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.
Demasking of peptide bonds during proteolysis of β-casein and β-lactoglobulin by trypsin was monitored by the measurement of the overall spectral shift of intrinsic protein fluorescence. A clear shift of the apparent emission maxima from approximately 340–345 nm to 355–360 nm during proteolysis was observed, with a time course, which follows protein degradation and structural opening. In contrast to procedures using extrinsic fluorescence labels, this label-free procedure does not bear the risk of structural alterations. It is easy to perform, fast, and has a relatively high accuracy of determination. Proteolysis was modelled as simple two-step process with consecutive demasking and hydrolysis stages. It was shown that the fluorescence shift can be attributed to the demasking stage. Formally, kinetics of the peptide bond demasking obeys a first-order kinetic law. Both the theoretical simulations and experiment are in accordance giving the similar dependences of the hydrolysis degree on the degree of peptide bond demasking.  相似文献   

2.
3.
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   

4.
Peptides derived from hydrolysis of αS1-casein(f1-9) [αS1-CN(f1-9)] and β-CN(f193-209) with cell extracts of Lactobacillus helveticus CNRZ32 and single-peptidase mutants (ΔpepC, ΔpepE, ΔpepN, ΔpepO, and ΔpepX) were isolated by using reverse-phase high-performance liquid chromatography and were characterized by mass spectrometry. The peptides identified suggest that there was activity of an endopeptidase, distinct from previously identified endopeptidases (PepE and PepO), with specificity for peptide bonds C terminal to Pro residues. Identification of hydrolysis products derived from a carboxyl-blocked form of β-CN(f193-209) confirmed that the peptides were derived from the activity of an endopeptidase.  相似文献   

5.
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   

6.
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   

7.
We recently observed two 2,4-dinitrophenylhydrazine (DNPH)-reactive proteins of 40 and 120 kDa in the bronchoalveolar lavage fluids of rats exposed to >95% O2 for 48 h. The N-terminal sequences of these proteins were both identical over 16 amino acids with rat β-casein, which, in addition to its more common association with milk, is produced by cytotoxic T-lymphocytes, and has been found to have proinflammatory properties. Because of the inflammatory response that accompanies hyperoxic lung injury, we investigated the oxidation of bovine β-casein by HOCl. Following exposure to HOCl at 4°C for 15 min, derivatization with DNPH, washing, and digestion with trypsin, the resultant peptides were separated by reverse-phase HPLC. One peptide isolated from a peak absorbing at 365 nm was identified as AVP(Y*)PQR, corresponding to amino acids 177–183 of bovine β-casein. Analysis of the peptide by both electrospray and matrix assisted laser desorption ionization (MALDI) mass spectrometry identified a molecular ion MH+ of 1008.5 Da, which represents an increase of 178 Da from the calculated monoisotopic MH+ of the unmodified peptide of 830.45 Da. Daughter ion spectra of the doubly charged parent ion of the peptide further support the oxidation of the tyrosine to the quinone methide, with subsequent conversion to the corresponding hydrazone with DNPH. A second pair of products were identified as arising from oxidation of Y193 within the tryptic peptide constituted by amino acids 184–202, and the corresponding chymotryptic cleavage side product, 191–202. Exposure of β-casein to increasing amounts of HOCl revealed that M and Y residues were the most susceptible, although bovine β-casein contains no C, and a single W, which would not be detected by our methods. The approach described in the present report can be used to evaluate the contributions of distinct mechanisms of oxidation in other experimental or pathological models. © 1997 Elsevier Science Inc.  相似文献   

8.
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   

9.
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   

10.
A fluorescent ATP analog, β-naphthyl triphosphate, was hydrolyzed to β-naphthyl diphosphate and orthophosphate by heavy meromyosin ATPase. In the process of hydrolysis the fluorescence intensity of β-naphthyl triphosphate changed remarkably. Thus, the rate of β-naphthyl triphosphate hydrolysis is evaluated directly and continuously by measuring the time course of fluorescence intensity.In the presence of Ca2+, the Michaelis constant (Km) of β-naphthyl triphosphate hydrolysis by heavy meromyosin was similar to that of ATP hydrolysis. While, in the presence of Mg2+ the Km of β-napthyl triphosphate hydrolysis was 9.0·10−6 M, much larger than the value of ATP hydrolysis, indicating that the apparent affinity of the enzyme for β-naphthyl triphosphate is less than that for ATP.The pH dependence of β-naphthyl triphosphatase activity resembled that of ATPase activity, suggesting a similarity in the mechanism of hydrolysis of the two substrates.  相似文献   

