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 14 毫秒
1.
2.
The effect of pHs and heating on the protein–polysaccharide complexation between the 0.5 wt% soy globulin (7S or 11S) and 0.1 wt% chitosan was studied. Electrophoretic and light scattering techniques were used to examine the electrical charge and aggregation of the individual biopolymers and complexes. At pH 3.0–6.5, 7S (or 11S) globulin in the presence of chitosan had significantly higher ζ‐potentials and lower particles size than 7S (or 11S) globulin alone did (e.g. 600–6000 nm at pH 5.5), indicating the formation of complexes. After heating 7S (or 11S)–chitosan mixtures had higher positive value of ζ‐potential. 7S (or 11S)–chitosan mixtures exhibited a significant increase in positive value of ζ‐potential and stability after heating at lower pH values (pH 3.3 instead of pH 4.5). Compared with other mixtures, at pH 2.5–6.0, the most remarkable decrease in aggregation was obtained for 11S–chitosan mixtures after heating at pH 3.3.  相似文献   

3.
Three types of crystalline lactose hydrates having melting points (MP) of 218°C, 195°C and 169°C were crystallised from 680 ml litre?1 ethanol solution with a lactose to solvent ratio of 1:14-1:24 at three pH values, ~ 4, 2.5 and 1.3. The three types were characterised on the basis of their optical rotation, phase transition and X-ray powder diffraction pattern. The samples with high and low MP were α-lactose hydrates whereas that with the intermediate MP (195°C) was a hydrated lactose containing both α- and β-forms.  相似文献   

4.
The selective precipitation of α‐lactalbumin (α‐La) is the basis for one of the possible methods in whey protein fractionation. Calcium concentration, type of acid added and pH play important roles in α‐La precipitation and on the following resolubilisation. Two washing steps are enough for quantitative removal of β‐lactoglobulin entrapped in the precipitate. α‐La losses are minimised during washing steps (5%) when NaCl is used as washing agent. The most important parameter to control during the resolubilisation step is pH, the maximum amount of the initial re‐dissolved α‐La being 76% when the pH is adjusted to 7.5, CaCl2 concentration is 0.2 m and prior precipitation is carried out adding citric acid. Addition of CaCl2 is not necessary to dissolve α‐La because of the fact that there is enough calcium in the precipitate to join all α‐La; however, its presence improves the solubilisation yield (66% vs. 75%).  相似文献   

5.
Thermal denaturation of whey protein solutions was investigated from a structural perspective utilising attenuated total reflectance – Fourier transformed infrared spectroscopy (ATR‐FTIR). Solutions (100 g protein/L, pH 7) of commercial whey protein isolate (WPI) powders and enriched protein fractions of β‐lactoglobulin (β‐lg) and α‐lactalbumin (α‐la) were heat‐treated at temperatures of 50–90 °C. Subsequent analysis by ATR‐FTIR highlighted the structural changes occurring as a direct result of heat treatments. Molecularly, WPI dispersions exhibited pronounced differences in denaturation behaviour depending on their method of manufacture. ATR‐FTIR is an informative tool to discern the structural molecular interactions not apparent through physical analysis of concentrated ingredients.  相似文献   

6.
Fluorescence spectroscopy was used to investigate the interaction between resveratrol and whey proteins. The whey proteins examined were lactoferrin, holo‐lactoferrin, apo‐lactoferrin, whey protein isolate (WPI) and the β‐lactoglobulin‐ and α‐lactalbumin‐rich fractions of WPI. Both an analytical‐grade and food‐grade resveratrol were examined. In all the systems studied, it was found that resveratrol interacted with the whey proteins to form a 1:1 complex. The binding constant, Ks, for the protein–resveratrol complex for all the proteins examined varied from 1.7 × 104 to 1.2 × 105 m ?1. Furthermore, the interaction between the whey proteins and resveratrol did not affect the secondary structure of the proteins.  相似文献   

7.
Yang Y  Anvari M  Pan CH  Chung D 《Food chemistry》2012,135(2):555-561
The interactions between fish gelatin (FG) and gum arabic (GA) in aqueous solutions were investigated by turbidimetry, methylene blue spectrophotometry, zeta potentiometry, dynamic light scattering, protein assay, and state diagram at 40 °C and a total biopolymer concentration (C(T)) of 0.05%. FG underwent complex coacervation with GA, possibly via its conformational change, depending on pH and FG to GA ratio (FG:GA). The formation of FG-GA complexes was the most intense when pH 3.55 and FG:GA=50:50 (6.6:1 M ratio), however, the coacervate phase was found to be composed of a much higher FG fraction. The pH range of complex formation shifted to a higher pH region with increasing FG:GA. Soluble and insoluble FG-GA complexes were formed even in a pH region where both biopolymers were net-negatively charged. Varying C(T) significantly influenced not only the formation of FG-GA complexes but also the development and composition of coacervate phase.  相似文献   

