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 406 毫秒
1.
Casein micelles of bovine skimmed milk were fractionated by permeation chromatography on porous glass (CPG-10, 50 nm followed by CPG-10, 300 nm) at 30 degrees C. Micelles were pooled in eight eluant fractions and their size distribution was determined by electron microscopy. The composition of casein in the eight fractions was determined by quantitative hydroxyapatite chromatography. Micelle size decreased progressively with increasing elution volume, and volume-to-surface average diameter ranged from 154 nm in fraction 1 to 62 nm in fraction 8. Concurrently there was a decrease in relative proportions of alpha s- and beta-caseins and a large enrichment of kappa-casein, which changed from 4.1% total casein in fraction 1 to 12.1% total casein in fraction 8. At least half the decrease in alpha s-casein proportions was attributed to the alpha s1-casein component, but the data also suggested a decline in proportions of alpha s2-casein in the smallest micelle fractions. A plot of kappa-casein fractional content versus micelle surface-to-volume ratio gave a straight line (correlation coefficient from linear regression 0.98) from which an average kappa-casein surface coverage of 1.5 m2/mg or 47.3 nm2/molecule was obtained. If a constant surface coverage for kappa-casein is assumed, the parameters of the linear equation predict that micelle voluminosity is inversely related to micelle diameter, being approximately 30% larger in fraction 8 compared to fraction 1.  相似文献   

2.
Isolation and properties of human kappa-casein   总被引:1,自引:0,他引:1  
Human kappa-casein was isolated from human whole casein by gel filtration with Sephadex G-200 and hydroxylapatite chromatography in the presence of sodium dodecyl sulfate (SDS). The kappa-casein was calcium-insensitive and did stabilize human beta-casein and bovine alpha s1-casein against precipitation by calcium ions. Formation of micelles from human beta- and kappa-caseins, and calcium ions was confirmed by electron microscopic observation. On SDS-polyacrylamide gel electrophoresis (SDS-PAGE), a single band was obtained. The formation of para-kappa-caseins by chymosin was confirmed by SDS-PAGE. Two para-kappa-caseins with apparent molecular weights of 13,000 and 11,000 appeared. The molecular weight of intact human kappa-casein was estimated to be approximately 33,000. The human kappa-casein contained about 40% carbohydrate (15% galactose, 3% fucose, 15% hexosamines, and 5% sialic acid) and 0.10% (1 mol/mol) phosphorus. Its amino acid composition was similar to that of bovine kappa-casein except for serine, glutamic acid, and lysine contents.  相似文献   

3.
Human casein micelles were reconstituted from isolated κ- and β-caseins and calcium ions. Micelle formation was recognized in the presence of calcium chloride even at the low concentration of 5mM. At pH levels ranging from 5.5 to 8.0, the re-formed micelles were quite stable so that precipitation of β-casein was not observed. The large micelles were constituted by a higher ratio of β-casein to κ-casein (16:1 by weight) than the small micelles (3: 1). The κ-casein in the small micelles contained carbohydrates to about 43% (w/w) in the molecule, whereas that in the large micelles was only about 25%. When the casein micelles were re-formed from κ-easein and fractionated β-casein components, the extent of phosphorylation of the β-casein component was found to influence the micelle formation; i.e., the β-casein component with no phosphate (the 0-P form) was disadvantageous to form micelles, but the component with 5 phosphates (the 5-P form) formed micelles most easily.  相似文献   

4.
Bovine casein micelles were fractionated on controlled pore granule (CPG-10/3000) chromatography by size and the chemical properties of the fractionated micelles were compared. The results indicated the presence of two types of micelles distinguishable as large and small micelles. In skim milk, 72.7% of casein was calculated to be in the form of small micelles, 13.6% in the form of large micelles and 13.8% in non-micellar casein form.

The αs1-casein content decreased, but β- and κ-casein content increased as the micelle size became smaller. κ-Casein in large micelles had a much higher sialic acid content than in small micelles. It was found that this difference in sialic acid content was due to the presence of non-glycosylated κ-casein in small micelles. In large micelles, non-glycosylated κ-casein was almost undetectable.

