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.
Whey proteins (WP) gelation process with addition of Arabic gum (AG) was studied. Two different driving processes were employed to induce gelation: (1) heating of 12% whey protein isolate (WPI) solutions (w/w) or (2) acidification of previous thermal denatured WPI solutions (5% w/w) with glucono-δ-lactone (GDL). Protein concentrations were different because they were minimal to form gel in these two processes, but denaturation conditions were the same (90 °C/30 min). Water-holding capacity and mechanical properties of the gels were evaluated. The BST equation was used to evaluate the nonlinear part of the stress–strain data. Cold-set gels were weaker than heat-set gels at the pH range near the isoelectric point (pI) of the main whey proteins, but heated gels were more deformable (did not exhibit rupture point) and showed greater elasticity modulus. However, gels formed by heating far from the pI (pH 6.7 or 3.5) showed more fragile structure, indicating that, in these mixed gels, there are prevailing biopolymers interactions. Cold-set and heat-set gels at pH near or below the WP pI showed strain-weakening behavior, but heated gels at neutral pH showed strong strain-hardening behavior. Such results suggest that differences in stress–strain curve at the nonlinear part of the data could be correlated to structure particularities obtained from different gelation processes.  相似文献   

2.
The effects of pH (6.7 or 5.8), protein concentration and the heat treatment conditions (70 or 90 °C) on the physical properties of heat-induced milk protein gels were studied using uniaxial compression, scanning electron microscopy, differential scanning calorimetry, and water-holding capacity measurements. The systems were formed from whey protein isolate (10–15% w/v) with (5% w/v) or without the addition of caseinate. The reduction in pH from 6.7 to 5.8 increased the denaturation temperature of the whey proteins, which directly affected the gel structure and mechanical properties. Due to this increase in the denaturation temperature of the β-lactoglobulin and α-lactalbumin, a heat treatment of 70 °C/30 min did not provide sufficient protein unfolding to form self-supporting gels. However, the presence of 5% (w/v) sodium caseinate decreased the whey protein thermo stability and was essential for the formation of self-supporting gels at pH 6.7 with heat treatment at 70 °C/30 min. The gels formed at pH 6.7 showed a fine-stranded structure, with great rigidity and deformability as compared to those formed at pH 5.8. The latter had a particulate structure and exuded water, which did not occur with the gels formed at pH 6.7. The addition of sodium caseinate led to less porous networks with increased gel deformability and strength but decreased water exudation. The same tendencies were observed with increasing whey protein concentration.  相似文献   

3.
Stability of oil-in-water emulsions during freezing and thawing is regulated by the phase transitions occurring in the continuous and dispersed phases upon thermal treatments and by the composition of the interfacial membrane. In the present study, the impact of the water phase formulation (0–2.5–5–10–20–30–40% w/w sucrose), the interfacial composition [whey protein isolates (WPI) or sodium caseinate (NaCas) used at different concentrations], and the particle size on the stability of hydrogenated palm kernel oil (30% w/w)-in-water systems was investigated. Phase/state behaviour of the continuous and dispersed phases and emulsion destabilisation were studied by differential scanning calorimetry. System morphology was observed by particle size analysis and optical microscopy. The presence of sucrose in the aqueous phase and reduced particle size distribution significantly improved emulsion stability. WPI showed better stabilising properties than NaCas at lipid to protein ratios of 10:1, 7.5:1, 5:1 and 4:1. Increased WPI concentration significantly improved emulsion resistance to breakdown during freeze–thaw cycling. NaCas showed poor stabilising properties and was ineffective in reducing emulsion destabilisation at 0% sucrose at all the lipid to protein ratios.  相似文献   

