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 62 毫秒
1.
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   

2.
This study aimed to evaluate amino acids content and the electrophoretic profile of camel milk casein from different camel breeds. Milk from three different camel breeds (Majaheim, Wadah and Safrah) as well as cow milk were used in this study.Results showed that ash and moisture contents were significantly higher in camel milk casein of all breeds compared to that of cow milk. On the other hand, casein protein of cow milk was significantly higher compared to that of all camel milk breeds. Molecular weights of casein patterns of camel milk breeds were higher compared to that of cow milk.Essential (Phe, Lys and His) and non-essential amino acids content was significantly higher in cow milk casein compared to the casein of all camel milk breeds. However, there was no significant difference for the other essential amino acids between cow casein and the casein of Safrah breed and their quantities in cow and Safrah casein were significantly higher compared to the other two breeds. Non-essential amino acids except Arg and the essential amino acids (Met, Ile, Lue and Phe) were also significantly higher in cow milk α-casein compared to α-casein from all camel breeds. Moreover, essential amino acids (Val, Phe and His) and the non-essential amino acids (Gly and Ser) content was significantly higher in cow milk β-casein compared to the β-casein of all camel milk breeds and the opposite was true for Lys, Thr, Met and Ile. However, Met, Ile, Phe and His were significantly higher for β-casein of Majaheim compared to the other two milk breeds. The non-essential amino acids (Gly, Tyr, Ala and Asp) and the essential amino acids (Thr, Val and Ile) were significantly higher in cow milk κ-casein compared to that for all camel milk breeds. There was no significant difference among all camel milk breeds in their κ-casein content of most essential amino acids.Relative migration of casein bands of camel milk casein was not identical. The relative migration of αs-, β- and κ-casein of camel casein was slower than those of cow casein. The molecular weights of αs-, β- and κ-casein of camel caseins were 27.6, 23.8 and 22.4 KDa, respectively. More studies are needed to elucidate the structure of camel milk.  相似文献   

3.
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   

4.
The stabilizing action of carboxymethyl cellulose (CMC-1 and CMC-2) on caseins was studied in the acidic pH region. CMC-1 stabilized 1% whole, α-, αS- and β-casein at pH 4.6 and 5.0, and at 5°C. But CMC-2 could not completely stabilize these caseins at pH 5.0. Interaction between κ-casein and CMC-1 commenced when pH was adjusted to 6.3, but CMC-2 interacted with κ-casein below pH 5.6. An αS- and κ-casein mixture (4 : 1) with CMC-2 was destabilized by the addition of 0.02 m NaCl or NaH2PO4 at pH 5.0. The αS/κ ratio of the precipitated casein was about 10. But the same system with CMC-1 was not destabilized by the salts.  相似文献   

5.
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   

6.
Elevated homocysteine levels are resulting in N-homocysteinylation of lysyl residues in proteins and they correlate with a number of human pathologies. However, the role of homocysteinylation of lysyl residues is still poorly known. In order to study the features of homocysteinylation of intrinsically unstructured proteins (IUP) bovine caseins were used as a model. α(S1)-, β- and κ-caseins, showing different aggregations and micelle formation, were modified with homocysteine-thiolactone and their physico-chemical properties were studied. Efficiency of homocysteine incorporation was estimated to be about 1.5, 2.1 and 1.3 homocysteyl residues per one β-, α(S1)-, and κ-casein molecule, respectively. Use of intrinsic and extrinsic fluorescent markers such as Trp, thioflavin T and ANS, reveal structural changes of casein structures after homocysteinylation reflected by an increase in beta-sheet content, which in some cases may be characteristic of amyloid-like transformations. CD spectra also show an increase in beta-sheet content of homocysteinylated caseins. Casein homocysteinylation leads in all cases to aggregation. The sizes of aggregates and aggregation rates were dependent on homocysteine thiolactone concentration and temperature. DLS and microscopic studies have revealed the formation of large aggregates of about 1-3μm. Homocysteinylation of α(S1)- and β-caseins results in formation of regular spheres. Homocysteinylated κ-casein forms thin unbranched fibrils about 400-800nm long. In case of κ-casein amyloidogenic effect of homocysteinylation was confirmed by Congo red spectra. Taken together, data indicate that N-homocysteinylation provokes significant changes in properties of native caseins. A comparison of amyloidogenic transformation of 3 different casein types, belonging to the IUP protein family, shows that the efficiency of amyloidogenic transformation upon homocysteinylation depends on micellization capacity, additional disulphide bonds and other structural features.  相似文献   

7.
αS-Casein, the major milk protein, comprises αS1- and αS2-casein and acts as a molecular chaperone, stabilizing an array of stressed target proteins against precipitation. Here, we report that αS-casein acts in a similar manner to the unrelated small heat-shock proteins (sHsps) and clusterin in that it does not preserve the activity of stressed target enzymes. However, in contrast to sHsps and clusterin, α-casein does not bind target proteins in a state that facilitates refolding by Hsp70. αS-Casein was also separated into α- and α-casein, and the chaperone abilities of each of these proteins were assessed with amorphously aggregating and fibril-forming target proteins. Under reduction stress, all α-casein species exhibited similar chaperone ability, whereas under heat stress, α-casein was a poorer chaperone. Conversely, αS2-casein was less effective at preventing fibril formation by modified κ-casein, whereas α- and αS1-casein were comparably potent inhibitors. In the presence of added salt and heat stress, αS1-, α- and αS-casein were all significantly less effective. We conclude that αS1- and α-casein stabilise each other to facilitate optimal chaperone activity of αS-casein. This work highlights the interdependency of casein proteins for their structural stability.  相似文献   

