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 671 毫秒
1.
1. Zonal electrophoresis in a column stabilized by a density gradient has been applied to the small-scale fractionation of the proteins of the casein complex of cow's milk. 2. The αs- and β-fractions from the milk of individual Ayrshire cows have been shown to behave as single homogeneous proteins on electrophoresis at two pH values in starch gels. 3. The αs-fraction has been found to be indistinguishable from αs-casein prepared by Ca2+ fractionation of the same milk samples. 4. On the evidence of their electrophoretic behaviour in starch gels and their elementary analyses, α1- and αs-casein are concluded to be substantially the same protein.  相似文献   

2.
The cell wall proteinase fraction of Streptococcus cremoris HP has been isolated. This preparation did not exhibit any activity due to either specific peptidases known to be located near the outside surface of and in the membrane or intracellular proteolytic enzymes. By using thin-layer chromatography for the detection of relatively small hydrolysis products which remain soluble at pH 4.6, it was shown that β-casein is preferentially attacked by the cell wall proteinase. This was also the case when whole casein or micelles were used as the substrate. κ-casein hydrolysis is a relatively slow process, and αs-casein degradation appeared to proceed at an extremely low rate. These results could be confirmed by using 14CH3-labeled caseins. A relatively fast and linear initial progress of 14CH3-labeled β-casein degradation is not inhibited by αs-casein and only slightly by κ-casein at concentrations of these components which reflect their stoichiometry in the micelles. Possible implications of β-casein degradation for growth of the organism in milk are discussed.  相似文献   

3.
Partially purified cell wall proteinases of eight strains of Streptococcus cremoris were compared in their action on bovine αs1-, β-, and κ-casein, as visualized by starch gel electrophoresis, sodium dodecyl sulfate-polyacrylamide gel electrophoresis, and thin-layer chromatography. Characteristic degradation profiles could be distinguished, from which the occurrence of two proteinases, represented by strain HP and strain AM1, was concluded. The action of the HP-type proteinase P1 (also detectable in strains Wg2, C13, and TR) was established by electrophoretic methods to be directed preferentially towards β-casein. The AM1-type proteinase PIII (also detectable in strain SK11) was also able to degrade β-casein, but at the same time split αs1- and κ-casein more extensively than did PI. Strain FD27 exhibited mainly PI activity but also detectable PIII degradation characteristics. The cell wall proteinase preparation of strain E8 showed low PI as well as low PIII activity. All proteinase preparations produced from κ-casein positively charged degradation products with electrophoretic mobilities similar to those of degradation products released by the action of the milk-clotting enzyme chymosin. The differences between PI and PIII in mode of action, as detected by gel electrophoresis and thin-layer chromatography, were reflected by the courses of the initial degradation of methyl-14C-labeled β-casein and by the effect of αs1- plus κ-casein on these degradations. The results are discussed in the light of previous comparative studies of cell wall proteinases in strains of S. cremoris and with respect to the growth of this organism in milk.  相似文献   

4.
5.
This work evaluated the angiotensin-converting-enzyme (ACE)-inhibitory activities of a bovine sodium caseinate fermentate generated using the proteolytic capabilities of the porcine small intestinal isolate Lactobacillus animalis DPC6134 (NCIMB deposit 41355). The crude 10-kDa L. animalis DPC6134 fermentate exhibited ACE-inhibitory activity of 85.51% (±15%) and had a 50% inhibitory concentration (IC50) of 0.8 mg protein/ml compared to captopril, which had an IC50 value of 0.005 mg/ml. Fractionation of the crude L. animalis DPC6134 fermentate by membrane filtration and reversed-phase high-performance liquid chromatography (HPLC) generated three bioactive fractions from a total of 72 fractions. Fractions 10, 19, and 43 displayed ACE-inhibitory activity percentages of 67.53 (±15), 83.71 (±19), and 42.36 (±11), respectively, where ACE inhibition was determined with 80 μl of the fractions with protein concentrations of 0.5 mg/ml. HPLC and mass spectrometry analysis identified 25 distinct peptide sequences derived from α-, β-, and κ-caseins. In silico predictions, based on the C-terminal tetrapeptide sequences, suggested that peptide NIPPLTQTPVVVPPFIQ, corresponding to β-casein f(73-89); peptide IGSENSEKTTMP, corresponding to αs1-casein f(201212); peptide SQSKVLPVPQ, corresponding to β-casein f(166-175); peptide MPFPKYPVEP, corresponding to β-casein f(124133); and peptide EPVLGPVRGPFP, corresponding to β-casein f(210-221), contained ACE-inhibitory activities. These peptides were chosen for chemical synthesis to confirm the ACE-inhibitory activity of the fractions. Chemically synthesized peptides displayed IC50 values in the range of 92 μM to 790 μM. Additionally, a simulated gastrointestinal digestion confirmed that the ACE-inhibitory 10-kDa L. animalis DPC6134 fermentation was resistant to a cocktail of digestive enzymes found in the gastrointestinal tract.  相似文献   