4.
A mechanism describing the denaturation and aggregation behavior during heat-treatment of pure β-lactoglobulin and β-lactoglobulin in whey protein isolate (WPI) under selected conditions (20 to 90 gL−1 in water at pH 7.0, 78 °C) is presented. A combination of reversed-phase and gel permeation chromatography was used to study the disappearance of native β-lactoglobulin and the formation of non-native intermediates in the aggregation process. The mean reaction order for pure β-lactoglobulin and β-lactoglobulin in WPI were the same, 1.4. While the rate of β-lactoglobulin denaturation was greater in WPI there was less aggregation compared to that of pure β-lactoglobulin. More of the β-lactoglobulin in WPI remained in a non-native monomer intermediate state after 30 min of heating. After an initial lag period, during which non-native monomers appeared, aggregates formed and rapidly reached a plateau in terms of their size. These aggregates were visualized using atomic force microscopy. There was no significant effect of protein concentration on either aggregate size or the number of exposed sulfhydryls in the heated solutions.  相似文献   

5.
The suitability of water-in-oil-in-water multiple emulsions to encapsulate resveratrol was assessed. Multiple emulsions were prepared by emulsifying a primary emulsion (40 wt.%) in water containing 0.5 wt.% sodium caseinate and 0.1 M NaCl. Four primary emulsions of canola oil (20 wt.%) stabilized by 8 wt.% polyglycerol polyricinoleate were chosen. The dispersed phase of the primary emulsions contained 0.1 M NaCl and either water, 20 wt.% ethanol in water, 2.5 wt.% whey protein isolate (WPI) in water, or 2.5 wt.% WPI and 5 wt.% gelatine in water. Resveratrol was incorporated into these primary emulsions at 0.25 wt.% to give a final 0.02 wt.% resveratrol in the multiple emulsions. Slight increase in particle size with storage at 23 °C for up to 2 weeks was observed. Further, less than 10% of the total encapsulated resveratrol is released to the external, continuous, aqueous phase. This work demonstrates the potential of multiple emulsions to encapsulate resveratrol for food applications.  相似文献   

6.
Power O  Hallihan A  Jakeman P 《Amino acids》2009,37(2):333-339
The insulinotropic response to the ingestion of whey protein and whey protein hydrolysate, independent of carbohydrate, is not known. This study examined the effect of protein hydrolysis on the insulinotropic response to the ingestion of whey protein. Sixteen healthy males ingested a 500 mL solution containing either 45 g of whey protein (WPI) or whey protein hydrolysate (WPH). The estimated rate of gastric empting was not altered by hydrolysis of the protein [18 (3) vs. 23 (3) min, n = 16; P = 0.15]. Maximum plasma insulin concentration (C max) occurred later (40 vs. 60 min) and was 28% [234 (26) vs. 299 (31) mM, P = 0.018] greater following ingestion of the WPH compared to the WPI leading to a 43% increase [7.6 (0.9) vs. 10.8 (2.6) nM, P = 0.21] in the AUC of insulin for the WPH. Of the amino acids with known insulinotropic properties only Phe demonstrated a significantly greater maximal concentration [C max; 65 (2) vs. 72 (3) μM, n = 16; P = 0.01] and increase (+22%) in AUC following ingestion of the WPH. In conclusion, ingestion of whey protein is an effective insulin secretagogue. Hydrolysis of whey protein prior to ingestion augments the maximal insulin concentration by a mechanism that is unrelated to gastric emptying of the peptide solution.  相似文献   

7.
The phase separation behavior of whey protein isolate (WPI) aggregates and κ-carrageenan (κ-car) mixtures was studied using the Vrij's theory and image analysis method. The intrinsic parameter (molecular mass and radius of gyration) for κ-car and the WPI aggregates was determined using intrinsic viscosity and reduced viscosity of each biopolymer. Confocal microscopy observations revealed the appearance of protein aggregate domains when phase separation occurred, with microgel droplets of WPI included in a continuous κ-car phase. The occurrence of aggregate droplet has not been reported before for the phase-separating WPI/κ-car mixtures. So far, network emulsion-like microstructures have been observed with WPI in a network structure. By using different WPI concentrations (4% or 6%), the microstructure of the systems changes while increasing the κ-car concentration. The size of the microgels (1–2.5 μm) depends on both κ-car and WPI concentration. Confocal microscopy combined with image analysis (method of the variance) was used effectively as objective means to determine the phase boundary of the phase-separating systems. Additional information on the depletion layer thickness, Δ, was obtained using self-consistent field theory. The results show that Δ has a constant value of 80.5 nm for ck - car \prec 2 g/l {{\hbox{c}}_{\kappa {\rm{ - car}}}} \prec {\hbox{2 g}}/{l} , in agreement with ∆ ≈ R g (radius of gyration). Above this concentration, Δ decreases as a function of κ-car concentration. The experimental phase boundary was well predicted using Vrij's theory. This work showed a new approach to generate phase diagrams (e.g., under shear) of phase-separating systems.  相似文献   

