首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Sugar beet pectin (SBP) is a marginally utilized co-processing product from sugar production from sugar beets. In this study, the kinetics of oxidative gelation of SBP, taking place via enzyme catalyzed cross-linking of ferulic acid moieties (FA), was studied using small angle oscillatory measurements. The rates of gelation, catalyzed by horseradish peroxidase (HRP) (EC 1.11.1.7) and laccase (EC 1.10.3.2), respectively, were determined by measuring the slope of the increase of the elastic modulus (G′) with time at various enzyme dosages (0.125–2.0 U mL−1). When evaluated at equal enzyme activity dosage levels, the two enzymes produced different gelation kinetics and the resulting gels had different rheological properties: HRP (with addition of H2O2) catalyzed a fast rate of gelation compared to laccase (no H2O2 addition), but laccase catalysis produced stronger gels (higher G′). The main effects and interactions between different factors on the gelation rates and gel properties were examined in response surface designs in which enzyme dosage (0.125–2.0 U mL−1 for HRP; 0.125–10 U mL−1 for laccase), substrate concentration (1.0–4.0%), temperature (25–55 °C), pH (3.5–5.5), and H2O2 (0.1–1.0 mM) (for HRP only) were varied. Gelation rates increased with temperature, substrate concentration, and enzyme dosage; for laccase catalyzed SBP gelation the gel strengths correlated positively with increased gelation rate, whereas no such correlation could be established for HRP catalyzed gelation and at the elevated gelation rates (>100 Pa min−1) gels produced using laccase were stronger (higher G′) than HRP catalyzed gels at similar rates of gelation. Chemical analysis confirmed the formation of ferulic acid dehydrodimers (diFAs) by both enzymes supporting that the gelation was a result of oxidative cross-linking of FAs.  相似文献   

2.
Whey protein isolate (WPI) was subjected to limited tryptic hydrolysis and the effect of the limited hydrolysis on the rheological properties of WPI was examined and compared with those of untreated WPI. At 10% concentration (w/v in 50 mM TES buffer, pH 7.0, containing 50 mM NaCl), both WPI and the enzyme-treated WPI (EWPI) formed heat-induced viscoelastic gels. However, EWPI formed weaker gels (lower storage modulus) than WPI gels. Moreover, a lower gelation point (77 °C) was obtained for EWPI as compared with that of WPI which gelled at 80 °C only after holding 1.4 min. Thermal analysis and aggregation studies indicated that limited proteolysis resulted in changes in the denaturation and aggregation properties. As a consequenece, EWPI formed particulated gels, while WPI formed fine-stranded gels. In keeping with the formation of a particulate gel, Texture Profile Analysis (TPA) of the heat-induced gels (at 80 °C for 30 min) revealed that EWPI gels possessed significantly higher (p < 0.05) cohesiveness, hardness, gumminess, and chewiness but did not fracture at 75% deformation. The results suggest that the domain peptides, especially β-lactoglobulin domains released by the limited proteolysis, were responsible for the altered gelation properties.  相似文献   

3.
This study describes the formation of materials with novel textural characteristics by controlled heteroaggregation of oppositely charged protein-coated lipid droplets. Heteroaggregation was induced by mixing a suspension of β-lactoglobulin (β-Lg)-coated lipid droplets (ζ = −51 mV, d43 ∼ 0.35 μm, 20 wt.%) with a suspension of lactoferrin (LF)-coated lipid droplets (ζ = +32 mV, d43 ∼ 0.35 μm, 20 wt.%) under conditions where the two proteins had opposite charges (pH 7). The mean particle size, flow behaviour, and shear modulus of the materials depended on positive-to-negative particle ratio (0–100%), pH (3–9), ionic strength (0–400 mM), and temperature (30–90 °C). The largest particle sizes, highest viscosities, and largest gel strengths were observed at intermediate particle ratios (40% LF:60% β-Lg), which was attributed to a strong electrostatic attraction between oppositely charged droplets (0 mM NaCl, pH 7, 25 °C). A reduction in particle aggregation, viscosity, and gel strength occurred at intermediate ionic strengths due to screening of the electrostatic attraction between oppositely charged droplets. However, increased aggregation, thickening, and gelation occurred at higher ionic strengths due to screening in electrostatic repulsion between similarly charged droplets. Thermal treatment of the samples (90 °C) promoted a substantial increase in gel strength due to protein denaturation and increased droplet attraction. This study has important implications for the utilisation of controlled particle aggregation to create novel structures in foods, cosmetics, personal care, and other products.  相似文献   