11.
Alzheimer’s disease is the most common form of neurodegenerative diseases in humans, characterized by the progressive accumulation and aggregation of amyloid-β peptides (Aβ) in brain regions subserving memory and cognition. These 39-43 amino acids long peptides are generated by the sequential proteolytic cleavages of the amyloid-β precursor protein (APP) by β- and γ-secretases, with the latter being the founding member of a new class of intramembrane-cleaving proteases (I-CliPs) characterized by their intramembranous catalytic residues hydrolyzing the peptide bonds within the transmembrane regions of their respective substrates. These proteases include the S2P family of metalloproteases, the Rhomboid family of serine proteases, and two aspartyl proteases: the signal peptide peptidase (SPP) and γ-secretase. In sharp contrast to Rhomboid and SPP that function as a single component, γ-secretase is a multi-component protease with complex assembly, maturation and activation processes. Recently, two low-resolution three-dimensional structures of γ-secretase and three high-resolution structures of the GlpG rhomboid protease have been obtained almost simultaneously by different laboratories. Although these proteases are unrelated by sequence or evolution, they seem to share common functional and structural mechanisms explaining how they catalyze intramembrane proteolysis. Indeed, a water-containing chamber in the catalytic cores of both γ-secretase and GlpG rhomboid provides the hydrophilic environment required for proteolysis and a lateral gating mechanism controls substrate access to the active site. The studies that have identified and characterized the structural determinants critical for the assembly and activity of the γ-secretase complex are reviewed here.  相似文献   

12.
Jones BL  Poulle M 《Plant physiology》1990,94(3):1062-1070
The hydrolytic specificity of a 30 kilodalton cysteine proteinase purified from germinated barley (Hordeum vulgare L. cv Morex) was investigated using high performance liquid chromatography to characterize its hydrolysis of two small barley seed proteins, the α- and β-hordothionins. The reduced and pyridylethylated thionins were rapidly cleaved, resulting in the production of a limited number of peptides. Peptide bonds Gly9-Arg10, Cys 16-Arg17, Cys25-Ala26, and Thr34-Ser35 were most susceptible to hydrolysis, the peptide bonds Arg5-Ser6, Arg19-Gly20 in both thionins and Lys38-Cys39 in β-hordothionin and Cys29-Arg30 of α-hordothionin being broken at much slower rates. The hydrolysis patterns were highly reproducible from assay to assay and with various enzyme preparations. The specificity was apparently defined by the amino acids in the P2 position, not those immediately adjacent to the susceptible bonds. The P2 amino acid residues of the released peptides were always either leucine, valine, tyrosine, or pyridylethylcysteine. From these observations and from the rates of release of the various peptides, it appears that the barley 30 kilodalton endoproteinase has an S2 subsite that preferentially binds the leucine side chain: i.e. for hydrolyzing the peptide bond P1-P1′ in the general sequence NH2—P2-P1-P1′—COOH, the enzyme is selective for leucine and, to a lesser extent, valine and tyrosine at position P2. The barley proteinase thus resembles two other cysteine proteinases, papain and Streptococcal proteinase, in its specificity.  相似文献   

13.
The peptidases of thermophilic lactic acid bacteria have a key role in the proteolysis of Swiss cheeses during warm room ripening. To compare their peptidase activities toward a dairy substrate, a tryptic/chymotryptic hydrolysate of purified β-casein was used. Thirty-four peptides from 3 to 35 amino acids, including three phosphorylated peptides, constitute the β-casein hydrolysate, as shown by tandem mass spectrometry. Cell extracts prepared from Lactobacillus helveticus ITG LH1, ITG LH77, and CNRZ 32, Lactobacillus delbrueckii subsp. lactis ITG LL14 and ITG LL51, L. delbrueckii subsp. bulgaricus CNRZ 397 and NCDO 1489, and Streptococcus thermophilus CNRZ 385, CIP 102303, and TA 060 were standardized in protein. The peptidase activities were assessed with the β-casein hydrolysate as the substrate at pH 5.5 and 24°C (conditions of warm room ripening) by (i) free amino acid release, (ii) reverse-phase chromatography, and (iii) identification of undigested peptides by mass spectrometry. Regardless of strain, L. helveticus was the most efficient in hydrolyzing β-casein peptides. Interestingly, cell extracts of S. thermophilus were not able to release a significant level of free proline from the β-casein hydrolysate, which was consistent with the identification of numerous dipeptides containing proline. With the three lactic acid bacteria tested, the phosphorylated peptides remained undigested or weakly hydrolyzed indicating their high intrinsic resistance to peptidase activities. Finally, several sets of peptides differing by a single amino acid in a C-terminal position revealed the presence of at least one carboxypeptidase in the cell extracts of these species.  相似文献   

