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 15 毫秒
1.
To study whether the phosphoserine residue is associated with the antigenicity of bovine αs1- casein, we examined the antigenic reactivity of dephosphorylated αs1-casein, peptide 1~25 from bovine β-casein and three chemical reagents with IgG antibody specific to native αs1-casein by an enzyme-linked immunosorbent assay.

The reaction between native αs1-casein and its IgG antibody was inhibited more strongly by native αs1-casein than by dephosphorylated αs1-casein. Peptide 1~25, having a phosphoserine residue-concentrated region from bovine β-casein, noticeably inhibited the reaction between native αs1 -casein and its antibody. Furthermore, the O-phospho-l-serine residue inhibited the reaction of peptide 61~123 with anti-native αs1-casein antibody, although l-serine and sodium phosphate showed no measurable inhibition.

These results suggest that the phosphoserine residue associated with part of an antigenic site in bovine αsl-casein.  相似文献   

2.
κ-Casein components having various carbohydrate contents were prepared by diethylaminoethyl-cellulose chromatography and the interactions of each κ-casein component both with αs1-casein and with β-casein were examined by Sepharose 4B gel chromatography, ultra-centrifugal experiments and viscosity measurements. Each κ-casein component could form complex with αs1- and β-casein in the absence and presence of CaCl2. Molecular weight of complexes of unfractionated κ-casein both with αs1-casein and with κ-casein were about 70 × 104 at 37°C in the absence of CaCl2, while those of complexes of each κ-casein component with αs1 and β-casein were about 50 × 104. Stokes radii of complexes increased with increasing calcium ion. While sedimentation coefficient at 37°C of complex with β-casein had almost the same value, those of complexes with αs1-casein decreased with increase of carbohydrate content of κ-casein components. Intrinsic viscosity of complex of κ-casein component having much carbohydrate was almost the same among tested temperatures. It is suggested that heterogeneity of κ-casein is necessary to form large complex and that the carbohydrate moiety of κ-casein contributes the stability of casein complex.  相似文献   

3.
It was indicated from fluorescence spectra and fluorescence titration that a hydrophobic probe, 1-anilino-8-naphthalenesulfonate (ANS), binds to casein components (αs-, β- and κ-caseins). Fluorescence intensity and affinity of ANS-κ-casein complex were larger than that of ANS-αs- and ANS-β-casein complexes. Enhancements of fluorescence intensity of complexes of casein components were observed by the addition of KCI or CaCl2. Reason for the enhancement was postulated to be the increase of the quantum yield of the ANS fluorescence caused by the environmental change of ANS binding region of the casein components.

Marked increase of sedimentation coefficient of β-casein in the presence of KCl or CaCl2 at 10°C was caused by the addition of ANS. This may be responsible for the stimulation of the Ca-dependent precipitation of β-casein by the addition of ANS.

It was found that αs · κ-association was prevented by ANS and that hydrophobic interaction have an important role for αs · κ-association.  相似文献   

4.
5.
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   

6.
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   

7.
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   

8.
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   

9.
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   

10.
Whole casein, αs-casein and k-casein were dephosphorylated with a phosphoprotein phosphatase prepared from beef spleen and their calcium-binding capacities were compared with those of respective native caseins by a ultracentrifugal method.

The bindings of the calcium to 94% dephosphorylated whole casein and to 97 % dephosphorylated αs-casein at neutral pH were approximately one third of those to respective native caseins. The decrease of calcium-binding capacity of k-casein due to dephosphorylation was also significant.

The effect of pH on the state and the calcium-binding capacity of dephosphorylated caseins was also examined and the role of organic phosphate groups of casein as calcium-binding sites was discussed.  相似文献   

11.
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   

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.
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   

14.
Pyrenebutyrate-conjugated αs1-casein was prepared and the complex formation between αs1- and κ-casein polymers was investigated by fluorescence polarization. The complex formation was also investigated by a microcalorimetric technique. The positive enthalpy and entropy changes and endothermic nature suggested the hydrophobic interaction between αs1- and κ-casein polymers.

The degree of polarization of κ-casein polymer decreased with the addition of 1-anilino-8-naphthalenesulfonate (ANS), while that of αs1-casein polymer and αs1-κ-casein complex was invariant. Moreover the reaction of κ-casein polymer and ANS was exothermic. These facts suggested that the intermolecular hydrophobic regions in κ-casein polymer were disrupted by the adsorption of ANS. The rotational relaxation time of pyrenebutyrate conjugated complex between cyanoethyl-κ-casein and αs1-casein polymer was smaller than that of cyanoethyl-κ-casein alone. From these results, it was postulated that the dissociation of κ-casein polymer by the complex formation with αs1-casein polymer might be caused by the disruption of the intermolecular hydrophobic bonds in κ-casein polymer.  相似文献   