The addition of wheat germ lectin to micelles resulted in the formation of aggregates through intermicellar bridges between the carbohydrate chains of κ-casein located on the surface of the micelles. Both large and small micelles formed aggregates after the addition of wheat germ lectin. Large micelles were more sensitive to wheat germ lectin than small ones.  相似文献   

5.
The casein complexes of bovine milk consist of four major protein fractions, alpha s1, alpha s2, beta, and kappa. Colloidal particles of casein (termed micelles) contain inorganic calcium and phosphate; they are very roughly spherical with an average radius of 650 A. Removal of Ca2+ leads to the formation of smaller protein aggregates (submicelles) with an average radius of 94 A. Two genetic variants, A and B, of the predominant fraction, alpha s1-casein, result in milks with markedly different physical properties, such as solubility and heat stability. To investigate the molecular basis for these differences, small-angle X-ray scattering was performed on the respective colloidal micelles and submicelles. Scattering curves for submicelles of both variants showed multiple Gaussian character; data for the B variant were previously interpreted in terms of two concentric regions of different electron density, i.e., a "compact" core and a relatively "loose" shell. For the submicelle of A, there was a third Gaussian, reflecting a negative contribution due to interparticle interference. Molecular parameters for submicelles of both A and B are in agreement with hydrodynamic data in the literature. Data for the micelles, for which scattering yields cross-sectional information, were fitted by a sum of three Gaussians for both variants; for these, the corresponding two lower radii of gyration represent the two concentric regions of the submicelles, while the third reflects the average packing of submicelles within the micellar cross section. Most of the molecular parameters obtained showed small but consistent differences between A and B, but for submicelles within the micelle several differences were particularly notable: A has a greater molecular weight for the "compact" region of the constituent submicelle (82,000 vs 60,000) and a much greater submicellar packing number (6:1 vs 3:1). Reasons for these and other differences are to be sought in sequence differences and in differences in calcium-binding sites and charge distribution.  相似文献   

6.
Potato acid phosphatase (EC 3.1.3.2) was used to remove the eight phosphate groups from alphas1-casein. Unlike most acid phosphatases, which are active at pH 6.0 or below, potato acid phosphatase can catalyze the dephosphorylation of alphas1-casein at pH 7.0. Although phosphate inhibition is considerable (K1=0.42 mM phosphate), the phosphate ions produced by the dephosphorylation of casein can be removed by dialysis, allowing the reaction to go to completion. The dephosphorylated alphas1-casein is homogeneous on gel electrophoresis with a slower mobility than native alphas1-casein and has an amino acid composition which is identical to native alphas1-casein. Thus the removal of phosphate groups from casein does not alter its primary structure. Potato acid phosphatase also removed the phosphate groups from other phosphoproteins, such as beta-casein, riboflavin binding protein, pepsinogen, ovalbumin, and phosvitin.  相似文献   

7.
The structure of thin casein films prepared with spin-coating is investigated as a function of the calcium concentration. Grazing incidence small-angle x-ray scattering and atomic force microscopy are used to probe the micelle structure. For comparison, the corresponding casein solutions are investigated with dynamic light-scattering experiments. In the thin films with added calcium three types of casein structures, aggregates, micelles, and mini-micelles, are observed in coexistence with atomic force microscopy and grazing incidence small-angle x-ray scattering. With increasing calcium concentration, the size of the aggregates strongly increases, while the size of micelles slightly decreases and the size of the mini-micelles increases. This effect is explained in the framework of the particle-stabilizing properties of the hairy layer of kappa-casein surrounding the casein micelles.  相似文献   

8.
Casein of cynomolgus monkey was compared with those from human and bovine milk. Cynomolgus monkey casein showed similar electrophoretical patterns to those of human casein on Disc- and SDS-electrophoresis. It consisted of beta- and kappa-casein-like components. The component corresponding to bovine alpha s1-casein was not detected. The beta-casein-like fraction of cynomolgus monkey showed 9 bands on Disc-PAGE. These were suggested to be the same protein binding different levels of phosphorus by dephosphorylation experiment using an acid phosphatase. The kappa-casein-like component of cynomolgus monkey was highly glycosylated (about 50% carbohydrate) similarly as human kappa-casein and the constituent carbohydrates were same as those detected in human kappa-casein (galactose, fucose, N-acetylgalactosamine, N-acetylglucosamine, and sialic acid). Amino acid composition of cynomolgus monkey kappa-casein bore a resemblance to those of both human and bovine kappa-caseins. Amino acid composition of cynomolgus monkey beta-casein was also similar to those of human and bovine beta-caseins.  相似文献   

