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.
Water-soluble (WSP) and insoluble non-starch polysaccharides (WIP) were isolated from wheat flour to evaluate the effects of WSP and WIP on starch gel properties. Isolated WSP and WIP were added to two types of isolated wheat starch with different amylose content at a concentration of 3% based on the dry weight of starch. 30% starch gels were prepared and stored at 5 °C for 1, 8, or 24 h. The dynamic viscoelasticity of 30% starch gels mixed with WSP and WIP was measured using parallel plate geometry, showing that WSP and WIP affected the elastic component of starch gels in opposite ways. Adding WIP increased the storage shear modulus (G′) of starch gels, while adding WSP decreased G′ and dramatically increased the loss tangent (tan δ=G″/G′).  相似文献   

2.
In a continuation of the work reported in the preceding paper, the effect of chicory root inulin (Mr≈4.5 kD) and branched dextrans of Mr 67, 464 and 2000 kD on gelation of low methoxy pectin (DE 31; 2.0 wt%; pH≈2.9–3.0) on cooling from 90 to 5 °C in the presence of stoichiometric Ca2+ has been characterised by low amplitude oscillatory measurements of G′ and G″. As found for other polymeric cosolutes studied previously (oxidised starch, potato maltodextrin, guar gum and locust bean gum), increasing concentrations of dextran or inulin caused a progressive increase of G′ and G″ in the solution state at 90 °C (attributed to segregative interactions promoting formation of calcium-mediated ‘egg-box’ junctions between pectin chains) and a progressive reduction in final moduli at 5 °C (attributed to excessive association of calcium pectinate into large aggregated bundles), with cooling curves crossing those for calcium pectinate alone at 55 °C. For the dextran samples, the ability to promote initial association of pectin chains at high temperature decreased with increasing molecular weight (i.e. with the smallest molecules having the greatest effect, as was also found for guar gum), but the reduction in final modulus on completion of cooling to 5 °C became more pronounced. By the criterion of the decrease in the change in log G′ during cooling, the three dextran samples studied seem roughly comparable in their overall effectiveness in promoting self-association of pectin, with inulin having a substantially smaller effect. Both materials, however, are less effective than the partially depolymerised starches and galactomannans studied previously. It is tentatively suggested that this may be associated with the flexibility introduced by the presence of three single bonds in the predominant glycosidic linkages of both dextran and inulin.  相似文献   

3.
Mixed gels of κ-carrageenan (κ-car) from Hypnea musciformis and galactomannans (Gal) from Cassia javanica (CJ) and locust bean gum (LBG) were compared using dynamic viscoelastic measurements and compression tests. Mixed gels at 5 g/l of total polymer concentration in 0.1 M KCl showed a synergistic maximum in viscoelastic measurements for κ-car/CJ and κ-car/LBG at 2:1 and 4:1 ratios, respectively. The synergistic maximum obtained from compression tests carried out for mixed gels at 10 g/l of total polymer concentration in 0.25 M KCl was the same for both κ-car/CJ and κ-car/LBG gels. An enhancement in the storage modulus (G′) and the loss modulus (G″) was observed in the mechanical spectra for the mixtures in relation to κ-car. The proportionally higher increase in G″ compared with G′, as indicated by the values of the loss tangent (tan δ), suggests that the Gal adhere non-specifically to the κ-car network.  相似文献   

4.
Addition of gum arabic (average Mr≈450 kDa; 0.5–2.0 wt%) to solutions of low methoxy pectin (DE 31; 2.0 wt%; pH≈2.9–3.0) with stoichiometric Ca2+ caused massive increases in G′ and G″ in the pre-gel state at 90 °C (attributed to segregative interactions promoting formation of calcium-mediated ‘egg-box’ junctions between pectin chains) but had little effect on the gels formed on cooling to 5 °C. This is in marked contrast to the behaviour of other polymeric cosolutes studied in the investigations reported in the two preceding papers, which caused large reductions in gel moduli (attributed to excessive association of calcium pectinate into large aggregated bundles); the difference is tentatively ascribed to strengthening of the calcium pectinate network by divalent counterions to the uronate residues in gum arabic. When the complication of cation exchange was eliminated by extensive dialysis of gum arabic against 100 mM Na+ and use of the final dialysate in preparation of mixtures with calcium pectinate, massive increases in G′ and G″ at high temperature were again observed, but with accompanying reductions in moduli at low temperature, which, at gum arabic concentrations above 1.0 wt%, arose from collapse of the developing calcium pectinate network during cooling. The tentative conclusion from this work, and from the two preceding papers, is that enthalpically unfavourable (segregative) interactions between low methoxy pectin and polymeric cosolutes can be relieved in two ways: (i) Ca2+-mediated self-association of pectin into compact ordered assemblies which occupy less of the total volume, and (ii) conformational rearrangement of the cosolute molecules to minimise segmental interactions with pectin; conformational rearrangement is inhibited by chain stiffness and by branching; thus polymeric cosolute molecules of limited flexibility are more effective in promoting self-association of pectin than more flexible molecules of comparable size, and branched molecules are more effective than linear chains of comparable stiffness.  相似文献   