4.
Extensive static and dynamic light scattering (DLS) measurements were done on sodium caseinate solutions as a function of the ionic strength (3–500 mM NaCl), pH (5–8) and temperature (10–70 °C). DLS results were analysed in terms of two populations: the caseinate and a small weight fraction of large particles with a hydrodynamic radius (Rh) of about 65 nm that was independent of the ionic strength, pH and temperature. Caseinate was present as individual molecules at low ionic strength (3 mM), but formed small aggregates (Rh=11 nm) at high ionic strength (>100 mM). The aggregation number (Nagg) increased weakly with decreasing pH between pH 8 and 6, but extensive acid-induced aggregation occurred below pH 5.4 at 250 mM and below pH 6.0 at 3 mM. Nagg increased reversibly with increasing temperature.  相似文献   

5.
刘晶  唐传核 《现代食品科技》2012,28(11):1450-1453
研究了菜豆属球蛋白(芸豆和绿豆)在低pH低离子强度条件下热诱导形成的自组装纤维聚集机理。通过对制备的球蛋白进行热性质及临界形成凝胶浓度分析,选定85℃和0.5%作为自组装纤维化的参数。通过硫代黄色素T(ThT)荧光法,动态光散射(DLS)和原子力显微镜(AFM)表征蛋白纤维化聚集程度及形态。结果表明:芸豆和绿豆球蛋白的Th T最大荧光强度在1h之内剧烈增大(分别是765和1093),表明芸豆和绿豆球蛋白自组装纤维化的"构筑单元"主要产生于加热的起始阶段。DLS结果显示芸豆自组装纤维聚集的能力比绿豆强。但是二者的自组装机理有所不同。AFM图清晰地显示了芸豆球蛋白在加热12 h时自组装形成规则的高度有序的"念珠串状"长线性纤维,绿豆则形成无规则的短棒状纤维和大量碎片状纤维。这为研究进一步研究超低固形物下基于自组装技术的纤维型植物蛋白凝胶的制备提供参考。  相似文献   

6.
Protein gel matrices are responsible for the texture of many foods. Therefore an understanding of the chemical reactions and physical processes associated with fracture properties of gels provides a fundamental understanding of select mechanical properties associated with texture. Globular proteins form thermally induced gels that are classified as fine-stranded, mixed or particulate, based on the protein network appearance. The fundamental properties of true shear stress and true shear strain at fracture, used to describe the physical properties of gels, depend on the gel network. Type and amount of mineral salt in whey protein and β-lactoglobulin protein dispersions determines the type of thermally induced gel matrix that forms, and thus its fracture properties. A fine-stranded matrix is formed when protein suspensions contain monovalent cation (Li+, K+, Rb+, Cs+) chlorides, sodium sulfate or sodium phosphate at ionic strengths ≤0.1 mol/dm3. This matrix has a well-defined network structure, and varies in stress and strain at fracture at different salt concentrations. At ionic strengths >0.1 mol/dm3 the matrix becomes mixed. This network appears as a combination of fine strands and spherical aggregates, and has high stress values and minimum strain values at fracture. Higher concentrations of monovalent cation salts cause the formation of particulate gels, which are high in stress and strain at fracture. The salt concentration required to change microstructure depends on the salt's position in the Hofmeister series. The formation of a particulate matrix also occurs when protein suspensions contain low concentrations (10–20 mol/dm3) of divalent cation (Ca2+, Mg2+, Ba2+) chloride salts or di-cationic 1,6-hexanediamine at pH 7.0. The divalent cation effect on β-lactoglobulin gelation is associated with minor changes in tertiary structure involving amide—amide interproton connectivities (determined by 1H NMR) at 40–45°C, increasing hydrophobicity and intermolecular aggregation. The type of matrix formed appears to be related to the dispersed or aggregated state of proteins prior to denaturation. Mixed and particulate matrices result from conditions which favor aggregation at temperatures (25–45°C) which are much lower than the denaturation temperature (~65°C). Therefore, general (e.g. Hofmeister series) and protein-specific factors can affect the dispersibility of proteins and thereby determine the microstructure and fracture properties of globular protein gels.  相似文献   