6.
During the ripening of Gouda-type cheese, two kinds of endopeptidases were found to participate in the degradation of αs1-CN(f1-23), a specific product from αs1-casein hydrolyzed by chymosin. One of the endopeptidases, lactic acid bacteria endopeptidase (LEP-II), which can recognize the size of its substrates, has already been purified and characterized (T. R. Yan, N. Azuma, S. Kaminogawa, and K. Yamauchi, Eur. J. Biochem. 163:259-265, 1987). The other endopeptidase, LEP-I, was purified to homogeneity by conventional chromatographic techniques from Streptococcus cremoris H61. The enzyme appeared to be monomeric, with an apparent molecular weight of 98,000, and its isoelectric point was 5.1. For the hydrolysis of αs1-CN(f1-23), the enzyme had an optimum pH and temperature of 7.0 to 7.5 and 40°C, respectively. Its activity was inhibited by such chelating agents as EDTA and 1,10-phenanthrolin, and it could be fully reactivated by Mn2+. Inhibitors specific for serine and thiol proteases had no effect on the protease activity. The enzyme showed a high affinity toward the Glu-Asn peptide bond of αs1-CN(f1-23) and αs1-CN(f91-100) but showed no hydrolysis activity toward αs1-CN(f1-52), αs1-CN(61-122), αs1-CN(136-196), αs1-casein, β-casein, κ-casein, α-lactalbumin, and β-lactoglobulin. The Km and Vmax of LEP-I for αs1-CN(f1-23) were 14.2 pM and 139 U, respectively.  相似文献   

7.
Caseins, the main milk proteins, interact with colloidal calcium phosphate to form the casein micelle. The mesostructure of this supramolecular assembly markedly influences its nutritional and technological functionalities. However, its detailed molecular organization and the cellular mechanisms involved in its biogenesis have been only partially established. There is a growing body of evidence to support the concept that αs1-casein takes center stage in casein micelle building and transport in the secretory pathway of mammary epithelial cells. Here we have investigated the membrane-associated form of αs1-casein in rat mammary epithelial cells. Using metabolic labelling we show that αs1-casein becomes associated with membranes at the level of the endoplasmic reticulum, with no subsequent increase at the level of the Golgi apparatus. From morphological and biochemical data, it appears that caseins are in a tight relationship with membranes throughout the secretory pathway. On the other hand, we have observed that the membrane-associated form of αs1-casein co-purified with detergent-resistant membranes. It was poorly solubilised by Tween 20, partially insoluble in Lubrol WX, and substantially insoluble in Triton X-100. Finally, we found that cholesterol depletion results in the release of the membrane-associated form of αs1-casein. These experiments reveal that the insolubility of αs1-casein reflects its partial association with a cholesterol-rich detergent-resistant microdomain. We propose that the membrane-associated form of αs1-casein interacts with the lipid microdomain, or lipid raft, that forms within the membranes of the endoplasmic reticulum, for efficient forward transport and sorting in the secretory pathway of mammary epithelial cells.  相似文献   

8.
9.
H. Bovenhuis  J. I. Weller 《Genetics》1994,137(1):267-280
Maximum likelihood methodology was used to estimate effects of both a marker gene and a linked quantitative trait locus (QTL) on quantitative traits in a segregating population. Two alleles were assumed for the QTL. In addition to the effects of genotypes at both loci on the mean of the quantitative trait, recombination frequency between the loci, frequency of the QTL alleles and the residual standard deviation were also estimated. Thus six parameters were estimated in addition to the marker genotype means. The statistical model was tested on simulated data, and used to estimate direct and linked effects of the milk protein genes, β-lactoglobulin, κcasein, and β-casein, on milk, fat, and protein production and fat and protein percent in the Dutch dairy cattle population. β-Lactoglobulin had significant direct effects on milk yield and fat percent. κ-Casein had significant direct effects on milk yield, protein percent and fat yield. β-Casein had significant direct effects on milk yield, fat and protein percent and fat and protein yield. Linked QTL with significant effects on fat percent were found for κ-casein and β-casein. Since the β-casein and κ-casein genes are closely linked, it is likely that the same QTL was detected for those two markers. Further, a QTL with a significant effect on fat yield was found to be linked to κ-casein and a QTL with a significant effect on protein yield was linked to β-lactoglobulin.  相似文献   

