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 656 毫秒
1.
The aim of this study was to produce pilot scale batches of β-casein concentrate and micellar casein concentrate with reduced β-casein level. The isolation of β-casein was done using membrane filtration at cold temperatures (≤5 °C). A micellar casein concentrate was obtained from skim milk by means of warm microfiltration (MF) at 50 °C (0.1 μm pore size, ceramic membranes). The concentrate was stored at 2–3 °C for approximately 40 h to induce temperature-dependent dissociation of β-casein from casein micelles. β-casein was separated from the cold-stored concentrate using MF (0.3 μm pore size, organic membranes) at ≤5 °C. β-Casein permeate was warmed up to 50 °C to lead self-association of β-casein micelles before ultrafiltration at 50 °C (10 kDa cut-off, organic membranes). Two streams, a β-casein-depleted and a β-casein concentrate, were generated. A purity of 92.64% and a yield of up to 18.07% were achieved for β-casein.  相似文献   

2.
Milk was processed with high hydrostatic pressure in order to modify the casein micelles. Images, that in details showed the casein micelle structure in untreated and pressure-treated skim milk, were obtained by using cryo-transmission electron microscopy (cryo-TEM). Sizes and shapes adopted by casein micelles in pressurised milk are concluded to be a result of an equilibrium distribution between self-assembling casein molecules in the serum phase and caseins adsorbed to surfaces of casein micelles and are governed by an initial pressure-dependent displacement of caseins into the serum phase. Pressurisation of milk at moderately high pressure, in the range 150–300 MPa, favoured formation of a large number of small micelles that coexisted with a fraction of large micelles, and which appeared perfectly spherical with smooth and well-defined surfaces, features which are suggested to originate from secondary adsorption of caseins. Pressurisation of milk at 400 MPa favoured formation of smaller casein assemblies, with sizes between 30 nm and 100 nm. Measurements of free calcium concentration [Ca2+] showed that calcium was rebound to casein micelles after pressurisation of milk. Furthermore, the electron microscopy images indicated that the substructures were similar for pressure-modified casein micelles and casein micelles in untreated milk.  相似文献   

3.
An indigenous inhibitor in raw milk inhibits cross-linking by transglutaminase (TG). The enzymatic cross-linking of micellar casein, compared with sodium caseinate, taking thermal inactivation of the TG inhibitor in the milk serum into consideration, was investigated. Inhibitor-free micellar casein was prepared by membrane separation combined with heat treatment of the UF permeate. The inhibitor permeated through MF (nominal pore size 0.1 μm) and UF (cutoff 25 kDa) membranes. TG-catalyzed cross-linking of casein micelles was clearly enhanced by UHT-treatment of UF permeate. Variation of the enzyme concentration showed that the inhibitory effect could not be compensated by higher enzyme concentrations when the casein micelles were suspended in unheated milk serum. Sodium caseinate, however, underwent high degrees of cross-linking even in unheated milk serum. By mixing an unheated milk serum and a UHT-treated milk serum at different ratios, the relative TG inhibitor activity was analysed. High inactivation (>80%) of the TG inhibitor is necessary to achieve high degrees of protein cross-linking.  相似文献   