7.
To study possible applications of microalgae proteins in foods, a colourless, protein-rich fraction was isolated from Tetraselmis sp. In the present study the emulsion properties of this algae soluble protein isolate (ASPI) were investigated. Droplet size and droplet aggregation of ASPI stabilized oil-in-water emulsions were studied as function of isolate concentration (1.25–10.00 mg/mL), pH (3–7), and ionic strength (NaCl 10–500 mM; CaCl2 0–50 mM). Whey protein isolate (WPI) and gum arabic (GA) were used as reference emulsifiers. The lowest isolate concentrations needed to reach d32 ≤ 1 μm in 30% oil-in-water emulsions were comparable for ASPI (6 mg/mL) and WPI (4 mg/mL). In contrast to WPI stabilized emulsions ASPI stabilized emulsions were stable around pH 5 at low ionic strength (I = 10 mM). Flocculation only occurred around pH 3, the pH with the smallest net droplet ζ-potential. Due to the charge contribution of the anionic polysaccharide fraction present in ASPI its droplet ζ-potential remained negative over the whole pH range investigated. An increase in ionic strength (≥100 mM) led to a broadening of the pH range over which the ASPI stabilized emulsions were unstable. GA emulsions are not prone to droplet aggregation upon changes in pH or ionic strength, but much higher concentrations are needed to produce stable emulsions. Since ASPI allows the formation of stable emulsions in the pH range 5–7 at low protein concentrations, it can offer an efficient natural alternative to existing protein–polysaccharide complexes.  相似文献   

8.
Low temperature cross-linking of denatured whey protein through pH-cycling is proposed to develop nanoparticles with controlled size and properties. Soluble polymers were produced by heating whey protein dispersions at low ionic strength and neutral pH. Nanoparticulation was induced by acidification of diluted polymer dispersions followed by pH neutralization. The effect of aggregation conditions on the physicochemical characteristics and stability of nanoparticles was studied. Nanoparticles with a diameter ranging from 100 to 300 nm were produced depending on the pH of aggregation (5.0, 5.5, 6.0), the added calcium concentration (0, 2.5, 5 mM) and the ageing time at the aggregation pH (0–75 h). The size and the turbidity of nanoparticle dispersions increased with increasing ageing time and calcium concentration. Nanoparticle voluminosity decreased with increasing calcium concentration during pH-cycling, suggesting a more compact and less porous internal structure. The stability of nanoparticles in the presence of different dissociating buffers (EDTA, urea, SDS and DTT) was evaluated and the results showed that whey protein nanoparticles were covalently cross-linked by disulphide bonds.  相似文献   

9.
The objective was to prepare sheared gels of potato protein concentrate and evaluate the effect of pH (3, ~4, ~7), ionic strength (15 or 200 mM) and protein drying conditions (spray or freeze drying) on the final appearance and rheological characteristics. Heat‐set gels 3 % (w/w) at a high ionic strength (200 mM) resulted in an inhomogeneous appearance with presence of clots, while low ionic strength (15 mM) gave homogenous structures. Gels prepared at pH 3 became transparent while preparation above pH 3.0 resulted in high turbidity. Heat treatment and cooling resulted in gelation for all samples except freeze dried powder at pH 3.0. Flow curves during shear from 0.1 to 100 s?1 were fitted by the Herschel–Bulkley model indicating shear thinning behavior for all samples except the freeze dried sample at pH 3 which displayed a Newtonian behavior. Oscillatory measurements after shear indicated viscus behavior (phase angle above 45°) for the spray dried sample at pH 3, and gelled behavior (phase angle above 45°) for the remaining gelled samples. Structure recovery was observed after shear in all samples except at pH 3.0. The data shows potato protein can be used as ingredient in protein beverages.  相似文献   

