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 78 毫秒
1.
In order to clarify the interaction of calcium ion with casein, the volume change associated with the interaction was measured by dilatometric procedures. When CaCl2 was added to the casein solutions at neutral pH, a volume increase occurred and reached a constant saturated value of about 700 ml per 106 g protein with increasing CaCl2 concentrations for whole-, αs- and β-casein solutions, but there was no volume change for κ-casein solution. On the other hand, the binding of calcium ion to the casein fractions was determined by a gel filtration procedure at pH 6.0 to 9.0. The number of Ca2+ ions bound to the caseins increased with the CaCl2 concentration and pH value, and the relative order of binding capacities for the caseins was: αs-casein > whole-casein > β-casein > κ-casein.

It was found that the volume changes obtained by the dilatometry were smaller than the calculated volume increases based on the assumption that these are caused by the binding of Ca2+ ion to the caseins. Therefore it is necessary to introduce another factor which reduces the volume increase due to the Ca2+ ion binding in order to reasonably explain the measured volume changes. At present it is presumed that there occurs the unfolding of peptide chain of casein molecule on Ca2+ ion binding, which has been known to decrease the volume of the protein solution.  相似文献   

2.
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   

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.
UDP-N-acetyl-d-galactosamine: κ-casein polypeptide N-acetylgalactosaminyltransferase was purified from a crude Golgi apparatus of lactating bovine mammary gland after solubilization with Triton X-100. Through chromatography on DEAE-Sephadex A-50, apomucin-Sepharose 4B, FPLC mono S, and Sephacryl S-200, and then electrofocusing, the enzyme was purified up to 7500-fold from the homogenate.

The molecular weight of the enzyme was estimated at 200,000 from gel filtration. The pI value of the enzyme was 6.4 on electrofocusing. The purified enzyme transferred GalNAc from UDP-GalNAc, not to the carbohydrate chains but to the polypeptide chains of the substrates, κ-casein and mucin. The enzyme required Mn2+, DTT, and Triton X-100 for maximal activity. The Km value for UDP-GalNAc was 16.2μm. Km values for K-subcomponents 1 and 7, and apomucin were 1.15, 5.10, and 0.192mg/ml, and Vmax values were 254, 259, and 581 nmol/hr/mg, respectively. Thermal stability and the effects of pH, milk components, lectins, and nucleotides were examined.

αs1-Casein strongly inhibited GalNAc transfer to κ-casein. The inhibitory effect of αs1-casein was canceled by the addition of Ca2+, which causes casein micelle formation. This means that the glycosylation of κ-casein occurs after casein micelle formation triggered by the accumulation of Ca2+ in vivo.  相似文献   

5.
The heterogeneity and chemical composition were investigated in κ-casein from colostrum. The acid casein was obtained from four different Holstein cow colostra. The yield of acid casein from colostrum was higher than that from normal milk. κ-Casein from colostrum was prepared by the gel filtration method of Yaguchi et al. The gel filtration profiles differed among the four colostrum acid caseins.

Colostrum κ-casein was fractionated on a DEAE-cellulose column into one nonadsorbed and six adsorbed fractions with increasing salt concentration. Six adsorbed fractions had the same molecular weight and stabilizing ability for αs1-casein in the presence of calcium ion. The amino acid composition and the phosphorus content of the adsorbed fractions were identical, but fractions eluted with high salt concentrations had more carbohydrates (galactose, sialic acid, glucosamine, galactosamine). Colostrum κ-casein was characterized by a higher content of carbohydrate moiety in comparison with normal κ-casein. Also glucosamine which has not been found in normal κ-casein was detected in colostrum κ-casein. The κ-casein component from colostrum contained at least one molecule of carbohydrate, though the carbo hydrate-free component was detected in normal κ-casein.  相似文献   

6.
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   

7.
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   

8.
9.
The amount of casein found in the milk of the tammar wallaby increases as lactation progresses. The increase is due to increasing amounts of β-casein; the α-casein remains largely constant. The α-casein is the more highly phosphorylated; the most abundant form is the 10-P, throughout lactation. The level of phosphorylation of β-casein shifts to lower average values in late lactation, possibly indicating the enzymatic reaction is overloaded by the increasing amounts of β-casein. Unlike bovine casein micelles, the wallaby micelles are not completely disrupted at pH 7.0 by sequestration of their calcium content with ethylene diamine tetraacetic acid (EDTA). Complete disruption only follows the addition of sodium dodecyl sulphate, indicating considerably greater importance for hydrophobic bonds in maintaining their integrity. This micellar behaviour indicates that, despite the evolutionary divergence of marsupials millennia ago, the caseins of wallaby milk assemble into micelles in much the same fashion as in bovine milk.  相似文献   