4.
The objectives of this study were to investigate the effect of the solvent on the accuracy of casein micelle particle size determination by dynamic light scattering (DLS) at different temperatures and to establish a clear protocol for these measurements. Dynamic light scattering analyses were performed at 6, 20, and 50°C using a 90Plus Nanoparticle Size Analyzer (Brookhaven Instruments, Holtsville, NY). Raw and pasteurized skim milk were used as sources of casein micelles. Simulated milk ultrafiltrate, ultrafiltered water, and permeate obtained by ultrafiltration of skim milk using a 10-kDa cutoff membrane were used as solvents. The pH, ionic concentration, refractive index, and viscosity of all solvents were determined. The solvents were evaluated by DLS to ensure that they did not have a significant influence on the results of the particle size measurements. Experimental protocols were developed for accurate measurement of particle sizes in all solvents and experimental conditions. All measurements had good reproducibility, with coefficients of variation below 5%. Both the solvent and the temperature had a significant effect on the measured effective diameter of the casein micelles. When ultrafiltered permeate was used as a solvent, the particle size and polydispersity of casein micelles decreased as temperature increased. The effective diameter of casein micelles from raw skim milk diluted with ultrafiltered permeate was 176.4 ± 5.3 nm at 6°C, 177.4 ± 1.9 nm at 20°C, and 137.3 ± 2.7 nm at 50°C. This trend was justified by the increased strength of hydrophobic bonds with increasing temperature. Overall, the results of this study suggest that the most suitable solvent for the DLS analyses of casein micelles was casein-depleted ultrafiltered permeate. Dilution with water led to micelle dissociation, which significantly affected the DLS measurements, especially at 6 and 20°C. Simulated milk ultrafiltrate seemed to give accurate results only at 20°C. Results obtained in simulated milk ultrafiltrate at 6°C could not be explained based on the known effects of temperature on the casein micelle, whereas at 50°C, precipitation of amorphous calcium phosphate affected the DLS measurement.  相似文献   

5.
The protein composition of bovine skim milk was modified using pilot scale membrane filtration to produce a whey protein-dominant ingredient with a casein profile closer to human milk. Bovine skim milk was processed at low (8.9 °C) or high (50 °C) temperature using ceramic microfiltration (MF) membranes (0.1 μm mean pore diameter). The resulting permeate stream was concentrated using polyethersulfone ultrafiltration (UF) membranes (10 kDa cut-off). The protein profile of MF and UF retentate streams were determined using reversed phase-high performance liquid chromatography and polyacrylamide gel electrophoresis. Permeate from the cold MF process (8.9 °C) had a casein:whey protein ratio of ∼35:65 with no αS- or κ-casein present, compared with a casein:whey protein ratio of ∼10:90 at 50 °C. This study has demonstrated the application of cold membrane filtration (8.9 °C) at pilot scale to produce a dairy ingredient with a protein profile closer to that of human milk.  相似文献   

6.
By studying the hydration of casein micelles using a variety of techniques, a distinction could be made between water that appeared bound by the protein (∼0.5 g g−1 protein), water associated with the κ-casein brush (∼1.0 g g−1 protein) and water entrapped in the casein micelles (∼1.8 g g−1 protein), yielding a total micellar hydration of ∼3.3 g g−1 protein, in line with casein micelle voluminosity derived from intrinsic viscosity measurements. For caseinate particles, however, the main contributor to intrinsic viscosity was not protein hydration but the non-spherical particle shape. These non-spherical particles in caseinate are likely to be naturally present as primary casein particles (PCP) in casein micelles. PCP could be used to build casein micelles by controlled introduction of micellar salts. Based on the findings of this study, casein micelles could be described as a porous network of non-spherical PCP linked by calcium phosphate nanoclusters.  相似文献   

7.
The objective of this study was to investigate the effects of whey protein denaturation and whey protein:casein-ratio on the structural, rheological and sensory properties of high protein (8% true protein), low fat (<0.5% fat) yoghurt. Yoghurt milk bases were made by adding undenatured whey proteins from native whey protein concentrate (NWPC) to casein concentrate in different whey protein:casein-ratios. The degree of whey protein denaturation was then controlled by the temperature treatment of the yoghurt milk bases. Addition of NWPC in low (whey protein:casein-ratio 25:75) or medium levels (whey protein:casein-ratio 35:65) in combination with heat treatment at 75 °C for 5 min gave yoghurts with significantly lower firmness, lower storage modulus (G′), and better sensory properties (less coarse and granular and more smooth), compared with corresponding yoghurts produced from yoghurt milk bases heat-treated at 95 °C for 5 min or with control yoghurts (no addition of NWPC).  相似文献   