5.
Pumpkin pectin: gel formation at unusually low concentration   总被引:2,自引:0,他引:2  
The gel properties of high-methoxy pectin from pumpkins have been investigated to assess the potential of this material as a hard-currency export from the former Soviet Union. Comparison was made with commercial slow-set, medium-set and rapid-set pectins from citrus peel. Gels were formed by cooling pectin solutions (pH 3·0; 60% (w/w) sucrose; 5% (w/w) corn syrup) from 95°C to 25°C, and the time-temperature course of network formation was monitored by small-deformation oscillatory measurements of storage modulus (G′). At concentrations above 1% (w/w) the pumpkin pectin gave weaker gels than the other three samples, but its minimum critical gelling concentration (c0) was found to be much lower (by at least a factor of five). Compression testing gave similar results, with pumpkin pectin giving useful breaking-stress (‘hardness’) at concentrations down to 0·5% (w/w), about a factor of two lower than for the citrus samples. Its gelation was also less rapid, giving G′ values below those of the other three samples at temperatures down to 60°C, but then setting sharply; this behaviour could be useful in avoiding ‘pregelation’ in commercial processing. The commercial slow-set pectin showed typical ‘weak gel’ properties in the solution state at 95°C, with systematic reduction in gel-like character with increasing ester content in the other samples. The rigidity of the final gels also decreased systematically through the series: rapid-set < medium-set < slow-set. These observations are tentatively ascribed to stable association of unesterified galacturonate chain segments at low pH, where electrostatic repulsion is suppressed.  相似文献   

6.
Recent studies on dilute solutions (Borgström et al. (1996), Int. J. Biol. Macromol. 18, 223) have shown that kappa-carrageenan helices associate into superhelical rigid rods in mixed 0.1 M aqueous solutions of NaI and CsI above a critical mole fraction (xCs = 0.4) of Cs. This work concerns the temperature-dependent rheology of more concentrated systems in mixed and pure solutions of the same salts. Gels with low moduli were even found in NaI alone, although this salt is known to impede the gelation of kappa-carrageenan, but only above 0.9% (w/w) of carrageenan. These gels were reminiscent of iota-carrageenan gels in two respects: the (low) magnitude of the shear storage modulus (G′), and the absence of hysteresis in the sol-gel transition. On the other hand, both the threshold concentration for gelation and the ratio between the loss and storage moduli were substantially higher for the kappa-carrageenan gels in NaI. In mixed solutions of CsI and NaI, two types of kappa-carrageenan gels could be distinguished, depending on the cesium content. The transition occurred at xCs = 0.4, as in the previous studies on dilute solutions. Below xCs = 0.4, the gels were similar to those in NaI alone. Above xCs = 0.4, the gels were similar to ‘conventional’ kappa-carrageenan gels, formed in salts such as KC1: a pronounced thermal hysteresis appeared in the sol-gel transition, the gels showed tendencies for syneresis, and G′ increased dramatically with increasing cesium content.  相似文献   

7.
As a prelude to studies of co-gelation with galatin, the gelation behaviour of Paselli maltodextrins SA-6 and SA-2 (DE ≈ 6 and 2, respectively) was mapped out over the experimentally-accessible range of temperature (T) and concentration (c), using a simple visual method to determine the time required for formation of a self-supporting network (tg). For both samples, log tg decreased linearly with log c and increased linearly with T. At equivalent temperatures and concentrations, SA-2 gelled between 20 and 60 times faster than SA-6.

Selected samples were monitored more rigorously by mechanical spectroscopy, taking tg as the time at which elastic response (G′) became greater than viscous response (G″). In all cases the values of tg obtained by this procedure were lower than those from visual inspection, by a constant factor of about 3·4.