10.
The structure of casein micelles has been studied by small-angle neutron scattering and static light scattering. Alterations in structure upon variation of pH and scattering contrast, as well as after addition of chymosin, were investigated. The experimental data were analyzed by a model in which the casein micelle consists of spherical submicelles. This model gave good agreement with the data and gave an average micellar radius of about 100–120 nm and a submicellar radius of about 7 nm both with a polydispersity of about 40–50%. The contrast variation indicated that the scattering length density of the submicelles was largest at the center of the submicelles. The submicelles were found to be closely packed, the volume fraction varying slightly with pH. Upon addition of chymosin the submicellar structure remained unchanged within the experimental accuracy. Correspondence to: S. Hansen  相似文献   

11.
Proteolytic activity in the extract from the cells of Streptococcus cremoris increased in the presence of casein, lactose, glucose, and CaCl2 in the media but was negligibly detectable in the extract of the cells harvested from the culture containing succinate or citrate. The intracellular proteinase from S. cremoris harvested from tomato medium was purified 150-fold in this experiment. The enzyme had a molecular weight of 140,000, optimum pH at 6.5 to 7.0, and maximum activity at 30 C. The proteinase was activated by Ca2+ and inhibited by Zn2+, Cu2+, Hg2+, Fe2+, ethylenediaminetetraacetate, and sodium lauryl sulfate. The Km value of the enzyme towards each casein fraction was almost the same, and the Vmax of the enzyme towards αs-casein was smaller than those towards the other casein fractions.  相似文献   

12.
Bovine κ-casein, a phosphoglycoprotein, has mucin-type carbohydrate chains. Subcellular distribution of enzymes that take part in the post-translational modification of κ-casein was examined. In lactating mammary glands from rats and cows, N-acetyl-galactosaminyl transferase, galactosyl transferase, sialyl transferase, and casein kinase were localized specifically in the Golgi apparatus.

The substrate specificities indicate that these enzymes are actually responsible for the processing of κ-casein.

The presence of a phosphate group attached to κ-casein did not affect the rate of glycosylation by N-acetyl-galactosaminyl transferase, while the presence of carbohydrate chains attached to κ- casein strongly reduced the rate of phosphorylation by casein kinase. These results suggest that in the Golgi apparatus, phosphorylation of κ-casein precedes glycosylation.  相似文献   

13.
Time-dependent light-scattering studies have been made on mixtures of αs1 -casein and Ca2+ at fixed temperature over a range of [Ca2+] and [αs1 -casein], and also as functions of temperature- Measurements were also made of the extent of precipitate formation in the casein/Ca2+ mixtures, using centrifugation. The results are analysed in terms of a monomeroctamer equilibrium between calcium caseinate particles followed by a Smoluchowski aggregation in which only the octamers can participate. The equilibrium constant is dependent upon the charge on the protein/Ca2+ particles, and hence can be related to the extent of binding of Ca2+ to the αs1 -casein. The Smoluchowski constant is likewise shown to be charge-dependent. The variation of the reaction rate with temperature can be ascribed solely to the changing charge of the αs1 -casein/Ca2+ complex caused by changed binding of Ca2+ at different temperatures.  相似文献   

14.
The catalytic subunit of rabbit muscle cyclic AMP-dependent protein kinase (EC 2.7.1.37; ATP:protein transferase) has been tested on a variety of caseins. The B variant of β-casein was phosphorylated at a much greater rate than other β-caseins, αs1-caseins, and κ-caseins. Whole casein homozygous for β-casein B was phosphorylated at 2.5 times the rate of commercial whole casein. Gel electrophoresis experiments indicate that β-casein is the predominant component phosphorylated in commerical casein. It is therefore suggested that phosphorylation of whole casein depends on its content of the specific genetic variant, β-casein B.  相似文献   

15.
1. Fractions have been obtained from human whole casein closely resembling the αs- and κ-fractions of cow casein. 2. The αs-fraction (human αs-casein) is calcium-sensitive, heterogeneous in zone analysis and inert towards rennin. 3. The κ-fraction (human κ-casein) is calcium-insensitive, heterogeneous in zone analysis, and forms a soluble glycopeptide when acted upon by rennin. 4. Human κ-casein stabilizes human αs-casein in the presence of Ca2+ ions. 5. The glycopeptides released by rennin from human casein and from cow casein have been compared. There are important differences in both the peptide and non-peptide structures of the two compounds. 6. In both human and bovine glycopeptides some of the carbohydrate residues are joined to the peptide by O-glycosidic links with threonine, and possibly with serine.  相似文献   