8.
Cryoconcentration combined with a cascade effect was used to concentrate skim milk up to 25.12% total dry matter. Size, shape, and inter-micellar distance of casein micelles were characterized by ZetasizerNano-ZS, transmission electron microscopy, and ImageJ analyses. Flow properties of the cryoconcentrated skim milk were evaluated during 5 weeks of storage under refrigerated condition at 4 °C. Milk color was also evaluated according to the L*, a*, and b* system. The cryoconcentrated skim milk obtained after three cryoconcentration cycles was characterized by a monomodal distribution of its micelles with a tendency to smaller casein micelles. Approximately 60% of the total micellar volume was occupied by the casein micelles with a size of 100–200 nm, less than 18% of the volume with a size of 50–100 nm and only less than 1% was occupied by micelles with a size > 350 nm. This result shows that cryoconcentration changed the distribution of the mean size of the casein micelles to smaller units. No significant difference was observed on the inter-micellar distance. Cryoconcentration significantly improved the color of skim milk by increasing the L* value up to 67 which was similar to that of whole milk. Transition from a Newtonian to a non-Newtonian behavior was observed from the fourth week storage with a slight increase of casein micelle size.Industrial relevanceA concentration procedure of skim milk based on a complete block cryoconcentration technique was proposed. Application of this sub-zero technology permitted the concentration of skim milk total dry matter up to 25%. The casein micelle size was positively affected by moving the major part of the micelles toward the smaller size, whereas the inter-micellar distance was not affected. This new knowledge can be exploited in milk-based products to enhance the product stability. The cryoconcentrated skim milk color was positively affected since its L* value, which represents the milk whiteness, was significantly improved. The flow behavior of the cryoconcentrated milk was of Newtonian type up to 4 weeks of storage at 4 °C. The generated knowledge in this study can be easily used by the milk processing industry in order to make stable milk product with high dry matter content without adding milk powder, which negatively affects the product sensory properties (floury consistency).  相似文献   

9.
The microstructure of acid skim milk gels (14% w/w milk protein low heat powder) with or without addition of locust bean gum (LBG), xanthan gum (XG) and LBG/XG blends was determined by transmission electron microscopy (TEM), phase-contrast light microscopy (PCLM) and scanning electron microscopy (SEM). Three polysaccharide concentrations (0.001%, 0.02% and 0.1%, w/w) were used for binary mixtures. In the case of ternary mixtures, three LBG/XG weight ratios were used (4/16, 11/9 and 16/4) at 0.02% total polysaccharide concentration. Control acid skim milk gels were structured by a homogeneous network of casein particles (0.1–0.7 μm in diameter) and clusters immobilizing whey in small pores (1–5 μm in diameter). Filamentous structures and small aggregates were observed at the surface of casein particles. Low concentration of LBG or XG (0.001% w/w) did not affect markedly the microstructure of acid skim milk gels. Conversely, LBG or XG at 0.02 or 0.1% concentration and LBG/XG blends at the three ratios selected had a great influence on the gel microstructure. Although the size and surface structure of the casein particles were not modified by the presence of polysaccharides, the primary casein network appeared very compact with a decrease of pore size and a large increase in the porosity of the network at the supramolecular level (sponge-like morphology). The effect is stronger for gels containing LBG and XG used at higher concentration and less apparent for gels containing LBG/XG blends. Skim milk/XG gels were highly organized into fibrous structures whereas skim milk/LBG gels were more heterogeneous. These structures were discussed in the light of volume-exclusion effects (demixing) and specific interactions between casein micelles and polysaccharides. At the three weight ratios, skim milk/LBG/XG gels displayed both jagged “coral-like”, “veil-like” and filamentous structures. These structures could originate from a secondary network constituted by the known LBG/XG synergistic interactions.  相似文献   