10.
Lactobacillus delbrueckii subsp. lactis ACA-DC 178, which was isolated from Greek Kasseri cheese, produces a cell-wall-bound proteinase. The proteinase was removed from the cell envelope by washing the cells with a Ca2+-free buffer. The crude proteinase extract shows its highest activity at pH 6.0 and 40°C. It is inhibited by phenylmethylsulfonyl fluoride, showing that the enzyme is a serine-type proteinase. Considering the substrate specificity, the enzyme is similar to the lactococcal PI-type proteinases, since it hydrolyzes β-casein mainly and α- and κ-caseins to a much lesser extent. The cell-wall-bound proteinase from L. delbrueckii subsp. lactis ACA-DC 178 liberates four main peptides from β-casein, which have been identified.  相似文献   

11.
Casein is used as an additive in binders or paints and as such exhibits unique properties which might be based on the properties of certain subproteins in the complex whole casein mixture. Therefore, the separation of whole casein (CN) from cow milk was performed on a gram scale in order to yield sufficient amounts of the protein subfractions α-, β-, and κ-casein for further testing utilizing fast protein liquid chromatography (FPLC) and preceding enrichment in the case of κ-casein. Construction chemical grade casein, which differs in quality from dairy grade casein, was used for separation because of our interest in the proteins responsible for plastification of cementitious systems such as mortar. The solubilized proteins were separated chromatographically via ion exchange chromatography (IEX) and the subsequently desalted protein fractions were tested for purity by isoelectric focusing (IEF).  相似文献   

12.
Higher eukaryotes encode various Y-family DNA polymerases to perform global DNA lesion bypass. To provide complete mutation spectra for abasic lesion bypass, we employed short oligonucleotide sequencing assays to determine the sequences of abasic lesion bypass products synthesized by human Y-family DNA polymerases eta (hPolη), iota (hPolι) and kappa (hPolκ). The fourth human Y-family DNA polymerase, Rev1, failed to generate full-length lesion bypass products after 3 h. The results indicate that hPolι generates mutations with a frequency from 10 to 80% during each nucleotide incorporation event. In contrast, hPolη is the least error prone, generating the fewest mutations in the vicinity of the abasic lesion and inserting dAMP with a frequency of 67% opposite the abasic site. While the error frequency of hPolκ is intermediate to those of hPolη and hPolι, hPolκ has the highest potential to create frameshift mutations opposite the abasic site. Moreover, the time (t50bypass) required to bypass 50% of the abasic lesions encountered by hPolη, hPolι and hPolκ was 4.6, 112 and 1 823 s, respectively. These t50bypass values indicate that, among the enzymes, hPolη has the highest abasic lesion bypass efficiency. Together, our data suggest that hPolη is best suited to perform abasic lesion bypass in vivo.  相似文献   

13.
We used a gene knockout approach to elucidate the specific roles played by the Jun-N-terminal kinase (JNK) and NF-κB pathways downstream of TNF-α in the context of α(2) type I collagen gene (COL1A2) expression. In JNK1−/−-JNK2−/− (JNK−/−) fibroblasts, TNF-α inhibited basal COL1A2 expression but had no effect on TGF-β-driven gene transactivation unless jnk1 was introduced ectopically. Conversely, in NF-κB essential modulator−/− (NEMO−/−) fibroblasts, lack of NF-κB activation did not influence the antagonism exerted by TNF-α against TGF-β but prevented repression of basal COL1A2 gene expression. Similar regulatory mechanisms take place in dermal fibroblasts, as evidenced using transfected dominant-negative forms of MKK4 and IKK-α, critical kinases upstream of the JNK and NF-κB pathways, respectively. These results represent the first demonstration of an alternate usage of distinct signaling pathways by TNF-α to inhibit the expression of a given gene, COL1A2, depending on its activation state.  相似文献   