8.
The present study attempts to characterize the effect of shear rate on the composition, size, and molecular weight of the protein aggregates present in the upper layer after phase separation of 5% whey protein isolate (WPI) mixed with 0.5% κ-carrageenan (κ-car) at pH 7.0. The mixtures were heated and sheared under different shearing rates. Size exclusion chromatography (SEC), dynamic light scattering, and static light scattering were employed to describe the effect of shear rate on the size and molecular mass of WPI aggregates. At the molecular level, the size of the aggregates increased with an increase in shear rate. Shear rate also caused a decrease in turbidity of the upper layer after centrifugation. SEC combined with multi-angle laser light scattering showed that the WPI aggregates molecular mass was between 106and 107 g/mol when the shear rate increased from 3.6 to 86.4 s−1. Two empirical models described well the effect of shear rate on the size of WPI aggregates, and both models gave comparable results. By varying process parameters such as flow behavior and temperature, it is possible to control WPI aggregation and, thus, obtain aggregates with a range of different characteristics (size).  相似文献   

9.
The objectives of this study were to investigate the moisture-induced protein aggregation of whey protein powders and to elucidate the relationship of protein stability with respect to water content and glass transition. Three whey protein powder types were studied: whey protein isolate (WPI), whey protein hydrolysates (WPH), and beta-lactoglobulin (BLG). The water sorption isotherms were determined at 23 and 45°C, and they fit the Guggenheim–Andersson–DeBoer (GAB) model well. Glass transition was determined by differential scanning calorimeter (DSC). The heat capacity changes of WPI and BLG during glass transition were small (0.1 to 0.2 Jg−1 °C−1), and the glass transition temperature (T g) could not be detected for all samples. An increase in water content in the range of 7 to 16% caused a decrease in T g from 119 down to 75°C for WPI, and a decrease from 93 to 47°C for WPH. Protein aggregation after 2 weeks’ storage was measured by the increase in insoluble aggregates and change in soluble protein fractions. For WPI and BLG, no protein aggregation was observed over the range of 0 to 85% RH, whereas for WPH, ∼50% of proteins became insoluble after storage at 23°C and 85% RH or at 45°C and ≥73% RH, caused mainly by the formation of intermolecular disulfide bonds. This suggests that, at increased water content, a decrease in the T g of whey protein powders results in a dramatic increase in the mobility of protein molecules, leading to protein aggregation in short-term storage.  相似文献   

10.
Bovine lactoferrin was enriched in various whey samples by affinity chromatography using immobilized gangliosides. Bovine gangliosides were isolated from fresh buttermilk using a combination of ultrafiltration and organic extraction. Isolated gangliosides were covalently immobilized onto controlled-pore glass beads. The immobilized matrix contained 66 micrograms of gangliosides per gram of beads. After loading the matrix with reconstituted whey protein isolate (WPI) or whey protein concentrate (WPC), the matrix was washed with sodium phosphate buffer (pH 7) followed by sodium acetate buffer (pH 4) before elution of lactoferrin with 1 M NaCl in sodium acetate buffer. From the intensities of the protein bands in SDS-PAGE, lactoferrin constituted a minimum of 40% of the total protein in the salt eluted sample. WPI, pretrated by heating and ultrafiltration, showed the highest lactoferrin purity among protein sources, while WPI (10% wt/vol) showed the highest recovery. These results show that immobilized gangliosides can be used to enrich the lactoferrin content of whey.  相似文献   