10.
Native casein micelles were isolated from raw skim milk by ultrafiltration (< 30 kDa) or microfiltration (< 0.2 μm) and subjected to high-pressure homogenization (HPH) at 100, 200, 250, 300, and 350 MPa. Of particular interest was the effect of HPH on casein micelle size in solutions varying in ionic strength (0, 5, 10, and 15 mM CaCl2) and micelle size populations. Particle size distribution reflected an initial decrease in micelle diameter in all samples at 100 MPa. In samples containing 10 and 15 mM CaCl2, there was an abrupt increase in particle size and subsequent casein precipitation followed by sedimentation upon centrifugation at elevated pressures (300 and 350 MPa). The amount of sedimentable casein protein increased as CaCl2 concentration (10 and 15 mM) and pressure (300 and 350 MPa) increased as determined by UV absorbance of sample supernatant. SDS-PAGE indicated extensive micellar disruption at elevated pressures (300 and 350 MPa) and confirmed that the sedimented portion of the samples contained casein proteins and minimal amounts of whey proteins. Results indicated that through HPH treatment casein micelle size can be modified based on CaCl2 concentration and pressure applied. Based on these findings, HPH in combination with an appropriate suspending medium has the ability to modify micelles to a desired size for a number of potential applications.Industrial relevanceThe modification of structure-function properties of the casein micelle from bovine milk by using high-pressure homogenization is relevant in (1) the development of new ingredients to change rheological/textural properties of dairy based foods, and (2) the discovery of new and/or improved functionalities for protein quaternary structures.  相似文献   

11.
The effects of whey protein concentrate (WPC) on the formation of soluble protein complexes and yoghurt texture were evaluated. Skim milk (SM) and skim milk enriched with 1% WPC (SM + 1%WPC) or 2% WPC (SM + 2%WPC) were left unheated or heated and then made into yoghurt gels. Yoghurt prepared from heated SM + 2%WPC had significantly higher storage modulus, water holding capacity and firmness values and a denser microstructure than those prepared only from skim milk. Electrophoretic analysis of the milk showed that the level of β-lactoglobulin and κ-casein in the serum phase increased with increasing WPC concentration, indicating that the content of disulfide-linked β-lactoglobulin and κ-casein was higher in SM + 2%WPC than in SM, suggesting that more soluble protein complexes had been formed. Consequently, yoghurt prepared from heated SM enriched with WPC may have more bonds and more protein complexes in the protein network than yoghurt prepared only from SM, thus resulting in firmer gels.Practical applicationsYoghurt, one of the most popular fermented milk products, is of high economic importance to the dairy industry worldwide. In particular, high-protein yoghurt, such as Greek-style or set-type yoghurt, has been driving its ongoing popularity over recent years. In current industrial production of high-protein yoghurt, protein fortification and heat treatment of milk are two of the most important processing parameters affecting yoghurt texture. Whey protein concentrate has been added to milk to reduce whey separation and to increase the firmness of the yoghurt. From a technological point of view, the interaction of the denatured whey proteins with casein micelles or with κ-casein in the serum phases is regarded as responsible for obtaining a good yoghurt structure. The present research has shown that it is possible to produce yoghurt with a range of textural properties by precisely controlling the rate of whey protein fortification during its manufacture. Therefore, this study provides a better understanding of the effect of WPC fortification and aims to extend this insight for the production of good-quality yoghurt.  相似文献   

12.
《Journal of dairy science》2022,105(5):3871-3882
The interactions among the proteins in sheep skim milk (SSM) during heat treatments (67.5–90°C for 0.5–30 min) were characterized by the kinetics of the denaturation of the whey proteins and of the association of the denatured whey proteins with casein micelles, and changes in the size and structure of casein micelles. The relationship between the size of the casein micelles and the association of whey proteins with the casein micelles is discussed. The level of denaturation and association with the casein micelles for β-lactoglobulin (β-LG) and α-lactalbumin (α-LA) increased with increasing heating temperature and time; the rates of denaturation and association with the casein micelles were markedly higher for β-LG than for α-LA in the temperature range 80 to 90°C; the Arrhenius critical temperature was 80°C for the denaturation of both β-LG and α-LA. The casein micelle size increased by 7 to 120 nm, depending on the heating temperature and the holding time. For instance, the micelle size (about 293 nm) of SSM heated at 90°C for 30 min increased by about 70% compared with that (about 174.6 nm) of unheated SSM. The casein micelle size increased slowly by a maximum of about 65 nm until the level of association of the denatured whey proteins with casein micelles reached 95%, and then increased markedly by a maximum of about 120 nm when the association level was greater than about 95%. The marked increases in casein micelle size in heated SSM were due to aggregation of the casein micelles. Aggregation of the casein micelles and association of whey protein with the micelles occurred simultaneously in SSM during heating.  相似文献   