10.
Casein glycomacropeptide (CMP) found in cheese whey is a C-terminal hydrophilic glycopeptide released from κ-casein by the action of chymosin during cheese making. In a previous work a self-assembly model for CMP at room temperature was proposed, involving a first step of hydrophobic assembly followed by a second step of electrostatic interactions which occurs below pH 4.5. The objective of the present work was to study, by dynamic light scattering (DLS), the effect of heating (35–85 °C) on the pH-driven CMP self-assembly and its impact on the dynamics of CMP gelation. The concentration of CMP was 3% w/w for DLS and 12% w/w for rheological measurements. The solutions at pH 4.5 and 6.5 did not show any change in the particle size distributions upon heating. In contrast the solutions at pH lower than 4.5 that showed electrostatic self-assembly at room temperature were affected by heating. The mean diameter of assembled CMP increased by decreasing pH. For all solutions with pH lower than 4.5, the particle size did not change on cooling, suggesting that the assembled CMP forms formed during heating were stable. The gel point determined as G′–G″ crossover, occurred in all systems at 70 °C, but at different times. The rate of self-assembly determined by DLS as well as the rate of gelation increased with increasing temperature and decreasing pH from 4 to 2. Increasing temperature and decreasing pH, the first step of CMP self-assembly by hydrophobic interactions is speed out. All the self-assembled structures and the gels formed at different temperatures were pH-reversible but did not revert to the initial size (monomer) but to associated forms that correspond mainly to CMP dimers.  相似文献   

11.
The effect of interactions between β-lactoglobulin (β-LG) and dextran sulfate (DS) on thermal stability at near neutral pH was investigated. Samples containing 6% w/w β-LG and DS (Mw = 5–500 kDa) at different biopolymer weight ratios, pH (5.6–6.2), and NaCl concentrations (0–30 mM) were heated at 85 °C for 15 min. Turbidity results showed that the presence of DS at appropriate biopolymer weight ratio and pH significantly lowered the turbidity of heated β-LG. Solutions containing DS:β-LG weight ratios of 0.02 or less showed improved heat stability as indicated by decreased turbidity. Analysis of the unheated mixture by size exclusion chromatography coupled with multi-angle laser light scattering (SEC–MALLS) showed an interaction between β-LG and DS. The size of the aggregates increased as pH decreased. The β-LG–DS aggregates had a greater negative charge as seen from electrophoretic mobility measurement. Addition of 30 mM NaCl inhibited complex formation and the effect of DS on reducing the turbidity of heated β-LG, suggesting that the interaction was electrostatic in nature. Other than charge property, the amount and size of native aggregates appeared to be the major factor in determining how DS altered heat-induced aggregation. The presence of DS decreased denaturation temperature of β-LG, indicating that DS did not improve thermal stability of β-LG by stabilizing its native state but rather by altering its aggregation. The results provide information that will facilitate the application of whey proteins and polysaccharides as functional ingredients in foods and beverages.  相似文献   

12.
Protein–polysaccharide complexes may be used for the development of delivery systems with applications in several industries. In the present work, the interaction of lactoferrin (LF, 0.2wt%) with a High-Methoxyl pectin (0.005–0.15wt%) in aqueous solutions was studied at different pH values (2–7) and temperatures (30–90 °C) at low ionic strength. ζ-potential and light-scattering techniques were used to provide information about the electrical charge and aggregation of individual biopolymers and complexes. At pH 7, the electrical charge went from positive to negative when increasing amounts of pectin were added to the LF solution, which was attributed to the formation of an electrostatic complex. These complexes remained soluble (low turbidity) from pH 7 to 3.5, but became turbid between pH 3.5 and 2, due to charge neutralization and bridging effects. At pH 7, the stability of LF–pectin complexes to aggregation during heating was much better than LF alone. The results of this study should provide information that will facilitate the utilization of lactoferrin as a bioactive component in food systems.  相似文献   