The concentration-dependence of gel moduli (G′) for SA-2 and for gelatin (second-extract limed ossein; LO-2) fitted accurately to the form anticipated from cascade theory for normal polymer networks. For SA-6, by contrast, log G′ varied linearly with log c over the entire range at which measurements could be made, indicating a different mechanism of structure-formation (such as the agglomeration of short, aggregated helices).  相似文献   


8.
A series of thermoreversible konjac glucomannan gels crosslinked by organic borate were prepared. The gel network was formed through the crosslinking reaction between borate ions dissociated from organic borate and the cis-diol hydroxyl groups on the mannose units of polysaccharide chains. The rheological properties of the complex gels were studied by dynamic viscoelastic measurement. The gelation kinetics of the gels were studied and the critical gelation points of the gels were exactly determined by the Winter–Chambon criterion. The effects of temperature and composite ratio on the shear storage modulus (G), the loss modulus (G), and the sol-gel transition points were investigated. The critical gel-sol temperatures of the complex gels were successfully elucidated by Winter–Chambon criterion. The effect of crosslinking density on the critical gelation temperature and the elasticity of the gels were discussed.  相似文献   

9.
Rheological measurements have been performed on three molecular weight fractions of purified canine submaxillary mucin (CSM) dissolved in the chaotropic solvent 6 M guanidine hydrochloride (GdnHCI). Solutions of the lower molecular weight fractions are viscoelastic sols, and their dynamic moduli can be scaled with respect to molecular weight and concentration according to linear viscoelasticity theory. In contrast, preparations of the highest molecular weight fraction form viscoelastic gels that exhibit an equilibrium shear modulus, Ge, which scales with mucin concentration as Ge c3. Amino acid and carbohydrate analyses of all three fractions are similar; thus, the differences in rheological behavior are attributed to molecular weight differences, which affect the degree of coil overlap in solutions of a given concentration. These observations demonstrate conclusively that mucin glycoproteins of high molecular weight form gels under conditions in which the mucin chains physically interpenetrate, even when non-covalent intermolecular interactions are extensively disrupted. A comparison of these results with previous studies of purified submaxillary and tracheobronchial mucins indicates that the carbohydrate side-chain length, in addition to molecular weight, is an important determinant of the observed elastic response and the ability to form physical gels  相似文献   

10.
Several constitutive models have been discussed to explain data for some foods in diluted and concentrated systems. Firstly, the theories of Rouse and Zimm, as well as rod-like theory, were used to study the conformation of the pectins in dilute solution. Among the dilute theories, the random coil theory of Zimm best explained the experimental data and suggested a certain level of intermolecular interaction present in the dilute pectin solution.

The Bird-Carreau constitutive theory with four empirical constants and zero shear limiting viscosity was used to describe the viscoelastic properties of the solutions of the guar, CMC/guar, glutenin, gluten and wheat flour doughs. The Bird-Carreau model was able to predict η and η′ in the high and low frequency regions for 1% guar solution. In the case of CMC/guar blend, the Bird-Carreau model explained steady shear and dynamic properties very well in the higher shear rate or frequency region of 1–100 s−1. However, η″/ω does not tend to a zero shear constant value. The Bird-Carreau model also gave good predictions on the rheological properties of gluten and glutenin biopolymers in the free-flow region.

The polydisperse type, Doi-Edwards model, fits the experimental G′ and G″ better than the monodisperse model for 5% apple pectin dispersion. However, there is still a discrepancy between experimental and predicted values.  相似文献   


11.
Dynamic oscillatory testing has been used to study the rheology of water-insoluble dextran. The rheological properties (storage and loss moduli) of dextran gel were measured and dextran was found to be neither a strong gel nor a weak gel, but an entanglement network at a concentration of 250 mg/ml. The extent of gelation, illustrated by the gel elastic modulus G′, is found to decrease with increasing concentration of calcium ions. This was confirmed by shift of crossover frequencies towards higher values on the dynamic spectra and lower yield stress τ values obtained from stress ramp experiments. Finally, a comparison between gelation of dextran and alginate (a similar biopolymer) was made for clear understanding of effect of calcium ions on the dextran gelation.  相似文献   