13.
The role of interfacial properties and size of casein micelles aggregates on foam stability of casein micelle dispersions (CMDs) was examined. CMDs were prepared by redispersing casein micelles pellets obtained by ultracentrifugation. The size of colloidal particles could be controlled by differences in redispersing temperature. CMDs redispersed at 20 °C (CMD20 °C) and 4 °C (CMD4 °C) had average particle sizes of around 200 nm (micelles) and 500 nm (micelles and aggregates), respectively. At 3% total protein, the foam half-life, t½, of CMD4 °C was significantly higher than that of CMD20 °C and skim milk. No correlation between foam stability and surface rheological properties or protein composition could be observed. Foam stability was strongly related to the size of colloidal particles present in CMD. This was confirmed by the observation that the foam stability of CMD4 °C decreased to that of CMD20 °C when the aggregates were broken down by homogenisation.  相似文献   

14.
Seasonal and farming practises can influence milk composition and functionality. An understanding of changes in milk fat globule (MFG) and casein micelle (CM) size may help to guide the selection of milk on the basis of MFG or CM size for manufacturing of different products and product quality. Milk was obtained from cows known to produce predominantly large or small MFG and CM. The rennet gelation properties of this milk were investigated by measuring the rheological properties during gel development. The structure of the CM and MFG network within the rennet gel were characterised by a series of microscopy techniques. Milk with small CM produced firmer curds, and the combination of large MFG (4.49–5.38 μm) with small CM (164–168 nm) produced the firmest curd of any of the combinations tested. MFG size can influence rennet gel firmness, an effect that is dependent upon the pore structure of the CM network.  相似文献   

15.
《International Dairy Journal》2005,15(10):1017-1025
This study examined the effect of hydrolysis of casein by added plasmin (6 mg L−1) on the heat stability of raw, pre-heated, serum protein-free or concentrated skim milk. Plasmin activity markedly affected the heat stability–pH profile of skim milk and serum protein-free milk, apparently by altering the properties of the casein micelles. It is probable that changes in the surface charge of the micelles, as a result of the hydrolysis of caseins, contributed to this effect. Hydrolysis by plasmin reduced the zeta-potential of the casein micelles from ∼−19 to ∼−16 mV. The effect of hydrolysis of casein by plasmin on the heat stability of pre-heated milk was less pronounced, shifting the heat stability–pH profile to more alkaline values; the heat stability of concentrated milk was unaffected by plasmin. A very high (50 mg L−1) level of added plasmin resulted in clearing of the skim milk; the L* value decreased from ∼75 to ∼35 after 24 h incubation at 37 °C. Clearing was correlated with a change in casein micelle diameter from an initial value of ∼175 to ∼250 nm. It is suggested that plasmin-induced changes in zeta-potential may promote micellar aggregation or changes in micelle stucture.  相似文献   

16.
When skim milk at pH 6.55 was heated (75 to 100 degrees C for up to 60 min), the casein micelle size, as monitored by photon correlation spectroscopy, was found to increase during the initial stages of heating and tended to plateau on prolonged heating. At any particular temperature, the casein micelle size increased with longer holding times, and, at any particular holding time, the casein micelle size increased with increasing temperature. The maximum increase in casein micelle size was about 30-35 nm. The changes in casein micelle size were poorly correlated with the level of whey protein denaturation. However, the changes in casein micelle size were highly correlated with the levels of denatured whey proteins that were associated with the casein micelles. The rate of association of the denatured whey proteins with the casein micelles was considerably slower than the rate of denaturation of the whey proteins. Removal of the whey proteins from the skim milk resulted in only small changes in casein micelle size during heating. Re-addition of beta-lactoglobulin to the whey-protein-depleted milk caused the casein micelle size to increase markedly on heat treatment. The changes in casein micelle size induced by the heat treatment of skim milk may be a consequence of the whey proteins associating with the casein micelles. However, these associated whey proteins would need to occlude a large amount of serum to account for the particle size changes. Separate experiments showed that the viscosity changes of heated milk and the estimated volume fraction changes were consistent with the particle size changes observed. Further studies are needed to determine whether the changes in size are due to the specific association of whey proteins with the micelles or whether a low level of aggregation of the casein micelles accompanies this association behaviour. Preliminary studies indicated lower levels of denatured whey proteins associated with the casein micelles and smaller changes in casein micelle size occurred as the pH of the milk was increased from pH 6.5 to pH 6.7.  相似文献   