8.
BACKGROUND: Milk proteins are used in a wide range of formulated food emulsions. The stability of food emulsions depends on their ingredients and processing conditions. In this work, β‐carotene nanodispersions were prepared with selected milk‐protein products using solvent‐displacement method. The objective of this work was to evaluate the stability of these nanodispersions against heating, salts and pH. RESULTS: Sodium caseinate (SC)‐stabilised nanodispersions possessed the smallest mean particle size of 17 nm, while those prepared with whey‐protein products resulted in larger mean particle sizes (45–127 nm). Formation of large particles (mean particle size of 300 nm) started after 1 h of heating at 60 °C in nanodispersions prepared with SC. More drastic particle size changes were observed in nanodispersions prepared with whey protein concentrate and whey protein isolate. The SC‐stabilised nanodispersions were fairly stable against Na+ ions at concentrations below 100 mmol L?1, but drastic aggregation occurred in ≥ 50 mmol L?1 CaCl2 solutions. Aggregation was also observed in whey protein‐stabilised nanodispersions after the addition of NaCl and CaCl2 solutions. All sample exhibited the smallest mean particle size at neutral pH, but large aggregates were formed at both ends of extreme pH and at pH around the isoelectric point of the proteins. CONCLUSION: The nanodispersions prepared with SC were generally more stable against thermal processing, ionic strength and pH, compared to those prepared with whey proteins. The stable β‐carotene nanodispersions showed a good potential for industrial applications. Copyright © 2008 Society of Chemical Industry  相似文献   

9.
10.
Study of heat denaturation of major whey proteins (β‐lactoglobulin or α‐lactalbumin) either in separated purified forms, or in forms present in fresh industrial whey or in recomposed mixture respecting whey proportions, indicated significant differences in their denaturation depending on pH, temperature of heating, presence or absence of other co‐denaturation partner, and of existence of a previous thermal pretreatment (industrial whey). α‐Lactalbumin, usually resistant to tryptic hydrolysis, aggregated after heating at ⪈85°C. After its denaturation, α‐lactalbumin was susceptible to tryptic hydrolysis probably because of exposure of its previously hidden tryptic cleavage sites (Lys‐X and Arg‐X bonds). Heating over 85°C of β‐lactoglobulin increased its aggregation and exposure of its peptic cleavage sites. The co‐denaturation of α‐lactalbumin with β‐lactoglobulin increased their aggregation and resulted in complete exposure of β‐lactoglobulin peptic cleavage sites and partial unveiling of α‐lactalbumin tryptic cleavage sites. The exposure of α‐lactalbumin tryptic cleavage sites was slightly enhanced when the α‐lactalbumin/β‐lactoglobulin mixture was heated at pH 7.5. Co‐denaturation of fresh whey by heating at 95°C and pH 4.5 and above produced aggregates stabilized mostly by covalent disulfide bonds easily reduced by β‐mercaptoethanol. The aggregates stabilized by covalent bonds other than disulfide arose from a same thermal treatment but performed at pH 3.5. Thermal treatment of whey at pH 7.5 considerably enhanced tryptic and peptic hydrolysis of both major proteins.  相似文献   

11.
12.
13.
14.
The heat stability of β-lactoglobulin (β-lg) is usually described with reference to a concentration-dependent pseudo-rate constant k. Kinetic and thermodynamic parameters for the irreversible denaturation of β-lg, in a whey protein mixture dissolved in Tris-HCl buffer, were examined over a wide temperature range 75–120°C and for degrees of denaturation [(Co-Ct)/Co] up to about 90%. The first-order kinetic model best described β-lg denaturation over the temperature range 75–85°C, whereas the second-order model applies in the range 90–120°C. A comparison between β-lg thermostability in buffer and literature data pertaining to more complex heating media (whey and milk), over the range 75–120°C, was carried out on the basis of changes in activation free energy (ΔG#), which itself takes into account changes in both activation enthalpy (ΔH#) and activation entropy (ΔS#). It was found that the thermal stability of β-lg in different media, and irrespective of the kinetic model assumed in the present study, can be ranked: buffer ≪ whey>milk for the temperature range 75–85°C. On the contrary, at the higher temperature range, 90–120°C, the ranking is buffer<whey<milk, when using the second-order model to describe the present data. For such comparisons to be valid the initial concentration of β-lg in various studies, as well as the reaction order applied, should be taken into consideration. © 1997 SCI.  相似文献   

