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 174 毫秒
1.
Mixed gels of cassava starch (CS) and a whey protein isolate (WPI), obtained by heating solutions of 10% total solids, pH 5.75 to 85°C, were characterized as a function of the starch fraction, θs, by axial compression, small-amplitude oscillatory rheometry, differential scanning calorimetry (DSC) and scanning electron microscopy (SEM). Gelation did not occur for θs > 0.7. In the range 0<θs < 0.4 mixed gels showed higher mechanical (E, elastic modulus) and rheological (G′, storage modulus) properties than pure gels, with maximum values for θs= 0.2–0.3. Viscoelastic measurements as a function of time showed that gels containing higher levels of WPI developed a larger G. Blends of both biopolymers showed independent thermal transitions in DSC measurements, related to gelatinization and denaturation. Microstructure of a mixed gel formed at θs= 0.2 showed a continuous matrix formed by strands of WPI particle aggregates and an independent CS phase.  相似文献   

2.
Hydrogels with moderate water retention capacity and salt stability have been prepared from starch, oxidized starch, and amylopectin all crosslinked with epichlorohydrin. The swelling power was measured in water and in various salt solutions of different concentrations. The swelling capacity decreases with increasing both the crosslinker and the polymer concentration. The plateau modulus (G′p) of the prepared gels was determined from complex dynamic viscosity measurements. G′p was measured as a function of the epichlorohydrin and polymer concentrations. The loss modulus (G″) for the prepared gels was also measured and was always found to be almost zero. The molecular weight between two entanglement points (mesh width) was estimated from the G′p values. The ratio between the number of elastically effective entanglement points, calculated from the G′p values, and that calculated from complete conversion of the cross-linker was also estimated for various gels. It was found that this ratio is very small (0.2–0.7%) for gels prepared with 10 wt-% starch, it increases however to 13% at higher starch concentration of 25 wt-%.  相似文献   

3.
Concentration dependence of mechanical modulus of two biopolymer systems, i.e., xanthan-locust bean (X/L) mixture and fish muscle protein (surimi) was evaluated and compared at a wide range of polymer concentrations. A small amplitude oscillatory shear test was performed to measure changes in storage (G′) modulus during gelation and after gelation. Critical concentration (Cc) of the X/L mixture and surimi gel was determined to be 0.15 g/100 mL of solvent and 2.04 g/100 g of solvent, respectively. Reduced concentration (CR=CL/Cc) was used to compare the power-law dependence of modulus of the two systems. The elasticity exponent of the X/L mixture and surimi gel was determined to be 2.4 and 1.97, respectively. The concentration dependence of two biopolymer gel systems such as physical gels (X/L) and chemicals gels (fish muscle protein) theoretically demonstrated that the difference of flexibility of junctions in the networks might distinguish the elasticity of each gel.  相似文献   

4.
Mixed starches with an amylose content of 5, 10, 18, 20, 23, and 25% were prepared by blending starches isolated from waxy and non‐waxy wheat at different ratios. The dynamic viscoelasticity of mixed 30% and 40% starch gels was measured using a rheometer with parallel plate geometry. The change in storage shear modulus (G′) over time at 5 °C was measured, and the rate constant of G′ development was estimated. As the proportion of waxy starch in the mixture increased, starch gels showed lower G′ and higher frequency dependence during 48 h storage at 5 °C. Since the amylopectin of waxy starch granules was solubilized more easily in hot water than that of non‐waxy starch granules, mixed starch containing more waxy starch was more highly solubilized and formed weaker gels. G′ of 30% and 40% starch gels increased steadily during 48 h. 30% starch gel of waxy, non‐waxy and mixed starches showed a slow increase in G′. For 40% starch gels, mixed starch containing more waxy starch showed rapidly developed G′ and had a higher rate constant of starch retrogradation. Waxy starch greatly influenced the rheological properties of mixed starch gels and its proportion in the mixture played a major role in starch gel properties.  相似文献   

