of temperatures and concentrations studied. Orange peel pectin was extracted by using HCl (pH 2.5, 90°C, 90 min) ammonium oxalate (0.25%, pH 3.5, 75°C, 90 min) and EDTA (0.5%, 90°C, 90 min) extraction procedures. The best result was obtained with ammonium oxalate extraction in which the pectin content of the final product was 30.12%, although the efficiency among the procedures varied.The average molecular weight was measured by light scattering technique. Magnitudes of intrinsic viscosity and molecular weight of pectins obtained by extraction with HCl, ammonium oxalate and EDTA were 0.262, 0.281, 0.309 m3/kg and 84 500, 91 400, 102 800 kg/kgmol, respectively. The molecular weight dependence of the intrinsic viscosity of the orange peel pectin solutions was expressed by Mark–Houwink–Sakurada equation. The data were fitted to equation as ηi=2.34×10−5(Mw,ave)0.8224 which helps to evaluate the average molecular weight of pectin solutions from orange peel with a knowledge of their intrinsic viscosity.  相似文献   

8.
Pectins were extracted from roots, petioles and leaves of sugar beet, and cross-linked using hydrogen peroxide and peroxidase. The effects on dehydrodiferulate formation were monitored by HPLC and TLC. Dehydrodimers were formed in different proportions to those found in vivo. There was a net loss of around 50% of the phenolic groups (monomers plus dimers) during dimerisation. Gel filtration showed that root and petiole pectin, but not leaf pectin, increased in molecular weight during cross-linking. The effects of varying the cross-linking conditions were investigated, and it was found that hydrogen peroxide concentration was the most important factor in controlling both the type and amount of dehydrodiferulate formed.  相似文献   

9.
A novel hydrogel system based on oligo(poly(ethylene glycol) fumarate) (OPF) is currently being investigated as an injectable carrier for marrow stromal cells (MSCs) for orthopedic tissue engineering applications. This hydrogel is cross-linked using the redox radical initiators ammonium persulfate (APS) and ascorbic acid (AA). In this study, two different persulfate oxidizing agents (APS and sodium persulfate (NaPS)) with three reducing agents derived from ascorbic acid (AA, sodium ascorbate (Asc), and magnesium ascorbate-2-phosphate (Asc-2)) and their combinations were examined to determine the relationship between pH, exposure time, and cytotoxicity for rat MSCs. In addition, gelation times for specific combinations were determined using rheometry. pH and cell viability data after 2 h for combinations ranging from 10 to 500 mM in each reagent showed that there was a smaller pH change and a corresponding higher viability at lower concentrations, regardless of the reagents used. At 10 mM, there was less than a 1.5 unit drop in pH and greater than 90% viability for all initiator combinations examined. However, MSC viability was significantly reduced with concentrations of 100 mM and higher of the initiator combinations. At 100 mM, exposure to NaPS/Asc-2 resulted in significantly more live cells than exposure to APS/AA or NaPS/Asc, but at this concentration, NaPS/Asc-2 exhibited significantly longer OPF gelation onset times than APS/AA. At all combination concentrations, exposure time (10 min vs 2 h) did not significantly affect MSC viability. These data indicate that final pH and/or radical formation have a large impact on MSC viability and that multiple, intertwined testing procedures are required for identification of appropriate initiators for cell encapsulation applications.  相似文献   

10.
Although isoelectric focusing patterns in ultrathin layers of polyacrylamide gel are distorted by the presence of salts, such as ammonium persulfate, this reagent is commonly used to promote polymerization of the gel. The amount of ammonium persulfate can be reduced but this causes the formation of sloppy gels or incomplete polymerization. A method is described here in which an ultrathin polyacrylamide gel is formed on a polyester sheet using ammonium persulfate in the absence of ampholyte. After complete polymerization, the ammonium persulfate is washed out and ampholyte is allowed to diffuse into the gel. Subsequent isoelectric focusing is then free from distortion caused by the presence of ammonium persulfate.  相似文献   

11.
Water extracts of rind, essential oil and juice from oranges, also citrus pectin and citric acid promoted the formation of lesions when spores of Penicillium digitatum were placed in wounds 1·0 mm deep in flavedo of oranges; fructose, glucose and sucrose had little effect. Rind extracts were less effective in wounds 0·5 mm deep but orange juice and pectin still increased infection. None of the substances allowed the parasite to infect fruit through unwounded surfaces. Germination of spores in water increased as spore concentration decreased but was poor even at low concentrations. Almost all spores germinated in aqueous extracts of flavedo, albedo or whole rind, or in wounds on the surface of fruit. Fructose, glucose, sucrose and xylose were less effective but still caused over two-thirds of spores to germinate but only in the presence of phosphate buffer. Without buffer, germination was little different from that in water. Arabinose and galactose stimulated germination to a lesser extent but with the same phosphate effect. Carboxymethylcellulose and pectin did not affect germination. A variety of substances containing nitrogen increased germination but to different degrees, decreasing in the order, casamino acids, yeast extract, ammonium salts, nitrate. Thiamin and to a lesser extent biotin were also effective. Volatile substances from rind infected with P. digitatum stimulated spore germination and growth of germ tubes. The significance of these results is discussed in relation to infection.  相似文献   