9.
Bovine casein components (αsl-, β-, and κ-caseins) were chemically phosphorylated and the properties of the modified components were compared with those of the native to clarify the function of the intrinsic phosphate groups of casein components in casein micelle formation. The calcium binding ability of casein components increased after chemical phosphorylation. The concentrations of calcium chloride required to precipitate modified αsl- and β-caseins were higher than those for native components. However, phosphorylation of αsl- and β-caseins did not affect their properties of forming micelles through interaction with κ-casein. The stabilizing ability of κ-casein for αsl- and β caseins was impaired by its phosphorylation, but the stability was recovered by treating phosphorylated κ-casein with phosphoprotein phosphatase. The results show that the phosphate content of κ-casein must be low to form a stable casein micelle. The results also explain why the specific phosphorylation of casein components in the mammary gland is required.  相似文献   

10.
Approximately 40% of the calcium existing in colloidal phase of skimmilk was estimated to be hardly exchanged with the calcium psesent in soluble phase by applying a radioisotopic technique. This type of calcium was designated hard-to-exchange calcium. Hard-to-exchange calcium was absent or nearly zero in calcium caseinate dispersion or colloidal phosphate-free milk, but was present in composite calcium caseinate phosphate dispersion. It is suggested that hard-to-exchange calcium is present in a part of colloidal phosphate portion of casein micelles.  相似文献   

11.
Sequestration of calcium phosphate by caseins occurs in the Golgi region of mammary secretory cells during lactation, where it helps to prevent calcification of the gland and to deliver high concentrations of calcium and phosphate to the neonate in the form of milk. Calcium phosphate nanoclusters are formed when a core of amorphous calcium phosphate is sequestered within a shell of casein or casein phosphopeptides. The nanoclusters can form spontaneously from a supersaturated solution or by dispersion of a precipitate of calcium phosphate, demonstrating that they are thermodynamically stable complexes. The average size and chemical composition of the complexes are largely independent of the solution conditions (pH, temperature, peptide concentration, salt composition and rate of reaction) under which they form. Larger, metastable, colloidal particles can form if there is not enough of the phosphopeptide to sequester all the calcium phosphate, or, transiently, if the salt and peptide solutions are mixed together without sufficient care. A thermodynamic model of the sequestration process is presented which makes use of an invariant ion activity product observed in nanocluster-containing solutions. In any given solution that has thermodynamic stability, the extent of the sequestration reaction can be calculated from the empirical formula of the nanoclusters using the criterion that the solution should have the equilibrium value of the invariant ion activity product. Other members of the paralogous group of secretory calcium-binding phosphoproteins to which caseins belong may also be able to sequester calcium phosphate in biological fluids such as saliva and in the extracellular matrix of mineralizing tissues.Abbreviations -PP s1-casein 5P (f59–79) - -PP -casein 4P (f1–25) - ACP amorphous calcium phosphate - Cit citrate - CPN calcium phosphate nanocluster - CPP commercial phosphopeptide - IAP ion activity product - MWCO molecular weight cut-off - PP phosphopeptide - SAXS small-angle X-ray scattering - SCPP secretory calcium-binding phosphoprotein - UF ultrafiltrate  相似文献   

12.
The caseins occur in milk as colloidal complexes of protein aggregates, calcium, and inorganic phosphate. As determined by electron microscopy, these particles are spherical and have approximately a 650 Å radius (casein micelles). In the absence of calcium, the protein aggregates themselves (submicelles) have been shown to result from mainly hydrophobic interactions. The fractional concentration of stable colloidal casein micelles can be obtained in a calcium caseinate solution by centrifugation at 1500g. Thus, the amount of stable colloid present with varying Ca2+ concentrations can be determined and then analyzed by application of equations derived from Wyman's Thermodynamic Linkage Theory. Ca2+-induced colloid stability profiles were obtained experimentally for model micelles consisting of only s1- (a calcium insoluble casein) and the stabilizing protein -casein, eliminating the complications arising from - and minor casein forms. Two distinct genetic variants s1-A andB were used. Analysis of s1-A colloid stability profiles yielded a precipitation (salting-out) constantk 1, as well as colloid stability (salting-in) parameterk 2. No variations ofk 1 ork 2 were found with increasing amounts of -casein. From the variation of the amount of colloidal casein capable of being stabilized vs. amount of added -casein an association constant of 4 L/g could be calculated for the complexation of s1-A and -casein. For the s1-B and -casein micelles, an additional Ca2+-dependent colloidal destabilization parameter,k 3, was added to the existingk 1 andk 2 parameters in order to fully describe this more complex system. Furthermore, the value ofk 3 decreased with increasing concentration of -casein. These results were analyzed with respect to the specific deletion which occurs in s1-caseinA in order to determine the sites responsible for these Ca2+-induced quaternary structural effects.  相似文献   