5.
Measurement of Gelpoint Temperature and Modulus of Pectin Gels   总被引:1,自引:0,他引:1  
Methods which permit the measurement of gelpoint (setting) temperature (Tgel) and rigidity modulus (G) of pectin gels were improved. Gel development on cooling was determined with an oscillatory pressure testing device capable of detecting a modulus as low as 3 Pa and strain values no greater than 5∞10?3. Sample was set in a glass “U” tube during pressure oscillatory assay for Tgel determination, and transferred after gelling in the same tube for G modulus determination with modified Saunders-Ward apparatus. Results confirmed rheological behavior reported for these kinds of gels and compared favorably with small amplitude oscillatory measurements performed with a stress controlled rheometer using cone and plate (4°, dia = 4 cm) geometry at different frequencies (0.5 to 1.5 Hz) and cooling rates (1 to 3°C/min).  相似文献   

6.
Mixed cold-set whey protein isolate (WPI)–flaxseed gum (FG) gels, induced by the addition of CaCl2 or NaCl at fixed ionic strength (150 mM), were evaluated with respect to their mechanical properties, water-holding capacity (WHC) and SEM microscopy. They were prepared by mixing FG and thermally denatured (90 °C/30 min) WPI solutions at room temperature, but the gels were formed at 10 °C using two methods of salt incorporation: diffusion through dialysis membranes and direct addition. The mixed systems formed using dialysis membranes showed phase separation with the development of two (axial) layers, and the CaCl2-induced gels presented radial phase separation. In general the CaCl2-induced gels were less discontinuous, stronger, and showing lower WHC and deformability than the NaCl-induced gels. An increase in the FG concentration reduced the gel strength and WHC for both systems, which was associated with a prevailing phase separation between the biopolymers over the gelation process. Using direct salt addition, apparently none of the mixed gels showed macroscopic phase separation, but the NaCl-induced gels showed much higher hardness and elasticity than the CaCl2-induced gels. Since the gelation process occurred more quickly by direct salt addition, and more effectively for the divalent salts, the more fragile structure of the CaCl2-induced gels was a consequence of disruption of the cross-link interactions of the aggregates during the agitation used to homogenize the salt added.  相似文献   

7.
Using the method of mathematical planning of experiments and taking the shear modulus of the gels as an equivalent of the gelation degree, the conditions of the thermal denaturation at the maximum shear modulus were found to be TD ≈? 100 °C and tD ≈? 60 min for the non-acetylated Vicia faba protein and TD ≈? 100 °C and tD ≈? 40 min for the acetylated Vicia faba protein. On these conditions the dependence of the shear modulus (G) on the protein concentration (C) is expressed by exponential functions G = α C4.3 and G = α C5.0, respectively. By an increasing acetylation of the Vicia faba protein the shear modulus, the thickness of the many-line chains, and the net density of gels are increased; the same is valid for the shear modulus and the masking of disulfide groups by an increasing concentration of non-acetylated and acetylated Vicia faba proteins.  相似文献   

8.
Egg white from which the precipitate occurring during dialysis against water had been removed gave a transparent gel on heating at lower salt concentrations and pH 3.54. The addition of NaCl or a shift of pH from 3.54 to 4.50 brought about the turbid gels. Creep analyses of these transparent and turbid gels were done using a four-element mechanical model. The instantaneous elastic modulus, E0, Newtonian modulus, ηn and retarded viscoelastic moduli, E1 and η1 increased with NaCl up to 125 mM and then decreased with increase in NaCl concentration at pH 3.54. At 125 mM NaCl and pH 3.54, a translucent gel was obtained. E0 and η1 decreased with the temperature of the gel, while E1 and η1did not change depending of temperature.  相似文献   

9.
Factors Affecting the Gelation Properties of Hydrolyzed Sunflower Proteins   总被引:2,自引:0,他引:2  
The effects of temperature and several chemicals on gelation time and strength of gels formed by heating (pH 8) 5% solutions of trypsin hydrolyzed sunflower proteins were studied by dynamic rheological methods. The storage modulus reached a maximum at 80°C. Ca2Cl (and NaCl at > 0.2M) accelerated gelation and weakened the gel. NaCOCH3Na2SO4 and NaSCN decreased the storage modulus. Urea decreased gelstrength and at high concentrations slowed gelation. Time for gelation diminished and gel strength increased with increasing mercaptoethanol concentration up to 0.1M. Propylene glycol at 5–20% concentrations accelerated gelation and at 5% also increased gel strength. Trypsin hydrolyzed sunflower proteins could be useful in products requiring strong gels at high temperatures.  相似文献   