15.
αsl-Casein can be made either soluble or insoluble by adjusting the concentration of coexisting calcium ions. In this study, we tried to make a soluble-insoluble interconvertible enzyme through the formation of a conjugate of an enzyme and αsl-casein using a heterobifunctional crosslinking reagent, N-succinimidyl 3-(2-pyridyldithio)propionate. The conjugate of phosphoglyceromutase and native αs1-casein did not exhibit sufficient calcium-dependent precipitation. However, conjugates of enzymes (phosphoglyceromutase, enolase or peroxidase) and αsl-casein polymerized by transglutaminase precipitated almost completely in the presence of more than 50 mM CaCl2. Most of the enzyme conjugates precipitated as calcium caseinates could be solubilized reversibly with EDTA, without a significant loss of activity. A mixture of the enzyme ? polymerized αs1-casein conjugates prepared with phosphoglyceromutase, enolase and pyruvate kinase could catalyze sequential reactions which convert d-3-phosphoglycerate into pyruvate with the same efficiency as a mixture of free enzymes. These results indicate that conjugates of enzymes and polymerized αs1-casein can be useful as soluble-insoluble interconvertible enzymes.  相似文献   

16.
The interaction of αs1-casein with β-, dephosphorylated β-,γ- and R-caseins was studied. It was proved by the sedimentation velocity experiments that αs1-casein formed a complex with each of these components at 25±C in the presence of 3 mm CaCl2.

In the presence of 10 mm CaCl2, β- and dephosphorylated β-casein prevented the precipitation of αs1-casein and gave micelle-like turbid solutions. However, γ- and R-caseins, fragments of β-casein, did not stabilize αs1-casein. It was concluded from these results that α-casein interacted with αs1-casein through its hydropholic region corresponding to R-casein and that hydrophilic region of β-casein was responsible for the stabilization of αs1-casein.  相似文献   

17.
αs1-Casein was dissolved in 50 mm cacodylate-HCl-70 mm KC1 buffer containing 0.02% of sodium azide (pH 7.1), and the size and shape of αs1-caseins in the absence and presence of calcium ions were observed with the electron microscope. In the absence of calcium ions, most αs1-caseins existed as spherical particles of which the smallest diameter was 5~6 nm. The particles were polymerized into bent chains by adding 3 mm calcium. It seemed that the smallest particles were the polymerizing units. The mean length of αs1-casein in the absence of calcium was 8.7 nm, and it increased as the calcium concentration increased. From these results, it was speculated that αs1-casein in the absence of calcium had one binding site and the calcium-induced conformational changes produced a second binding site. The probability distributions were calculated with the above speculation, and compared with the frequency distributions obtained from electron micrographs. It was suggested from the comparisons that the second binding site produced in αs1-casein might be responsible for the calcium-induced aggregation.  相似文献   

18.
The first genetic map for Hevea spp. (2n=36) is presented here. It is based on a F1 progeny of 106 individuals allowing the construction of a female, a male, and a synthetic map according to the pseudo-testcross strategy. Progeny were derived from an interspecific cross between PB260, a H. brasiliensis cultivated clone, and RO38, a H. brasiliensis×H. benthamiana interspecific hybrid clone. The disomic inheritance observed for all the codominant markers scattered on the 2n=36 chromosomes revealed that Hevea behaves as diploids. Homologous linkage groups between the two parental maps were merged using bridge loci. A total of 717 loci constituted the synthetic map, including 301 RFLPs, 388 AFLPs, 18 microsatellites, and 10 isozymes. The markers were assembled into 18 linkage groups, thus reflecting the basic chromosome number, and covered a total distance of 2144 cM. Nine markers were found to be unlinked. Segregation distortion was rare (1.4%). Average marker density was 1 per 3 cM. Comparison of the distance between loci in the parental maps revealed significantly less meiotic recombination in the interspecific hybrid male parent than in the female parent. Hevea origin and genome organisation are discussed. Received: 2 February 1999 / Accepted: 11 March 1999  相似文献   

19.
The aim of this study was to detect new polymorphisms in the bovine β‐casein (β‐CN) gene and to evaluate association of (new) β‐CN protein variants with milk production traits and milk protein composition. Screening of the β‐CN gene in genomic DNA from 72 Holstein Friesian (HF) bulls resulted in detection of 19 polymorphisms and revealed the presence of β‐CN protein variant I in the Dutch HF population. Studies of association of β‐CN protein variants with milk composition usually do not discriminate protein variant I from variant A2. Association of β‐CN protein variants with milk composition was studied in 1857 first‐lactation HF cows and showed that associations of protein variants A2 and I were quite different for several traits. β‐CN protein variant I was significantly associated with protein percentage and protein yield, and with αs1‐casein (αs1‐CN), αs2‐casein (αs2‐CN), κ‐casein (κ‐CN), α‐lactalbumin (α‐LA), β‐lactoglobulin (β‐LG), casein index and casein yield. Inferring β‐κ‐CN haplotypes showed that β‐CN protein variant I occurred only with κ‐CN variant B. Consequently, associations of β‐κ‐CN haplotype IB with protein percentage, κ‐CN, α‐LA, β‐LG and casein index are likely resulting from associations of κ‐CN protein variant B, while associations of β‐κ‐CN haplotype IB with αs1‐CN and αs2‐CN seem to be resulting from associations of β‐CN variant I.  相似文献   

20.
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   

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