14.
Template-based studies on antimicrobial peptide (AMP) derivatives obtained through manipulation of the amino acid sequence are helpful to identify properties or residues that are important for biological activity. The present study sheds light on the importance of specific amino acids of the milk-derived αs2-casein f(183–207) peptide to its antibacterial activity against the food-borne pathogens Listeria monocytogenes and Cronobacter sakazakii. Trimming of the peptide revealed that residues at the C-terminal end of the peptide are important for activity. Removal of the last 5 amino acids at the C-terminal end and replacement of the Arg at position 23 of the peptide sequence by an Ala residue significantly decreased activity. These findings suggest that Arg23 is very important for optimal activity of the peptide. Substitution of the also positively charged Lys residues at positions 15 and 17 of the αs2-casein f(183–207) peptide also caused a significant reduction of the effectiveness against C. sakazakii, which points toward the importance of the positive charge of the peptide for its biological activity. Indeed, simultaneous replacement of various positively charged amino acids was linked to a loss of bactericidal activity. On the other hand, replacement of Pro residues at positions 14 and 20 resulted in a significantly increased antibacterial potency, and hydrophobic end tagging of αs2-casein f(193–203) and αs2-casein f(197–207) peptides with multiple Trp or Phe residues significantly increased their potency against L. monocytogenes. Finally, the effect of pH (4.5 to 7.4), temperature (4°C to 37°C), and addition of sodium and calcium salts (1% to 3%) on the activity of the 15-amino-acid αs2-casein f(193–207) peptide was also determined, and its biological activity was shown to be completely abolished in high-saline environments.  相似文献   

15.
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   

16.
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   

17.
The shape of mechanically pierced giant vesicles is studied to obtain the elastic modulus of Gaussian curvature of egg lecithin bilayers. It is argued that such experiments are governed by an apparent modulus, ¯κapp, not the true modulus of Gaussian curvature, ¯κ. A theory of ¯κapp is proposed, regarding the pierced bilayer vesicle as a closed monolayer vesicle. The quantity measured, i.e. ¯κapp/κ, where κ is the rigidity, agrees satisfactorily with the theory. We find ¯κapp = -(1.9 ± 0.3) · 10-12 erg (on the basis of κ = (2.3 ± 0.3) · 10-12 erg). The result may have implications for bilayer fusion.  相似文献   

18.
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   

19.
We previously demonstrated that trans-10, cis-12 (10,12) conjugated linoleic acid (CLA) induced inflammation and insulin resistance in primary human adipocytes by activating nuclear factor κB (NFκB) and extracellular signal-related kinase (ERK) signaling. In this study, we demonstrated that the initial increase in intracellular calcium ([Ca2+]i) mediated by 10,12 CLA was attenuated by TMB-8, an inhibitor of calcium release from the endoplasmic reticulum (ER), by BAPTA, an intracellular calcium chelator, and by D609, a phospholipase C (PLC) inhibitor. Moreover, BAPTA, TMB-8, and D609 attenuated 10,12 CLA–mediated production of reactive oxygen species (ROS), activation of ERK1/2 and cJun-NH2-terminal kinase (JNK), and induction of inflammatory genes. 10,12 CLA–mediated binding of NFκB to the promoters of interleukin (IL)-8 and cyclooxygenase (COX)-2 and induction of calcium-calmodulin kinase II (CaMKII) β were attenuated by TMB-8. KN-62, a CaMKII inhibitor, also suppressed 10,12 CLA–mediated ROS production and ERK1/2 and JNK activation. Additionally, KN-62 attenuated 10,12 CLA induction of inflammatory and integrated stress response genes, increase in prostaglandin F, and suppression of peroxisome proliferator activated receptor γ protein levels and insulin-stimulated glucose uptake. These data suggest that 10,12 CLA increases inflammation and insulin resistance in human adipocytes, in part by increasing [Ca2+]i levels, particularly calcium from the ER.  相似文献   

20.
1,25-Dihydroxyvitamin D (1,25(OH)2D3) is known to suppress NF-κB activity, but the underlying mechanism remains poorly understood. Here we show that the vitamin D receptor (VDR) physically interacts with IκB kinase β (IKKβ) to block NF-κB activation. 1,25(OH)2D3 rapidly attenuates TNFα-induced p65 nuclear translocation and NF-κB activity in a VDR-dependent manner. VDR overexpression inhibits IKKβ-induced NF-κB activity. GST pull-down assays and coimmunoprecipitation experiments demonstrated that VDR physically interacts with IKKβ and that this interaction is enhanced by 1,25(OH)2D3. Protein mapping reveals that VDR-IKKβ interaction occurs between the C-terminal portions of the VDR and IKKβ proteins. Reconstitution of VDR−/− cells with the VDR C terminus restores the ability to block TNFα-induced NF-κB activation and IL-6 up-regulation. VDR-IKKβ interaction disrupts the formation of the IKK complex and, thus, abrogates IKKβ phosphorylation at Ser-177 and abolishes IKK activity to phosphorylate IκBα. Consequently, stabilization of IκBα arrests p65/p50 nuclear translocation. Together, these data define a novel mechanism whereby 1,25(OH)2D3-VDR inhibits NF-κB activation.  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