15.
Associative phase separation within lentil protein isolate (LPI), polysaccharide (κ‐carrageenan (κ‐CG) and gellan gum (GG)) mixtures was investigated as a function of pH (1.50–8.00) and mixing ratio (1:1–30:1; LPI/polysaccharide) by turbidity and electrophoretic mobility. Effects of salts (NaCl, KCl and CaCl2) on complex stability were also studied as a function of ionic strength. Coacervation typically follows two pH‐dependent forming events associated with the formation of soluble and insoluble complexes. The addition of polysaccharides to a LPI system (at all ratios) resulted in a significant drop in turbidity over the entire pH range and a shift in net neutrality to lower pH relative to LPI alone; where LPI aggregation was inhibited by repulsive forces between neighbouring polysaccharide chains. As the biopolymer mixing ratio increased, structure formation was less inhibited. The addition of salts resulted in the disruption of formed LPI/polysaccharide complexes.  相似文献   

16.
Polydatin is the main bioactive ingredient in many medicinal plants, such as Hu‐zhang (Polygonum cuspidatum), with many bioactivities. However, its poor aqueous solubility restricts its application in functional food. In this work, 6‐O‐α‐Maltosyl‐β‐cyclodextrin (Malt‐β‐CD), a new kind of β‐CD derivative was used to enhance the aqueous solubility and stability of polydatin by forming the inclusion complex. The phase solubility study showed that polydatin and Malt‐β‐CD could form the complex with the stoichiometric ratio of 1:1. The supermolecular structure of the polydatin/Malt‐β‐CD complex was characterized by ultraviolet–visible spectroscopy (UV), Fourier transform infrared spectroscopy (FT‐IR), X‐ray diffractometry (XRD), thermogravimetric/differential scanning calorimetry (TG/DSC), and proton nuclear magnetic resonance (1H‐NMR) spectroscopy. The changes of the characteristic spectral and thermal properties of polydatin suggested that polydatin could entrap inside the cavity of Malt‐β‐CD. Furthermore, to reasonably understand the complexation mode, the supermolecular structure of polydatin/Malt‐β‐CD inclusion complex was postulated by a molecular docking method based on Autodock 4.2.3. It was clearly observed that the ring B of polydatin oriented toward the narrow rim of Malt‐β‐CD with ring A and glucosyl group practically exposed to the wide rim by hydrogen bonding, which was in a good agreement with the spectral data.  相似文献   

17.
Structural differences in barley grains have been classified as either mealy or steely and their relative proportions have been determined using a light transflectance method in three barley samples varying in the degree of steeliness, Target being the most steely and Chariot most mealy with Blenheim being intermediate. These structural differences were found to be associated with differences in the concentration of endosperm components, particularly proteins and β‐glucan. Analysis of nitrogen within the endosperm showed that protein was mainly concentrated in the embryo and distal regions with the inner, mid‐endosperm containing lowest levels. As the total nitrogen (TN) of the grain increased, the mealier samples accumulated nitrogen mainly in the embryo whereas the steely sample had higher levels in the central endosperm. SDS‐PAGE showed no differences in the protein banding pattern at different TN levels. Electron microscopy using immuno‐gold labelling demonstrated that γ‐hordeins were present in sub‐aleurone and outer endosperm whereas the C‐hordeins were found throughout the central endosperm. However, steely areas of central endosperm contained γ‐hordeins. During malting, protein modification in Chariot was more extensive than in Target with 34kD and 97kD hordeins being completely degraded. In Chariot and Blenheim, level of β‐glucan was low and it was evenly distributed throughout the endosperm. In the steelier Target, however, the amount of β‐glucan was higher and was concentrated in the proximal and distal areas of the endosperm. Steely grains (containing high concentrations of protein and β‐glucan) displayed slower water distribution during steeping and later development and distribution of β‐glucanase during germination. As a consequence, the steely sample achieved a lower degree of modification during malting. The structure of the endosperm, therefore, has a prime influence on the evenness of distribution of moisture and enzymes which is crucial for homogeneous modification during malting. © 1999 Society of Chemical Industry  相似文献   

18.
19.
On solid-like rheological behaviors of globular protein solutions   总被引:1,自引:0,他引:1  
Dynamic viscoelastic and steady flow properties of β-lactoglobulin, bovine serum albumin, ovalbumin, and α-lactalbumin aqueous solutions were investigated at 20°C. When a sinusoidal strain in the linear viscoelastic region was applied, the solutions of the globular proteins except for α-lactalbumin showed typical solid-like rheological behavior: the storage modulus G′ was always larger than the loss modulus G″ in the entire frequency range examined (0.1–100 rad/s). Under a steady shear flow, strong shear thinning behavior was observed with increasing shear rate from 0.001 to 800 s−1, for the globular proteins except for α-lactalbumin. The values of the steady shear viscosity η were lower than those of the dynamic shear viscosity η* at a comparable time scale of observation, violating the Cox–Merz rule, and thus suggesting that a solid-like structure in a globular protein solution was susceptible to a steady shear strain. During isothermal gelation of the protein colloids at 70°C, no crossover between G′ and G″ was observed so that the gelation point was judged by an abrupt increase in the modulus or a sudden decrease in tanδ.  相似文献   

20.
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