12.
The flow characterization of blood through healthy and diseased flow geometries is of interest to researchers and clinicians alike, as it may allow for early detection, and monitoring, of cardiovascular disease. In this paper, we use a numerically efficient particle-based flow model called multiparticle collision dynamics (MPC for short) to study the effect of compressibility and slip of flow of a Newtonian fluid through a cylinder with a local constriction. We use a cumulative averaging method to compare our MPC results to the finite-element solution of the incompressible no-slip Navier-Stokes equations in the same geometry. We concentrate on low Reynolds number flows [ $Re \in (4,30)$ ] and quantify important differences observed between the MPC results and the Navier-Stokes solution in constricted geometries. In particular, our results show that upstream recirculating zones can form with the inclusion of slip and compressibility, which are not observed in the flow of an incompressible Newtonian fluid using the no-slip assumption, but have been observed experimentally for blood. Important flow features are also presented that could be used as indicators to observe compressibility and slip in experimental data where near-wall data may be difficult to obtain. Finally, we found that the cumulative averaging method used is ideal for steady particle-based flow methods, as macroscopic no-slip is readily obtained using the MPC bounce-back rule. Generally, a small spurious slip is seen using other averaging methods such as weighted spatial averages or averages over several runs, and the bounce-back rule has to be modified so as to achieve macroscopic no-slip. No modifications of the bounce-back rule were required for our simulations.  相似文献   

13.
The storage moduli (G′) of phase-separated co-gels formed by quench-cooling mixed solutions of gelatin and potato maltodextrin (Paselli SA-2 and SA-6) have been related quantitatively to the experimentally-determined concentration-dependence of G′ for the constituent polymers. Distribution of water between the phases was examined explicitly by using polymer blending laws to derive calculated moduli for gelatin-continuous and maltodextrin-continuous networks over the entire range of solvent partition. Allowance was made for the direct contribution of polymer chains, and for density differences between the phases, in calculating relative phase-volumes. The effect of gel formation prior to phase-separation was calculated using classical theory for network de-swelling. Good agreement with observed moduli for more than 30 gelatin/maltodextrin combinations was achieved using a single adjustable parameter, p (the ratio of solvent to polymer in the gelatin phase divided by the same ratio for the maltodextrin phase), with an optimum value of p ≈ 1·8 for both SA-6 and SA-2.  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.
Density functional theory (DFT) computations at the B3LYP/Lanl2DZ level were used to elucidate the oxygen atom transfer (OAT) and coupled electron proton transfer (CEPT) reaction steps involved in the biomimetic catalytic cycle performed by polymer-supported MoVIO2(NN′)2 complexes [NN′ = phenyl-(pyrrolato-2-ylmethylene)-amine] with water as oxygen source, trimethyl-phosphane as oxygen acceptor and one-electron oxidising agents. The DFT method employed has been validated against experimental data [X-ray crystal structures of a NN′ ligand and a MoVIO2(NN′)2 complex as well as kinetic data]. The rate-limiting step in the forward-OAT from [MoVIO2] to PMe3 is the attack of PMe3 at an oxo ligand with ΔG (298 K) = 64.6 kJ mol−1. Dissociation of the product OPMe3 is facile with ΔG (298 K) = 26.3 kJ mol−1 giving a mono-oxo [MoIVO] complex which fills its coordination sphere with a further PMe3 substrate with ΔG (298 K) = 39.2 kJ mol−1. One-electron oxidation to a Mo(V) phosphane complex precedes the coordination of water/hydroxide. Additionally, the comproportionation of [MoVIO2] and [MoIVO] to dinuclear oxo-bridged [OMoV–O–MoVO] species has been calculated as the thermodynamic sink in this system and the back-OAT from dmso to mono-oxo [MoIVO] to give [MoVIO2] has been shown to involve an equilibrium between stereoisomeric [MoVIO2] complexes with an activation barrier of ΔG (298 K) = 113.1 kJ mol−1.  相似文献   

16.
Water-extractable arabinoxylan (WEAX) gels exhibiting different structural, rheological and protein transport properties were obtained upon laccase treatment of WEAX solutions, by modifying (i) the initial ferulic acid (FA) content of WEAX from 2.3 to 1.6 μg/mg AX or (ii) the AX concentration of the WEAX solution from 0.5 to 2.0% (w/v). WEAX gels with di and tri-ferulic acid (di-FA, tri-FA) contents varying from 6.2 to 3.2 μg/ml gel and from 0.61 to 0.27 μg/ml gel, respectively, were obtained. In parallel, increases in gel mesh sizes from 201 to 331 nm and reduction of G′ of gels from 160 to 12 Pa were observed. The differences in structural and rheological characteristics of WEAX gels were reflected in their capacity to load and release proteins of Mw ranging from 43 to 669 kDa. The possibility of modulating protein release from WEAX gels makes these gels potential candidates for the controlled delivery of proteins.  相似文献   