10.
Mixed and pure gels of gelatin and whey protein concentrate (WPC) were formed by using temperature and high pressure simultaneously. Combining these gel formation methods enables the two polymer networks to set at the same time. The microstructure of the gels was studied by means of light microscopy and transmission electron microscopy, and the rheological properties by means of dynamic oscillatory measurements and tensile tests. The pH values investigated were 5.4, 6.8 and 7.5. The isoelectric point of the WPC is around pH 5.2 and that of gelatin between pH 7.5 and 9. At pH 5.4, the mixed gel formed a phase-separated system, with a gelatin continuous network and spherical inclusions of the WPC. The storage modulus (G) of the mixed gel was similar to that of a pure gelatin gel. At pH6.8, the mixed gel formed a phase-separated system, composed of an aggregated network and a phase with fine strands. The aggregated network proved to be made up of both gelatin and WPC, and the fine strands were formed of gelatin. The mixed gel at pH 6.8 showed a high G compared with the pure gels, which decreased significantly when the gelatin phase melted. At pH 7.5 the mixed gel was composed of one single aggregated network, in which gelatin and WPC were homogeneously distributed. It was impossible to distinguish the gelatin from the WPC in the mixed network. The mixed gel at pH 7.5 showed a significantly higher G than the pure gels. As the gelatin phase was melted out for the mixed gel, a large decrease in G was observed. The pure gelatin gels, formed by a temperature decrease under high pressure, proved to be pH-dependent, showing an increase in aggregation as the pH increased from 5.4 to 7.5. A fine-stranded, transparent gelatin gel was formed at pH 5.4, while an aggregated, opaque gel was formed at pH 7.5. The stress at fracture for the gelatin gels decreased as the aggregation, and consequently the pore size, increased.  相似文献   

11.
The waxy starches used in this investigation (maize (WM), barley (WB) and two rice starches RD4 and IR29) showed different gelatinization temperatures (GT) and enthalpies (ΔHG) measured with differential scanning calorimetry (DSC). The differences in GT and ΔHG could only partially be related to X-ray crystallinity. The high GT waxy starches WM and RD4 retrograded to a greater extent measured with DSC and the order of increased retrogradation agreed well with the order of X-ray crystallinity of the retrograded waxy starches. The melting temperature (TC) of the retrograded waxy starches in contrast to GT was very similar for all starches. This indicates that the temperature of the glass transition (Tg) of the amorphous regions in the starch granules controlled the onset GT and perhaps also the extent of retrogradation. Addition of cetyl trimethyl ammonium bromide (CTAB) decreased the extent of retrogradation more than 45% compared to the melting enthalpy (ΔHC) of the waxy starches without CTAB addition. The rice starch RD4 was most affected by the CTAB addition, and the WM starch the least. The viscoelastic behaviour in the temperature interval 25–90°C of 12%(w/w) gels differed between the waxy starches. The WB starch gels showed the highest storage modulus (G′) value and the lowest phase angle (δ), i. e. strongest and least viscous gels. The WM starch gels showed the lowest G′ value and the highest δ. The rice starches were in between with the RC4 starch (high GT) showing higher G′ value and lower δ than the IR29 (low GT). The viscoelastic parameters changed only slightly with increased temperature. The addition of CTAB to the waxy starch gels changed the viscoelastic behaviour of the stronger and less viscous starch gels of the WB and RD4 as their G′ value decreased and δ increased with increased temperature. The effect on WM and IR29 was only small.  相似文献   

12.
The phase transition temperatures, rheological properties and gel‐network characteristics for gelation of κ‐carrageenan‐salt (NaCl, KCl and CaCl2) solutions and their aged gels were investigated. The effectiveness of increasing gelling and gel‐melting temperatures at the salt concentrations examined followed the sequence of K+ > Ca2+ > Na+. This sequence was also true for the gel strength and the melting enthalpy (DH) of the most crosslinked junction zones of aged gels at low salt concentrations. Nonetheless, a different order (Ca++ > K+ and Na+) was found for increasing storage modulus and gelation rate during early‐stage gelation, thermal hysteresis and the DH of aged gels in some salt‐carrageenan systems.  相似文献   