16.
Membrane fractions from mature silver beet (Beta vulgaris) deveined leaf and leaf stem homogenates have associated Ca2+ -dependent protein kinase. The Ca2+ -dependent protein kinase activity is associated with plasma membranes (density 1.14-1.18 grams per cubic centimeter) as determined from copurification on isopycnic centrifugation with plasma membrane markers such as β-glucan synthetase, eosin-5-maleimidelabeling, and specific naphthylphthalamic acid-binding. The Ca2+ -dependent protein kinase is not specifically associated with chloroplasts or mitochondria. The membrane-bound Ca2+ -dependent protein kinases were solubilized with 0.8% (volume/volume) Nonidet P40. The solubilized enzymes were extensively purified by a protocol involving binding to diethylaminoethyl-cellulose (Whatman DE-52), Ca2+ -dependent binding to phenyl-Sepharose CL-4B, gradient elution from diethylaminoethyl-Sephacel (resolving two distinct Ca2+ -dependent protein kinases), and gel filtration on Ultrogel AcA 44. These two membrane-derived enzymes have similar molecular weights but differ in protein substrate specificity, in Km values for ATP, and in Ca2+ -independent activation by unsaturated fatty acids. The membrane-bound enzymes correspond closely in these properties to two Ca2+ -dependent protein kinases present in the soluble phase.  相似文献   

17.
The chymosin-induced flocculation of casein micelles of bovine milk can be explained and calculated in terms of three relationships, which are (i) the action of chymosin upon the κ-casein of the micelles; (ii) the probability that a micelle, with a given proportion of its κ-casein destroyed, will aggregate, and (iii) the aggregation of micelles by a Smoluchowski mechanism. Details of the calculations are given, and the theory is shown to be in good agreement with experimental observations of the dependence of the clotting time with variations in enzyme and substrate concentrations.  相似文献   

18.
S-carboxymethylated (SCM) κ-casein forms in vitro fibrils that display several characteristics of amyloid fibrils, although the protein is unrelated to amyloid diseases. In order to get insight into the processes that prevent the formation of amyloid fibrils made of κ-caseins in milk, we have characterized in detail the reaction and the roles of its possible effectors: glycosylation and other caseins. Given that native κ-casein occurs as a heterogeneous mixture of carbohydrate-free and carbohydrate-containing chains, kinetics of fibril formation were performed on purified glycosylated and unglycosylated SCM κ-caseins using the fluorescent dye thioflavin T in conjunction with transmission electron microscopy and Fourier transform infrared spectroscopy for morphological and structural analyses. Both unglycosylated and glycosylated SCM κ-caseins have the ability to fibrillate. Kinetic data indicate that the fibril formation rate increases with SCM κ-casein concentration but reaches a plateau at high concentrations, for both the unglycosylated and glycosylated forms. Therefore, a conformational rearrangement is the rate-limiting step in fibril growth of SCM κ-casein. Transmission electron microscopy images indicate the presence of 10- to 12-nm spherical particles prior to the appearance of amyloid structure. Fourier transform infrared spectroscopy spectra reveal a conformational change within these micellar aggregates during the fibrillation. Fibrils are helical ribbons with a pitch of about 120-130 nm and a width of 10-12 nm. Taken together, these findings suggest a model of aggregation during which the SCM κ-casein monomer is in rapid equilibrium with a micellar aggregate that subsequently undergoes a conformational rearrangement into a more organized species. These micelles assemble and this leads to the growing of amyloid fibrils. Addition of αs1-and β-caseins decreases the growth rate of fibrils. Their main effect was on the elongation rate, which became close to that of the limiting conformation change, leading to the appearance of a lag phase at the beginning of the kinetics.  相似文献   

19.
Summary The location of the glycosylated part of κ-casein in bovine casein micelles was investigated using gold particles (6 nm in diameter) labelled withRicinus communis lectin andLimulus polyphemus lectin. The pattern of marking of thin sections of micelles was similar with both lectins. Glycosylated κ-casein was distributed uniformly throughout most micelles of all sizes. Peripheral location of glycosylated κ-casein was observed only occasionally in some of the largest micelles. Quantitative data indicated that the concentration of the glycosylated protein was constant in micelles of increasing sizes. As larger micelles contain less total κ-casein than smaller ones, these data indicated that a greater proportion of their κ-casein is glycosylated. These results support models for casein micelle structure where κ-casein is distributed throughout the micelles. They do not agree with “coat-core” structures.  相似文献   

20.
Reversible and irreversible states of pressure-dissociated casein micelles were studied by in situ light scattering techniques and ex situ atomic force microscopy. AFM experiments performed at ambient pressure reveal heterogeneities across the micelle, suggesting a sub-structure on a 20 nm scale. At pressures between 50 and 250 MPa, the native micelles disintegrate into small fragments on the scale of the observed sub-structure. At pressures above 300 MPa the micelles fully decompose into their monomeric constituents. After pressure release two discrete populations of casein aggregates are observed, depending on the applied initial pressure: Between 160 and 240 MPa stable micelles with diameters near 100 nm without detectable sub-structures are formed. Casein micelles exposed to pressures above 280 MPa re-associate at ambient pressure yielding mini-micelles with diameters near 25 nm. The implications concerning structural models are discussed.  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