13.
The stability of internally cross-linked casein micelles against disruption by urea (which disrupts hydrogen bonds and hydrophobic interactions) and trisodium citrate (which sequesters micellar calcium phosphate) was investigated. Addition of urea (0-6 mol L-1) and/or citrate (0-50 mmol L-1) progressively reduced the turbidity of a suspension of casein micelles cross-linked by transglutaminase and increased particle size (determined by dynamic and static light scattering and small-angle neutron scattering), which was attributed to swelling of the micelles. Furthermore, model calculations, assuming a completely stable casein network, were performed to describe the decreases in turbidity on addition of urea and citrate. Measured and described turbidity values are in agreement, indicating that cross-linking of casein micelles with transglutaminase results in a covalently bound protein network, which is entirely stable to disruption by urea and/or citrate. This may offer potential applications for the use of cross-linked casein micelles as biocompatible protein micro-gel particles.  相似文献   

14.
The present studies show that the colloidal calcium phosphate of cow's milk has a (Ca + Mg)/Pi ratio of 1.67 (± 0.10; n = 22) and contains citrate, Mg and Zn at molar ratios to Ca averaging 0.05, 0.03 and 0.003, respectively. The composition of the natural colloidal phosphate of milk is similar to the precipitates formed by neutralization of ultrafiltrates obtained from acidified milks, and to that of the calcium phosphate-enriched fraction produced by extensive enzymic hydrolysis of the casein micelles in milk. Examination by electron microscopy of these artificial preparations of milk calcium phosphate revealed in both a very fine and uniform substructure which consisted of granules having an average, true diameter of approx. 2.5 nm. The size and shape of these tiny granules closely resemble the morphologies reported for the colloidal phosphate particles in native casein micelles, as well as for the subunits of amorphous calcium phosphate observed during calcification in other biological systems such as mitochondria and bone.  相似文献   

15.
Rat liver mitochondrial fractions corresponding to four morphological structures (matrix, inner membrane, intermembrane space and outer membrane) contain proteinases that cleave casein components at different rates. Proteinases of the intermembrane space preferentially cleave kappa-casein, whereas the proteinases of the outer membrane, inner membrane and matrix fractions degrade alpha S1-casein more rapidly. Electrophoretic separation of the degradation products of alpha S1-casein and kappa-casein in polyacrylamide gels shows that different polypeptides are produced when the substrate is degraded by the matrix, by both membranes and by the intermembrane-space fraction. Some of the degradation products resulting from incubation of the caseins with the mitochondrial fractions are probably the result of digestion by contaminating lysosomal proteinase(s). The matrix has a high peptidase activity, since glucagon, a small peptide, is very rapidly degraded by this fraction. These observations strongly suggest that distinct proteinases, with different specificities, are associated respectively with the intermembrane space and with both membrane fractions.  相似文献   

16.
The influence of enzymatic crosslinking by microbial transglutaminase (mTG) on the stability of casein micelles of ultrahigh temperature (UHT)-treated milk in the presence of EDTA (0-0.45 mM) or ethanol (0-74 vol%) as well as under high hydrostatic pressures up to 400 MPa was investigated. Disintegration of micelles and changes in micelle size were monitored by the measurement of turbidity as well as by dynamic light scattering. The results show that the incubation of UHTtreated milk with mTG resulted in an improved micelle stability toward disintegration on addition of EDTA, ethanol, or pressure treatment. Intramicellar formed isopetides significantly enhanced the stability of casein micelles. It is supposed that net-like crosslinks are formed within the external region of the micelles and they adopt the stabilizing role of colloidal calcium phosphate within the micelles, thus making the micelles less contestable for disrupting influences.  相似文献   