13.
《Food Hydrocolloids》2001,15(4-6):415-421
The gel properties of commercial egg white (EW) with high gelling ability were compared with those of standard type using creep and compression measurements. The effects of heating time, temperature, and protein concentration on gel properties were investigated at pH 7.5. The experimental creep curves were analyzed using four elements of the Maxwell–Voigt model. The difference in rheological properties of the two types of EW was reflected in the values of elastic modulus and viscosity. The experimental stress–strain curves for all EW gels were fitted using the BST equation with high accuracy. This equation contained two fitting parameters, i.e. Young's modulus and the elasticity parameter nBST. The elasticity parameter nBST characterized non-linear elastic behavior for large deformation. At 80°C, nBST of both types of EW gels did not change with increasing heating time. However, nBST decreased with increasing temperature in a range of 70–85°C.  相似文献   

14.
Heat Gelation Properties and Protein Extractability of Beef Myofibrils   总被引:3,自引:0,他引:3  
At a heating rate of 1oC/min suspensions (pH 6.0) of isolated beef myofibrils were found to start forming gels at 43-56oC, as detected by dynamic rheological measurements. The increase in gel storage modulus levelled off at temperatures > 65oC. At medium to high (0.3-0.6M) concentrations of sodium chloride, addition of pyrophosphate (plus magnesium chloride) had the following effects: (1) both the protein extractability of non-heated myofibrils and the storage moduli of heat-induced gels were markedly increased; (2) the apparent activation energy for gel formation was decreased. Increasing concentrations of sodium chloride, up to 0.5-0.6M, increased the protein concentration of the liquid phase of the gels.  相似文献   

15.
Formation of bovine myosin gels (10 mg ml?1) by heat treatment at pH 6 and an ionic strength of 0.24 M has been monitored by using the Bohlin Rheometer System in the oscillatory mode. Rheological thermograms were determined with a general repeatability of about 2% for a given suspension. A pronounced maximum and an accompanying minimum in storage modulus (G′) were found at about 50 and 55°C, respectively. The thermograms for the loss modulus (G″) and the phase angle (δ) displayed complex behaviour as well, suggesting a multitransition process. Presumably, denaturational events in parts of the molecule are responsible for the complex rheology observed. This complexity is not related to trivial wall slippage as data obtained from cells with different gap sizes were highly reproducible and consistent with other measurements. A decrease in heating rate from 2.5 to 0.1°C min?1 had a large effect on G′; it increased from 905 to 1600 N m?2 for gels at 75°C. The phase angle was also affected by the heating rate, especially at about 55°C. The effect of increasing the strain from 0.003 to about 0.1 was significant in two temperature regions; G′ at temperatures higher than 65°C and δ at temperatures lower than 54°C increased with increasing strain.  相似文献   

16.
Effect of pH (4.5–7.5) and Ca2+ (0.01–0.5 m ) on gelation of single and mixed systems of 10% β‐lactoglobulin (BLG) and 1% basil seed gum (BSG) was investigated. The gelling point of BLG and BSG gels was strongly pH‐dependent, and stiffer gels formed at higher pH. The BLG gels were formed upon heating to 90 °C and reinforced on cooling to 20 °C; however, the gelation of BSG occurred at temperatures below 70 °C. By increasing Ca2+ concentration, storage modulus of BLG and BSG gels were increased, although pH had a greater effect than Ca2+. In contrast, mixed systems showed two distinct types of behaviour: BLG gel formation and BSG network, suggesting that phase‐separated gels were formed. In addition, higher strength was obtained for BLG‐BSG mixture at higher Ca2+ concentration.  相似文献   

17.
Viscoelastic properties of two nontraditional hydrocolloid dispersions were evaluated. Prosopis chilensis seed gum was evaluated based on temperature (5–80 °C) and added CaCl2 (0.07%), whereas nopal mucilage was evaluated based on temperature (5–80 °C) and sucrose concentration (0–20%). Viscoelasticity was tested by the small strain oscillatory shear test; storage modulus (G′), loss modulus (G″) and tan δ were reported. Prosopis chilensis and nopal dispersions behaved as weak gels (G’ > G’’) regardless of experimental condition. Raising temperature from 20 to 80 °C significantly increased G’. The gel structure was strengthened by adding CaCl2 and G’ increased at 40 °C. The sucrose effect depended on concentration and temperature; at low sucrose concentrations, G’ modulus increased regardless of temperature level, but at high concentrations, it decreased at temperatures >40 °C. In conclusion, nopal and Prosopis chilensis dispersions show weak gel structure regardless of experimental condition. G′ increases as temperature increases, and these dispersions could be suitable for food applications requiring heat tolerance.  相似文献   