11.
The paper reports a study involving the use of Halomonas boliviensis, a moderate halophile, for co-production of compatible solute ectoine and biopolyester poly(3-hydroxybutyrate) (PHB) in a process comprising two fed-batch cultures. Initial investigations on the growth of the organism in a medium with varying NaCl concentrations showed the highest level of intracellular accumulation of ectoine (0.74 g L−1) at 10–15% (w/v) NaCl, while at 15% (w/v) NaCl, the presence of hydroxyectoine (50 mg L−1) was also noted. On the other hand, the maximum cell dry weight and PHB concentration of 10 and 5.8 g L−1, respectively, were obtained at 5–7.5% (w/v) NaCl. A process comprising two fed-batch cultivations was developed—the first culture aimed at obtaining high cell mass and the second for achieving high yields of ectoine and PHB. In the first fed-batch culture, H. boliviensis was grown in a medium with 4.5% (w/v) NaCl and sufficient levels of monosodium glutamate, NH4+, and PO43−. In the second fed-batch culture, the NaCl concentration was increased to 7.5% (w/v) to trigger ectoine synthesis, while nitrogen and phosphorus sources were fed only during the first 3 h and then stopped to favor PHB accumulation. The process resulted in PHB yield of 68.5 wt.% of cell dry weight and volumetric productivity of about 1 g L−1 h−1 and ectoine concentration, content, and volumetric productivity of 4.3 g L−1, 7.2 wt.%, and 2.8 g L−1 day−1, respectively. At salt concentration of 12.5% (w/v) during the second cultivation, the ectoine content was increased to 17 wt.% and productivity to 3.4 g L−1 day−1.  相似文献   

12.
The enzymatic cross-linking of adsorbed biopolymer nanoparticles formed between whey protein isolate (WPI) and sugar beet pectin using the complex coacervation method was investigated. A sequential electrostatic depositioning process was used to prepare emulsions containing oil droplets stabilized by WPI – nanoparticle – membranes. Firstly, a finely dispersed primary emulsion (10 % w/w miglyol oil, 1 % w/w WPI, 10 mM acetate buffer at pH 4) was produced using a high-pressure homogenizer. Secondly, a series of biopolymer particles were formed by mixing WPI (0.5 % w/w) and pectin (0.25 % w/w) solutions with subsequent heating above the thermal denaturation temperature (85 °C, 20 min) to prepare dispersions containing particles in the submicron range. Thirdly, nanoparticle-covered emulsions were formed by diluting the primary emulsion into coacervate solutions (0–0.675 % w/w) to coat the droplets. Oil droplets of stable emulsions with different interfacial membrane compositions were subjected to enzymatic cross-linking. We used cross-linked multilayered emulsions as a comparison. The pH stability of primary emulsions, biopolymer complexes and nanoparticle-coated base emulsions, as well as multilayered emulsions, was determined before and after enzyme addition. Freeze-thaw stability (?9 °C for 22 h, 25 °C for 2 h) of nanoparticle-coated emulsions was not affected by laccase. Results indicated that cross-linking occurred exclusively in the multilamellar layers and not between adsorbed biopolymer nanoparticles. Results suggest that the accessibility of distinct structures may play a key role for biopolymer-cross-linking enzymes.  相似文献   

13.

This study investigated the influence of thermal treatment (30 °C to 110 °C, 30 min) on the physicochemical and rheological properties of an emulsion stabilized by black tilapia (Oreochromis mossambicus) skin at pH 4. The protein pattern of tilapia gelatin did not have any significant difference after the gelatin was heated within a temperature range of 30 °C to 70 °C. However, at 90 °C and 110 °C, denaturation occurred where α-, β- and γ-chains of the gelatin were degraded, leading to a concomitant increase in low molecular peptides. The emulsion stability was investigated through a particle size analyzer, zeta potential, microscopic observation and creaming index. The gelatin emulsion was physically stable at 30 °C to 70 °C with a mean droplet size of less than 13 μm. When the heating temperature was increased to 90 °C and 110 °C, the emulsion showed a pronounced increase in droplet size due to coalescence. The gelatin emulsion heated at 90 °C and 110 °C also displayed instability against creaming after storage at room temperature for 7 days. As the heating temperature increased, the gelatin emulsion exhibited a decrease in apparent viscosity and the flow behavior changed from shear thinning to Newtonian. The rheological data also showed that the storage modulus (G′) of emulsion became more frequency dependent as the heating temperature increased, indicating weak droplet interactions. The tilapia gelatin emulsion was physically unstable when subjected to thermal treatment above 70 °C. The data reported in this study provides useful insight into the formulation of acidic food emulsions that require thermal treatment.