17.
Micellar Transition State in Casein Between pH 5.5 and 5.0   总被引:1,自引:0,他引:1  
pH-induced changes in casein micelles during direct acidification and bacterial fermentation of reconstituted skim milk at 20°C were monitored by scanning electron microscopy (SEM) in combination with biochemical and rheological methods. For SEM casein micelle observations, an original method of milk sample preparation with porous inorganic membranes was developed. Micrographs suggested that different stages of micellar association were related to pH and that between pH 5.5 and 5.0 casein micelles coalesced. Correlations between microstructural and biochemical changes in casein micelles, and rheological behavior of milk or gel, help to explain the different steps leading to the final protein network of the acid milk gel.  相似文献   

18.
通过纳米粒度分析仪和扫描电子显微镜,分别对水牛乳、牛乳及羊乳中的酪蛋白颗粒直径大小分布情况及酪蛋白胶束结构进行研究。结果表明:水牛乳、牛乳及羊乳中酪蛋白的粒径分布及胶束结构方面存在明显的差异。水牛乳酪蛋白平均颗粒直径为182.3nm,酪蛋白颗粒互相连接成较细长的胶束,胶束之间交联成网络状;牛乳酪蛋白平均颗粒直径为207.4nm,酪蛋白颗粒聚集成直径较大的胶束;羊乳酪蛋白平均颗粒直径为173.8nm,酪蛋白颗粒仅能够形成较短的胶束,也不能交联成网络状。  相似文献   

19.
The enzymatic cross-linking of casein micelles with transglutaminase had an adverse influence on rennet-induced coagulation. Incubation with transglutaminase at 30 °C progressively reduced the levels of monomeric caseins and increased rennet flocculation time (RFT) in a Berridge test. For incubation up to 3 h at 30 °C, the reciprocal of the RFT was linearly correlated with the level of residual monomeric κ-casein, indicating that at complete cross-linking flocculation is absent. After treatment for 4–24 h at 30 °C, no residual monomeric κ-casein was detected and no rennet-induced flocculation of the casein micelle suspension was observed. Monitoring rennet-induced coagulation by diffusing wave spectroscopy revealed that transglutaminase-induced inhibition of rennet-induced coagulation of casein micelles is primarily due to an inhibition of the secondary phase of rennet coagulation, i.e., the gelation and gel-firming phase of the casein micelle coagulation. The gelation and fusion of κ-casein-depleted para-casein micelles as in normal milk appears to be absent if the casein macropeptide remains attached to the casein micelle.  相似文献   

20.
Serum protein concentrates (SPCs) were generated from reconstituted skim milk (3.2% protein) using lab-scale tangential-flow filtration at 3–4 °C. The influence of membrane type on process performance (e.g., permeate flux) and protein-enrichment (e.g., protein profile) was assessed with polyvinylidene-difluoride membranes (0.1 μm and 0.45 μm pore-size), and a polyethersulfone membrane (1000 kDa cut-off). The 1000 kDa membrane exhibited the highest starting flux (6.7 L m−2 h−1), followed by the 0.1 μm (5.4 L m−2 h−1) and 0.45 μm (4.8 L m−2 h−1) membranes. Flux decreased by >40% during filtration with the 1000 kDa and 0.1 μm membranes, while the decrease was lower (<20%) with the 0.45 μm membrane. β-Casein comprised >97% of casein in SPCs from the 0.1 μm and 1000 kDa membranes. SPCs from the 0.45 μm membrane had higher β-casein:αs-casein ratios than the feed and higher levels of minor whey proteins (e.g., lactoferrin) relative to the other SPCs.  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