8.
The process of complex formation of casein from skimmed milk and purified casein with chitosan of different molecular weights was studied. It was shown that at pH 6.3 casein micelles and parts of whey proteins coagulated with positively charged chitosan molecules with molecular weights of 45.3, 25.4, 7.7 and 1.5 kDa. As a result of ionic interaction of chitosan with skimmed milk proteins the yield of target product reached 90–92%. It consisted of all forms of casein: α-casein, β-casein, κ-casein and small amount of whey proteins.  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.
κ-casein A was fractionated into 9 subcomponents, all of which were identified as κ-casein from immunological analyses. The microheterogeneity of the subcomponents was explained by stepwise increase of their carbohydrate contents (0~4mol/mol of GalNAc, and 0~8mol/mol of NANA). The micelle-stabilizing ability of κ-casein subcomponents increased with the increase of their carbohydrate contents: the carbohydrate rich subcomponent 7 possessed twice the stabilizing ability of the carbohydrate free subcomponent 1. The sensitivity of synthetic casein micelle composed of κ-casein subcomponents and αsl-casein to the wheat germ lectin-induced aggregation also increased with the increase of their NANA contents.  相似文献   

11.
Casein (αS1, αS2, β, κ) is the major protein fraction in milk and, together with heat denatured whey proteins, responsible for gel network formation induced by acidification. Rheological measurements during gelation typically reveal a maximum storage modulus (G') at a pH close to the isoelectric point (pI) of casein (~4.6). With further decreasing pH gel stiffness decreases because of increased electrostatic repulsion, which is referred to as overacidification. In this study we investigated the effect of casein cross-linking with microbial transglutaminase on gel structure weakening induced by acidification to pH below the pI. Although enzymatic cross-linking increased the maximum stiffness (G' MAX ) of casein gels the reduction of G' during overacidification, expressed as ratio of the plateau value (G' FINAL ) to G' MAX , was more pronounced. Almost no soluble protein was detected in the serum of gels from cross-linked casein, whereas considerable amounts of αS- and κ-casein were released from reference gels below the pI. This suggests that covalent cross-linking of casein retains charged molecules within the gel network and therefore causes a higher reduction of protein-protein interactions because of higher electrostatic repulsion. Furthermore, higher amounts of uncross-linked β-casein, which was the only casein type not found in the serum, resulted in higher G' FINAL to G' MAX ratios, underlining the important contribution of β-casein to acid gel formation and prevention of gel structure weakening.  相似文献   

12.
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   

13.
BackgroundEnvironmental conditions regulate the association/aggregation states of proteins and their action in cellular compartments. Analysing protein behaviour in presence of lipid membranes is fundamental for the comprehension of many functional and dysfunctional processes. Here, we present an experimental study on the interaction between model membranes and α-casein. α-casein is the major component of milk proteins and it is recognised to play a key role in performing biological functions. The conformational properties of this protein and its capability to form supramolecular structures, like micelles or irreversible aggregates, are key effectors in functional and pathological effects.MethodsBy means of quantitative fluorescence imaging and complementary spectroscopic methods, we were able to characterise α-casein association state and the course of events induced by pH changes, which regulate the interaction of this molecule with membranes.ResultsThe study of these complex dynamic events revealed that the initial conformation of the protein critically regulates the fate of α-casein, size and structure of the newly formed aggregates and their effect on membrane structures. Disassembly of micelles due to modification in electrostatic interactions results in increased membrane structure rigidity which accompanies the formation of protein lipid flower-like co-aggregates with protein molecules localised in the external part.General significanceThese results may contribute to the comprehension of how the initial state of a protein establishes the course of events that occur upon changes in the molecular environment. These events which may occur in cells may be essential to functional, pathological or therapeutical properties specifically associated to casein proteins.  相似文献   

14.
κ-Casein and αs1-κ-casein complex with a weight ratio of unity were dissolved in 50mm cacodylate-HCl-70 mm KC1 buffer containing 0.02% of sodium azide (pH 7.1), and their size and shape in the absence and/or presence of calcium ions were observed with the electron microscope. In the absence of calcium ions, both κ-casein and αs1-κ-casein complex were spherical particles. However, the mean length of αs1-κ-casein complex (12 nm) was smaller than that of κ-casein (17 nm), which suggested that complex formation led to dissociation of the κ-casein polymer. The addition of calcium ions to the complex led to the formation of bent chains, though micelle-like aggregates were not observed even at 20 nm calcium. Comparison of the frequency distributions of αs1-κ-casein complex at 0, 5, 10, 15 and 20 mm of calcium with the calculated probability distributions suggested that most αs1-κ-casein complexes had two binding sites above 10 mm of calcium, which seemed to be essential for the stability of casein micelle.  相似文献   

15.
16.
17.
18.
Sodium dodecyl sulfate (SDS) at low concentrations considerably enhanced insulin aggregation and reduced the chaperone-like activity of purified camel αS1-casein (αS1-CN). These observed changes were the result of repulsive electrostatic interactions between both negative charged head groups of SDS and αS1-CN, and the net negative charge of insulin molecules, resulting in the greater exposure of hydrophobic patches of insulin and its enhanced aggregation. In contrast, enhanced hydrophobic interactions were primarily responsible for the conformational changes observed in insulin and αS1-CN at high SDS concentrations, resulting in increased binding of SDS and αS1-CN to insulin and its reduced aggregation.  相似文献   

19.
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   

20.
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   

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