  相似文献   

14.
Loveday SM  Su J  Rao MA  Anema SG  Singh H 《Biomacromolecules》2011,12(10):3780-3788
Self-assembly of amyloid-like nanofibrils during heating of bovine whey proteins at 80 °C and pH 2 is accelerated by the presence of NaCl and/or CaCl(2), but the rheological consequences of accelerated self-assembly are largely unknown. This investigation focused on the impact of CaCl(2) on the evolution of rheological properties and fibril morphology of heated whey protein isolate (WPI), both during self-assembly at high temperature and after cooling. Continuous rotational rheometry of heated 2% w/w WPI showed a nonlinear effect of CaCl(2) on the viscosity of fibril dispersions, which we attributed to effects on fibril flexibility and thus the balance between intrafibril and interfibril entanglements. Small-amplitude oscillatory measurements made in situ during heating of 10% w/w WPI at 80 °C suggest that CaCl(2) is not involved in either fibril structure or gel structure, and this was confirmed with dialysis experiments.  相似文献   

15.

Freeze-dried banana powder represents an ideal source of nutrients and has not yet been used for probiotic incorporation. In this study, microencapsulation by freeze drying of probiotics Lactobacillus acidophilus and Lactobacillus casei was made using whey protein isolate (WPI), fructooligosaccharides (FOS), and their combination (WPI + FOS) at ratio (1:1). Higher encapsulation yield was found for (WPI + FOS) microspheres (98%). Further, microcapsules of (WPI + FOS) were used to produce a freeze-dried banana powder which was analyzed for bacterial viability under simulated gastrointestinal fluid (SGIF), stability during storage at 4 °C and 25 °C, and chemical and sensory properties. Results revealed that (WPI + FOS) microcapsules significantly increased bacteria stability in the product over 30 days of storage at 4 °C averaging (≥ 8.57 log CFU/g) for L. acidophilus and (≥ 7.61 log CFU/g) for L. Casei as compared to free cells. Bacteria encapsulated in microspheres (WPI + FOS) were not significantly affected by the SGIF, remaining stable up to 7.05 ± 0.1 log CFU/g for L.acidophilus and 5.48 ± 0.1 log CFU/g for L.casei after 90 min of incubation at pH 2 compared to free cells which showed minimal survival. Overall, encapsulated probiotics enriched freeze-dried banana powders received good sensory scores; they can therefore serve as safe probiotics food carriers.

  相似文献   

16.
Ethyl cellulose microcapsules were developed for use as a drug-delivery device for protecting folic acid from release and degradation in the undesirable environmental conditions of the stomach, whilst allowing its release in the intestinal tract to make it available for absorption. The controlled release folic acid-loaded ethyl cellulose microcapsules were prepared by oil-in-oil emulsion solvent evaporation using a mixed solvent system, consisting of a 9:1 (v/v) ratio of acetone:methanol and light liquid paraffin as the dispersed and continuous phase. Span 80 was used as the surfactant to stabilize the emulsion. Scanning electron microscopy revealed that the microcapsules had a spherical shape. However, the particulate properties and in vitro release profile depended on the concentrations of the ethyl cellulose, Span 80 emulsifier, sucrose (pore inducer), and folic acid. The average diameter of the microcapsules increased from 300 to 448 μm, whilst the folic acid release rate decreased from 52% to 40%, as the ethyl cellulose concentration was increased from 2.5% to 7.5% (w/v). Increasing the Span 80 concentration from 1% to 4% (v/v) decreased the average diameter of microcapsules from 300 to 141 μm and increased the folic acid release rate from 52% to 79%. The addition of 2.5–7.5% (w/v) of sucrose improved the folic acid release from the microcapsules. The entrapment efficiency was improved from 64% to 88% when the initial folic acid concentration was increased from 1 to 3 mg/ml.  相似文献   