17.
The possible mechanism of casein aggregation and micelle buildup was studied in a new approach by letting α-casein adsorb from low concentration (0.1 mg·ml(-1)) solutions onto the charged surfaces of polyelectrolyte films. It was found that α-casein could adsorb onto both positively and negatively charged surfaces. However, only when its negative phosphoseryl clusters remained free, i.e. when it adsorbed onto a negative surface, could calcium phosphate (CaP) nanoclusters bind to the casein molecules. Once the CaP clusters were in place, step-by-step building of multilayered casein architectures became possible. The presence of CaP was essential; neither Ca(2+) nor phosphate could alone facilitate casein aggregation. Thus, it seems that CaP is the organizing motive in the casein micelle formation. Atomic force microscopy revealed that even a single adsorbed casein layer was composed of very small (in the range of tens of nanometers) spherical forms. The stiffness of the adsorbed casein layer largely increased in the presence of CaP. On this basis, we can imagine that casein micelles emerge according to the following scheme. The amphipathic casein monomers aggregate into oligomers via hydrophobic interactions even in the absence of CaP. Full scale, CaP-carrying micelles could materialize by interlocking these casein oligomers with CaP nanoclusters. Such a mechanism would not contradict former experimental results and could offer a synthesis between the submicelle and the block copolymer models of casein micelles.  相似文献   

18.
A method was developed to distinguish between genotypic variants B and C of bovine alpha s1-casein, using the allele-specific polymerase chain reaction (ASPCR). The alpha s1-casein genotype determined for 17 Jersey cows by the ASPCR method was confirmed by typing the alpha s1-casein milk proteins on isoelectric focusing gels. Using the ASPCR method described, rapid analysis of the alpha s1-casein genotype of bulls is now possible. In addition, kappa-casein genotypes can be determined from the same PCR reaction.  相似文献   

19.
Chromatography of glutaraldehyde-fixed skim-milk on controlled-pore glass (CPG-10, 300 nm) gave three micellar fractions whose averaged diameters, measured by electron microscopy, decreased progressively with increasing elution volume. Casein micelles with diameters up to 680 nm were detected. The casein composition of the same fractions from unfixed skim-milk was determined. As the fraction elution volume increased, κ-casein varied from 7.7 to 11.4% of total casein, giving αs/κ ratios of 6.1, 4.7 and 3.3.A plot of κ-casein content versus micelle surface-to-volume ratio for skim-milk and the column fractions approximated to a straight line. Re-calculation of the published results from two other studies also gave linear relationships between κ-casein content and surface area for artificial micelles. The three regression lines thus obtained had small intercepts. It was concluded that the data indicated the same fundamental structure for casein micelles, with a pre-dominant surface location for κ-casein, whether the micelles are natural or artificial and whether they are aggregated or by Ca2+ alone oy Ca2+ together with calcium phosphate-citrate complex.  相似文献   

20.
In our previous paper (Nagy et?al. in J Biol Chem 285:38811–38817, 2010) by using a multilayered model system, we showed that, from α-casein, aggregates (similar to natural casein micelles) can be built up step by step if Ca-phosphate nanocluster incorporation is ensured between the protein adsorption steps. It remained, however, an open question whether the growth of the aggregates can be terminated, similarly to in nature with casein micelles. Here, we show that, in the presence of Ca-phosphate nanoclusters, upon adsorbing onto earlier α-casein surfaces, the secondary structure of α-casein remains practically unaffected, but κ-casein exhibits considerable changes in its secondary structure as manifested by a shift toward having more β-structures. In the absence of Ca-phosphate, only κ-casein can still adsorb onto the underlying casein surface; this κ-casein also expresses considerable shift toward β-structures. In addition, this κ-casein cover terminates casein aggregation; no further adsorption of either α- or κ-casein can be achieved. These results, while obtained on a model system, may show that the Ca-insensitive κ-casein can, indeed, be the outer layer of the casein micelles, not only because of its “hairy” extrusion into the water phase, but because of its “softer” secondary structure, which can “occlude” the interacting motifs serving casein aggregation. We think that the revealed nature of the molecular interactions, and the growth mechanism found here, might be useful to understand the aggregation process of casein micelles also in?vivo.  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