13.
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   

14.
The acid-induced gelation of natural actomyosin (NAM) from burbot (Lota lota) and Atlantic cod (Gardus morhua) added with d-gluconic acid-δ-lactone (GDL) during incubation at room temperature (22–23 °C) for 48 h was investigated. During acidification, pH values of both NAMs reached 4.6 within 48 h. Both NAMs underwent aggregation during acidification as evidenced by increases in turbidity and particle size, especially after 6 h of incubation. The decreases in Ca2+-ATPase activity and salt solubility of both NAMs were observed during incubation. Decreases in total sulphydryl content with the concomitant increases in disulphide bond content of NAM from both species were also noticeable. Additionally, surface hydrophobicity of NAM increased, suggesting the conformational changes in NAM induced by acidification. The storage modulus (G′) values increased with increasing incubation time and G′ development was greater in Atlantic cod NAM, compared with burbot NAM. Differential scanning calorimetry (DSC) revealed that Tmax and enthalpy of myosin peak shifted to the lower values and endothermic peak of actin completely disappeared. In general, gel development was more pronounced in Atlantic cod NAM, compared with the burbot counterpart. As visualised by transmission electron microscopy, network strands of aggregates from Atlantic cod were finer and more uniform than those of the burbot counterpart. Acid-induced gelation of NAM from both fish species therefore involved both denaturation and aggregation processes. However, gelation varied with fish species and had an impact on the resulting gels.  相似文献   

15.
We have investigated the influence of partial hydrolysis with an immobilized proteinase from Bacillus licheniformis on the thermal gelation of isolated beta-lactoglobulin B. Gelation behaviour was determined by dynamic rheological measurements (small deformation) and the gels were characterized with respect to microstructure and water-holding properties. A fine-stranded gel with a complex modulus of approximately 2000 Pa was formed from beta-lactoglobulin (50 g/l in 75 mM-Tris-HCl, pH 7.5). Limited hydrolysis prior to thermal gelation resulted in coarser gels with thicker protein strands and larger pores. Gel structure correlated with its permeability, proton mobility and water-holding capacity. Total stiffness gel increased with low degrees of hydrolysis, but decreased after prolonged hydrolysis. Maximal gel stiffness was 1.5-fold that gels made from of unhydrolysed beta-lactoglobulin. This was much lower than the stiffening effect obtained after partial hydrolysis of whey protein isolate, showing that the gel strengthening effect of partial hydrolysis was depedent on the protein composition and/or the hydrolysis and gelatin conditions. A mechanism to explain the observed effects of hydrolysis on gelation and gel properties is presented.  相似文献   

16.
Addition of CaCl2 to pre-heated whey protein isolate (WPI) suspensions caused an increase in turbidity when pre-heating temperatures were ≥ 64°C. Pre-heating to ≥ 70°C was required for gelation. WPI suspensions which contained CaCl2 became turbid at 45°C and formed thermally induced gels at 66°C. Thermally and Ca2+-induced gels showed significant time/temperature effects but the penetration force values in the Ca2+-induced gels were always lower. However, Ca2+-induced gels were higher in shear stress at fracture. The Ca2+-induced gels had a fine-stranded protein matrix that was more transparent than the thermally induced gels, which showed a particulate microstructure.  相似文献   

17.
Reconstituted skim milk was gelled with a crude protease extract from tamarillo [Cyphomandra betacea or Solanum betacea (syn.)] fruit and compared with gels prepared with calf rennet. The effects of temperature and pH on the gelation of skim milk were investigated by small deformation oscillatory rheology. The tamarillo extract-induced gels had a faster rate of increase in the elastic modulus (G′) at the early stage of gelation than rennet-induced milk gels. This was probably due to the broader proteolytic activity of tamarillo protease extracts as shown by sodium dodecyl sulfate–PAGE analysis. Confocal microscopy also showed that the milk gels resulting from the addition of tamarillo extracts had larger voids than rennet-induced milk gels. The proteolytic activity of tamarillo extracts was found to be optimal at pH 11. For both rennet and tamarillo extracts, the aggregation time was similar between pH 6.7 and 6.5, but the aggregation time of rennet-induced milk gels was lower than that of milk gels obtained by the addition of tamarillo extracts at pH lower than 6.5. An increase in temperature was found to have a significant effect on aggregation time, particularly at 20°C, where rennet did not coagulate milk in 3 h but the tamarillo extracts coagulated milk within 2 h. The results of this study suggest that extracts from tamarillo fruit could be used for milk gelation, particularly under lower temperature or high pH conditions.  相似文献   