17.
Byproduct utilization is an important consideration in the development of sustainable processes. Whey protein isolate (WPI), a byproduct of the cheese industry, and gelatin, a byproduct of the leather industry, were reacted individually and in blends with microbial transglutaminase (mTGase) at pH 7.5 and 45 °C. When a WPI (10% w/w) solution was treated with mTGase (10 U/g) under reducing conditions, the viscosity increased four-fold and the storage modulus (G′) from 0 to 300 Pa over 20 h. Similar treatment of dilute gelatin solutions (0.5–3%) had little effect. Addition of gelatin to 10% WPI caused a synergistic increase in both viscosity and G′, with the formation of gels at concentrations greater than 1.5% added gelatin. These results suggest that new biopolymers, with improved functionality, could be developed by mTGase treatment of protein blends containing small amounts of gelatin with the less expensive whey protein.  相似文献   

18.
The thermal properties of cowpea protein isolates (CPI) were studied by differential scanning calorimetry under the influence of various conditions. An increase in the pH of protein extraction, from 8.0 to 10.0, during CPI preparation promoted a partial denaturation of cowpea proteins. Increases in enthalpy change of denaturation (ΔH) and temperature of denaturation (Td) were detected with increasing protein concentration from 7.5 to 10.5% (w/w). This behavior suggests that denaturation involves a first step of dissociation of protein aggregates. Calcium induced thermal stabilization in cowpea proteins, the increase in Td was ca. 0.3 °C/mM for protein dispersions of 7.5% (w/w) for 0 to 40 mM CaCl2. High hydrostatic pressure (HHP) induced denaturation in CPI in a pressure level dependent manner. The presence of calcium protected cowpea proteins towards HHP-induced denaturation when pressure level was 400 MPa, but not when it was 600 MPa. Thermal properties of cowpea protein isolates were very sensitive to processing conditions, these behaviors would have implications in processing of CPI-containing foodstuff.  相似文献   

19.
Stability of whey protein-pectin complexes is an essential criterion for their application in different food matrices. The impact of process parameters on micro- and macro-structural characteristics of thermally stabilised whey protein-pectin complexes was investigated using fluorescence spectroscopy, ζ-potential measurements, dynamic light scattering and phase separation. Complexes prepared from whey protein isolate (WPI) and pectins with different degrees of esterification (HMP, LMP) were generated at different biopolymer concentrations (WPI + pectin: 5.0 % + 1.0 %, c h i g h ; 2.75 % + 0.55 %, c m e d ; 0.5 % + 0.1 %, c l o w ), heating temperatures (80-90°C) and pH levels (6.1-4.0). Micro- and macro-structural characteristics of the complexes depended on concentration level and degree of esterification, with complexes being more sensitive towards environmental changes at c l o w than at c m e d and c h i g h . WPI-LMP complexes exhibited sizes <1 μm suitable for micro-encapsulation, whereas WPI-HMP complexes at c m e d achieved sizes from 1-10 μm and at c h i g h from 10-200 μm underlining their potential as fat-replacers and structuring agents, respectively. Slopes and intercepts derived from intensity ratios of fluorescence spectra gave insights into the state of unfolding of β-lactoglobulin within the complexes and thus about the protective effect of pectin addition.  相似文献   

20.
The growth rate of abalone post larvae of Haliotis rufescens fed ad libitum with a benthic monoalgal diatom culture maintained as monocultures on a semi-commercial scale, was evaluated and correlated with the biochemical composition of the diatoms. The cell size (7.0 × 4.0 μm to 21.0 × 7.5 μm), protein percentage (7.42% to 13.66%), and ash content (49.03% to 59.61%) were different among diatom strains; lipid percentage, nitrogen free extract, and energy content (Kcal g−1) were similar among diatom strains. The values of essential and non-essential amino and fatty acids composition differed among diatom strains. Differences in the abalone shell length and orthogonal analyses revealed postlarval growth was dependent on the quality of the food source. Postlarvae abalone displaying the longest shell lengths were fed Nitzschia thermalis var. minor and Amphiprora paludosa var. hyalina (1,712.0 ± 61 μm and 1,709 ± 67 μm, respectively), followed by Navicula incerta (1,413.3 ± 43 μm). The fatty acid content of benthic diatoms and abalone growth rate were not correlated.  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