12.
Insolubilizing studies of water-soluble synthetic polypeptides containing lysine residues were examined using organic aliphatic and aromatic cross-linking agents such as dialdehydes, diacyl chlorides and diactive ester, together with an enzyme tyrosinase, in water and simulated seawater systems. The cross-linking reaction was characterized by the viscosity and turbidity changes. Among the organic cross-linking agents used aliphatic glutaraldehyde and aromatic o-phthalaldehyde were the most effective. When excess organic cross-linking agents were added to the lysine polypeptide systems, the corresponding solid gels were formed. As a whole, the molecular weight of the samples, the amino acid compositions, the cross-linking agent used, the molar ratios between cross-linking agents and functional residues and system pH were found to have roles in the insolubilizing reaction and the gel formation. The cross-linking results obtained were compared with those of the polypeptide-tyrosinase systems, whose deep brownish red colour was decolorized by the addition of L-ascorbic acid.  相似文献   

13.
《Carbohydrate research》1987,163(1):15-27
The actions of ammonium persulfate on (feruloylated) sugar-beet pectins and ferulate have been studied by spectrophotometry, viscometry, 1H-n.m.r. spectroscopy, and gel-permeation chromatography. The reactions followed a pseudo-first-order law with respect to pectin and ferulate, whereas the order with respect to ammonium persulfate was unity for pectins and varied from 0.5 to > 2 for ferulate. The rate constants mainly varied with the pH of the reaction mixture and there was an optimum at 3.8–5.7 for the gelation of the pectins. The results ruled out a simple condensation process between two ferulates (or feruloyl residues linked to the pectins) and suggeste a free-radical polymerisation reaction.  相似文献   

14.
The mechanism of association of pectin by calcium ions was studied to elucidate the gelling process. The molecular weight and size were determined by light scattering measurements on samples of pectin demethylated in gradation (ELM-pectin) by pectinesterase from Aspergillus japonicus, acid demethylated pectin (CLM-pectin), and sodium polygalacturonic acid (PGA). The molecular size of ELM-pectin which was prepared from identical materials increased quantitatively as demethylation progressed. The molecular size of CLM-pectin and PGA was larger than ELM-pectin even though the methoxyl content was similar. This probably resulted from differences in molecular structure. When Ca2+ was added to ELM-pectin, as demethylation progressed, molecular weight increased due to cross-linking induced by Ca2 + ; however, the increase was small, when Ca2+ was added to CLM-pectin, molecular weight increased greatly; however, the molecular size was small, and a slight contraction of molecular was caused by cross-linking, Ca2+ addition to PGA resulted in enhancement of phenomena observed with CLM-pectin.  相似文献   

15.
An endo-polygalacturonase from culture extracts of Aspergillus japonicus was purified about 34-fold by ammonium sulfate fractionation, SE-Sephadex column chromatography and gel filtration. The purified enzyme was homogeneous on ultracentrifugation and disc electrophoresis. Using gel filtration a molecular weight of 35,500 was estimated for the enzyme. The enzyme rapidly reduced the viscosity of pectic acid and released reducing groups in a random manner, yielding a mixture of mono-, di- and trigalacturonic acids as end products. The pH optimum of the enzyme for viscosity-reducing activity was 4.5 with pectin and pectic acid as substrates, and that for releasing reducing groups was also 4.5 with various pectic substances. The purified enzyme was able to macerate various kinds of plant tissues by itself.  相似文献   

16.
The aim of this study was to modify pectin by covalent attachment of the water-insoluble ligand 4-aminothiophenol to its polymeric backbone. 4-Aminothiophenol is a ligand which is highly prone to oxidation. Therefore, this ligand allows oxidative cross-linking of pectin under mild oxidative conditions. Additionally, hydrophobization of pectin can be achieved by the mentioned modification which offers certain advantages over highly hydrophilic native pectins. 4-Aminothiophenol was covalently attached to pectin via amide bond formation between carboxylic moieties of pectin and the amino-group of 4-aminothiophenol. Two different pectin–4-aminothiophenol conjugates were synthesized and investigated regarding the amount of coupled ligand, rheological behavior under oxidative conditions, swelling behavior, and cytotoxic effects. Within this study, 557.3 ± 49.0 and 158.8 ± 23.1 μmol 4-aminothiophenol have been coupled per gram pectin. Within both conjugates, around 75% of the bound ligand appeared in its reduced form. Within rheological studies, a 500-fold increase in viscosity was achieved by addition of hydrogen peroxide as an oxidizing agent. Investigations on the swelling behavior revealed that this hydrophobic modification of pectin results in decelerated water uptake on the one hand and improved cohesive properties after oxidation of thiol groups to disulfide bonds on the other hand. Thereby, the maximum amount of water which can be uptaken by pectin matrices could be increased. According to these results, Pec-ATP conjugates could be valuable tools for several pharmaceutical applications due to the established method of gelation and the altered swelling and disintegration behavior.  相似文献   