18.
Turbid solutions and fine-stranded gels of myosin from bovine semi-membranous muscle were investigated by transmission and scanning electron microscopy. Evidence is given that the turbidity was caused by filament formation upon dialysis to pH 5.5 and 0.25 M KCl and to pH 4.0 and 0.6 m KCl at 4°C. The filaments formed at pH5.5 and 0.25 m KCl had a backbone with a diameter of c. 25 nm with the myosin heads located close to the filament backbone. The total width of these filaments was c.45nm. The filaments were prepared for electron microscopy by adsorption on various substrates, negative staining, or freeze drying and rotary shadowing. Variations in the preparation technique did not affect the appearance of the filaments. The filaments formed at pH 4.0 and 0.6 m KCl had a more irregular appearance, and the total filament width varied between 20 and 45 nm. Fringes of globular material surrounding the filament backbone were seen but also clusters of myosin molecules protruding further out from the backbone and from the filament ends. Comparison of heat-treated filaments in dilute solutions with strands of the gel network confirmed that the gel strands originated from filaments formed upon dialysis prior to gelation. Typical features of the network structure were junction zones formed by parallel alignments of filaments in pairs and by end-to-side interactions forming so-called Y-junctions. At pH5.5 and 0.25 m KCl these interactions resulted in a rather loose and open network structure. At pH 4.0 and 0.6 M KCl the filaments often interacted approximately at right angles, which resulted in a denser network than that observed atpH5.5 and 0.25 m KCl. The efficient network formation at pH 4.0 gave rise to spontaneous gel formation upon dialysis without any heat treatment. Additional heating did not change the character of the network, and no differences could be observed between unheated and heat-treated gels at low magnifications. At higher magnifications it could be seen that heating resulted in loss of details of the filaments at both pH values and ionic strengths. The shape of the myosin heads was lost, and the heads fused together on the filament backbone.  相似文献   

19.
Effect of pH on the gel properties and secondary structure of fish myosin   总被引:3,自引:0,他引:3  
The relationships between gel properties and the secondary structures of silver carp myosin were investigated at pH 5.5–9.0 using dynamic rheological measurement, circular dichroism and scanning electron microscopy. The gel properties of fish myosin were strongly pH and temperature dependent. During heating at 1 °C/min, myosin formed gels in the pH range 5.5–7.5, but not at pH 8.0–9.0. α-Helix was the predominant structure at pH 7.0. The α-helix fraction declined with increasing temperature and the pH away from 7.0, whilst the other secondary structure fractions increased. The α-helix structure of myosin was more susceptive to acid-treatment than alkali-treatment. As pH increased, the gelation rate and gel strength decreased, and the water-holding capacity (WHC) showed an increasing trend followed by a plateau. High β-sheet and β-turn fractions prior to heating could improve G′ at 90 °C, but they depressed the WHC. A compact and uniform gel of fish myosin was obtained at pH 7.0.  相似文献   

20.
Biopolymer nanoparticles can be formed by heating globular protein/polysaccharide mixtures above the thermal denaturation temperature of the protein under pH conditions where the two biopolymers are weakly electrically attracted to each other. In this study, the influence of polysaccharide linear charge density on the formation and properties of these biopolymer nanoparticles was examined. Mixed solutions of globular proteins (β-lactoglobulin) and anionic polysaccharides (high and low methoxyl pectin) were prepared. Micro-electrophoresis, dynamic light scattering, turbidity and atomic force microscopy (AFM) measurements were used to determine the influence of protein-to-polysaccharide mass ratio (r), solution pH, and heat treatment on biopolymer particle formation. Biopolymer nanoparticles (d < 500 nm) could be formed by heating protein–polysaccharide complexes at 83 °C for 15 min at pH 4.75 and r = 2:1 in the absence of added salt. The biopolymer particles formed were then subjected to pH and salt adjustment to determine their stability. The pH stability was greater for β-lactoglobulin-HMP complexes than for β-lactoglobulin-LMP complexes. The addition of 200 mM sodium chloride to heated complexes greatly improved the pH stability of HMP complexes, but decreased the pH stability of LMP complexes. The biopolymer particles formed consisted primarily of β-lactoglobulin, which was probably surrounded by a pectin coating at low pH values. AFM measurements indicated that the biopolymer nanoparticles formed were spheroid in shape. These biopolymer particles may be useful as delivery systems or fat mimetics.  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