17.
The effect of ageing on the rheological properties of potato starch pastes was investigated by measurement of dynamic viscoelasticity. The value of the storage modulus (G′) increased rapidly with time for the first few hours, but attained equilibrium after a long period of ageing. The relation between the concentration and the storage modulus showed a c1·6 dependence. The Avrami exponent (n) obtained in this study was 0·66 and deviated significantly from the value (n = 1), corresponding to rod-like growth of crystal. The rate of storage modulus increase decreased with increasing concentrations up to 10%, and at higher concentrations little or no increase was observed. In addition, the rate of storage modulus increase increased dramatically when the NaCl concentration was raised to levels of 0·01 or above. An even greater increase in rate occurs on addition of AlCl3.  相似文献   

18.
The concentration-dependence of viscosity in solutions of purified glycoprotein from pig gastric mucus is of the form expected for simple polymer entanglement. At higher concentrations, however, a weak viscoelastic gel is formed, whose mechanical spectrum (over the frequency range 10−2---102 rad s−1) indicates a more stable mechanism of interchain association, and is closely similar to that of native mucus. On prolonged exposure to solvent, reconstituted gels redissolve, while native mucus retains its structural integrity (as characterized by the storage modulus, G′) but releases a significant, variable amount of glycoprotein. On proteolytic digestion or disulphide reduction of the glycoprotein to its component subunits, network structure is lost, but the mechanical spectra of the resulting solutions show interactions beyond simple entanglement. From this evidence we suggest that in the sub-micrometre-sized ‘domains’ in which native mucus is secreted, the carbohydrate side chains of component glycoprotein molecules are interdigitated in a comparatively stable arrangement, with the polymeric subunit structure of the glycoprotein conferring the branching required for development of a three-dimensional network, and with a substantial, variable sol-fraction of free glycoprotein within the interstices of the gel. On solubilization of native mucus, the ‘domain’ structure is destroyed irreversibly. Interaction between domains, and between individual molecules in gels reconstituted from the component glycoprotein after extraction and purification, is by more transient, non-specific interdigitation and entanglement, to confer the overall flow and spreading characteristics of the gel.  相似文献   

19.
The mechanical characterization of soft anisotropic materials is a fundamental challenge because of difficulties in applying mechanical loads to soft matter and the need to combine information from multiple tests. A method to characterize the linear elastic properties of transversely isotropic soft materials is proposed, based on the combination of dynamic shear testing (DST) and asymmetric indentation. The procedure was demonstrated by characterizing a nearly incompressible transversely isotropic soft material. A soft gel with controlled anisotropy was obtained by polymerizing a mixture of fibrinogen and thrombin solutions in a high field magnet (B?=?11.7 T); fibrils in the resulting gel were predominantly aligned parallel to the magnetic field. Aligned fibrin gels were subject to dynamic (20-40 Hz) shear deformation in two orthogonal directions. The shear storage modulus was 1.08?±?0. 42 kPa (mean?±?std. dev.) for shear in a plane parallel to the dominant fiber direction, and 0.58?±?0.21 kPa for shear in the plane of isotropy. Gels were indented by a rectangular tip of a large aspect ratio, aligned either parallel or perpendicular to the normal to the plane of transverse isotropy. Aligned fibrin gels appeared stiffer when indented with the long axis of a rectangular tip perpendicular to the dominant fiber direction. Three-dimensional numerical simulations of asymmetric indentation were used to determine the relationship between direction-dependent differences in indentation stiffness and material parameters. This approach enables the estimation of a complete set of parameters for an incompressible, transversely isotropic, linear elastic material.  相似文献   

20.
Differential scanning calorimetry (DSC), rheological measurements and granule size analyses were performed to characterize the influence of phosphorylation substitution levels on the properties of cross-linked potato starch. Phosphorus oxychloride (POCl3) was used to produce the cross-linked potato starch. The levels of the reagent used for the reaction ranged between 40 and 5000 ppm (dwb). Storage (G′) and loss (G″) moduli were measured for a 5% (w/w) gelatinized starch dispersion stored at 20 °C for 24 h after heating at 85 °C for 30 min. The samples from 80 to 500 ppm were recognized as ‘strong gel'systems, whereas native potato starch showed ‘weak gel'behavior. Steady shear and dynamic viscoelastic properties of gelatinized starch dispersion were compared. Furthermore, granule mean diameter was measured by laser scattering for a 1% (w/w) dispersion heated at 85 °C for 30 min. The granules in the 100 ppm sample swelled to a maximum of about 2.6 times the native starch granule mean diameter.  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