18.
《Food chemistry》2001,74(2):161-167
The gel properties of two different kinds of fish gelatins prepared from cod (Gadus morhua) and hake (Merluccius merluccius) and modified by the coenhancers glycerol, salt and microbial transglutaminase, were examined. Gel strength was substantially increased by the addition of coenhancers although results varied, depending on the species. In gelatin from hake (M. merluccius) skin, the highest values were obtained with 10 mg/g of transglutaminase, whereas magnesium sulphate was more effective at both concentrations (0.1 and 0.5 M) in gelatin from cod (G. morhua) skin. Although, in both gelatins, the addition of any ingredient increased the viscosity modulus (G″), the elastic modulus (G′) was only increased by the addition of glycerol 15% (w/v) and MgSO4 0.5 M in hake (M. merluccius) gelatins; in cod (G. morhua) it was increased by all ingredients. The gelling and melting points, very important properties in fish gelatin, showed a notable improvement, the behaviour being different, depending on the species.  相似文献   

19.
Texture and sensory properties of starch solutions or gels, resp., can be influenced by regulating the visco-elastic properties that cannot be adjusted by simply varying the waxy rice starch concentration. Therefore, the hydrocolloid galactomannans guar gum and locust bean gum were used for addition. A total polymer concentration of c = 5% was maintained through all tests, the hydrocolloid fractions varied (cHK = 0.5%, 1%, 2%). The mixtures were entirely decomposed in an autoclave at 130°C resulting in an almost complete destruction of the granular starch structure. With rheomechanical oscillation measurements in the linear-visco-elastic range the influence of the galactomannans onto viscosity yield and visco-elasticity was determined quantitatively. It was found that the influence on the viscosity is depending on the deformation rate. At low frequencies and with increasing galactomannan fraction the viscosity decreases, while above a critical frequency ωkr an increase in viscosity could be detected compared to pure starch. The mixtures also displayed an increase in viscous fractions versus elastic fractions compared to pure starch. With increasing galactomannan fraction, the ratio storage modulus/loss modulus is reduced continuously from G′:G″ > 10 (definitely mainly elastic) to G′:G″ = 1 with simultaneously decreasing influence of the frequency. Based on the presented data and knowledge on solution structure of the individual components, a model for describing network structures in mixtures was developed. Changes in viscous-elastic properties are based on a thermodynamic incompatibility between starch and galactomannan fractions. The mixtures contain mixed network structures, with waxy rice starch (permanent network points) and galactomannans (temporary network points) contributing individually and characteristically to the resulting flow behaviour.  相似文献   

20.
Young's modulus of heat-denatured gels of calcium alginate and bovine serum albumin (BSA) was determined and compared to the modulus of BSA gels containing sodium alginate and to pure BSA gels. Ionic strength, pH, and calcium concentration were varied. The BSA/Ca-alginate gels were either prepared with -glucono-δ-lactone (GDL) and CaCO3 to induce alginate gelation before the gelation of BSA, or by soaking heat-denatured BSA/Na-alginate gels in a CaCl2 solution. BSA/Ca-alginate gels were stronger than BSA/Na-alginate gels at all conditions, and stronger than pure BSA gels up to higher pH values and up to somewhat higher ionic strengths than BSA/Na-alginate gels. The strength of BSA/Ca-alginate gels was highly dependent on the strength of the alginate gel. This was shown by variation of the calcium concentration and by soaking the gels in EDTA, NaCl, and CaCl2 solutions. When BSA/Na-alginate or BSA/Ca-alginate gels prepared at optimum conditions were soaked in solutions of higher ionic strength or pH, no reduction in gel strength was observed. Consequently, they were much stronger than gels that were prepared directly at high pH or ionic strength. The results may suggest that the alginate network in a BSA/Ca-alginate gel increases the effectiveness of electrostatic BSA-alginate cross-links or entanglements. However, other explanations are also possible.  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