17.
This study describes a synthesis method of biodegradable macroporous hydrogels suitable as in situ cross-linkable biomaterials. Macroporous hydrogels were based on poly(propylene fumarate-co-ethylene glycol) and prepared via coupled free radical and pore formation reactions. Cross-linking was initiated by a pair of redox initiators, ammonium persulfate and L-ascorbic acid. Pores were formed by the reaction between L-ascorbic acid and sodium bicarbonate, a basic component, which evolved carbon dioxide. Sol fraction of the hydrogels was varied from 0.06 +/- 0.01 to 0.64 +/- 0.01. A stereological approach was used to analyze the morphological properties of the macroporous hydrogels by relating the morphological properties of thin sections to the original three-dimensional macroporous hydrogel. Prepared macroporous hydrogels had porosities between 0.43 +/- 0.08 and 0.84 +/- 0.02 and surface area densities between 55 +/- 3 and 108 +/- 7 cm(-1). Sodium bicarbonate concentration had the greatest effect on both the porosity and surface area density. The effect of copolymer formulation on the porosity and surface area density was insignificant. From thin sections of the macroporous hydrogels, the profile size distributions were determined as an estimate of the pore size distribution. Two formulations synthesized with varying L-ascorbic acid concentration of 0.05 and 0.1 M had median profile sizes of 50-100 and 150-200 microm, respectively. This novel synthesis method allows for the in situ cross-linking of biodegradable macroporous hydrogels with morpholological properties suitable for consideration as an injectable tissue engineering scaffold.  相似文献   

18.
采用不同的提取液,对10个小麦品种的非酶功能性种子储藏蛋白进行提取,分别进行梯度凝胶电泳分析。电泳依据提取液的不同,分别采用酸性或碱性系统。对酸性凝胶催化系统,采用Ap-Vc-FeSO4系统代替H2O2-Vc-FeSO4系统,克服了酸性凝胶的不足,提高了凝胶的性质性能并使之容易操作。应用新的催化系统配制的酸性梯度胶,提高了分辨率。并初步尝试以酸性系统分析种子谷蛋白,获得了成功,经过对不同提取液蛋白  相似文献   

19.
Three glucuronate-rich dermatan sulfate proteoglycan (DS-PG) subclasses were isolated and previously characterized from bovine aortic endothelial cell cultures (Kinsella, M. G., and Wight, T. N. (1988) J. Biol. Chem. 263, 19222-19231). In the present study, pulse-chase experiments indicate that the DS-PG of highest apparent Mr (approximately 1 x 10(6)), denoted previously as HMW-DS, is a relatively stable component of the endothelial extracellular matrix and is formed at the expense of lower Mr DS-PG species. The formation of HMW-DS is reduced in a dose-dependent manner in the presence of dansylcadaverine, an inhibitor of transglutaminase-catalyzed protein cross-linking, but not when the activity of other cross-linking enzymes such as lysyl oxidase is inhibited. The putative DS-PG precursor to HMW-DS accumulates during inhibition of cross-linking only when lysosomal degradation is also inhibited by ammonium chloride, suggesting that the precursor is degraded rapidly in the absence of cross-linking. HMW-DS is precipitable from endothelial cell monolayer extracts with antibodies against fibronectin, a known transglutaminase substrate. Thus, we conclude that the stability of HMW-DS in the subendothelial matrix in culture depends upon the cross-linking of a low Mr DS-PG precursor to matrical protein(s), including fibronectin, resulting in the formation of a DS-PG subclass of high apparent molecular mass.  相似文献   

20.
The kinetic behavior during gel formation and the microstructure of 0.75% high methoxyl (HM) pectin gels in 60% sucrose have been investigated by oscillatory measurements and transmission electron microscopy for three comparable citrus pectin samples differing in their degree of blockiness (DB). Ca2+ addition at pH 3.0 resulted in faster gel formation and a lower storage modulus after 3 h for gels of the blockwise pectin A. For gels of the randomly esterified pectin B, the Ca2+ addition resulted in faster gel formation and a higher storage modulus at pH 3.0. At pH 3.5, both pectins A and B were reinforced by the addition of Ca2+. In the absence of Ca2+, the shortest gelation time was obtained for the sample with the highest DB. Microstructural characterization of the gel network, 4 and 20 h after gel preparation, showed no visible changes on a nanometer scale. The microstructure of pectins A and B without Ca2+ was similar, whereas the presence of Ca2+ in pectin A resulted in an inhomogeneous structure.  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