14.
Although proteins consist exclusively of L-amino acids, we have reported that aspartyl (Asp) 58 and Asp 151 residues of αA-crystallin of eye lenses from elderly cataract donors are highly inverted and isomerized to D-β, D-α and L-β-Asp residues through succinimide intermediates. Of these Asp isomers, large amounts of D-β- and L-β-isomers are present but the amount of D-α-isomer is not significant. The difference in abundance of the Asp isomers in the protein may be due to the rate constants for the formation of the isomers. However, the kinetics have not been well defined. Therefore, in this study, we synthesized a peptide corresponding to human αA-crystallin residues 55 to 65 (T55VLD58SGISEVR65) and its isomers in which L-α-Asp at position 58 was replaced with L-β-, D-β- and D-α-Asp and determined the rate of isomerization and inversion of Asp residues under physiological conditions (37°C, pH7.4). The rate constant for dehydration from L-α-Asp peptide to L-succinimidyl peptide was 3 times higher than the rate constant for dehydration from L-β-Asp peptide to L-succinimidyl peptide. The rate constant for hydrolysis from L-succinimidyl peptide to L-β-Asp peptide was about 5 times higher than the rate constant for hydrolysis from L-succinimidyl peptide to L-α-Asp peptide. The rate constant for dehydration from L-α-Asp peptide to L-succinimidyl peptide was 2 times higher than the rate constant for dehydration from D-α-Asp peptide to D-succinimidyl peptide. The rate constants for hydrolysis from L-succinimidyl peptide to L-β-Asp peptide and for hydrolysis from D-succinimidyl peptide to D-β-Asp peptide were almost equal. Using these rate constants, we calculated the change in the abundance ratios of the 4 Asp isomers during a human lifespan. This result is consistent with the fact that isomerized Asp residues accumulate in proteins during the ageing process.  相似文献   

15.
The platelet integrin αIIbβ3 binds to a KQAGDV motif at the fibrinogen γ-chain C terminus and to RGD motifs present in loops in many extracellular matrix proteins. These ligands bind in a groove between the integrin α and β-subunits; the basic Lys or Arg side chain hydrogen bonds to the αIIb-subunit, and the acidic Asp side chain coordinates to a metal ion held by the β3-subunit. Ligand binding induces headpiece opening, with conformational change in the β-subunit. During this opening, RGD slides in the ligand-binding pocket toward αIIb, with movement of the βI-domain β1-α1 loop toward αIIb, enabling formation of direct, charged hydrogen bonds between the Arg side chain and αIIb. Here we test whether ligand interactions with β3 suffice for stable ligand binding and headpiece opening. We find that the AGDV tetrapeptide from KQAGDV binds to the αIIbβ3 headpiece with affinity comparable with the RGDSP peptide from fibronectin. AGDV induced complete headpiece opening in solution as shown by increase in hydrodynamic radius. Soaking of AGDV into closed αIIbβ3 headpiece crystals induced intermediate states similarly to RGDSP. AGDV has very little contact with the α-subunit. Furthermore, as measured by epitope exposure, AGDV, like the fibrinogen γ C-terminal peptide and RGD, caused integrin extension on the cell surface. Thus, pushing by the β3-subunit on Asp is sufficient for headpiece opening and ligand sliding, and no pulling by the αIIb subunit on Arg is required.  相似文献   

16.
Characterization of the length dependence of end-to-end loop-closure kinetics in unfolded polypeptide chains provides an understanding of early steps in protein folding. Here, loop-closure in poly-glycine-serine peptides is investigated by combining single-molecule fluorescence spectroscopy with molecular dynamics simulation. For chains containing more than 10 peptide bonds loop-closing rate constants on the 20–100 nanosecond time range exhibit a power-law length dependence. However, this scaling breaks down for shorter peptides, which exhibit slower kinetics arising from a perturbation induced by the dye reporter system used in the experimental setup. The loop-closure kinetics in the longer peptides is found to be determined by the formation of intra-peptide hydrogen bonds and transient β-sheet structure, that accelerate the search for contacts among residues distant in sequence relative to the case of a polypeptide chain in which hydrogen bonds cannot form. Hydrogen-bond-driven polypeptide-chain collapse in unfolded peptides under physiological conditions found here is not only consistent with hierarchical models of protein folding, that highlights the importance of secondary structure formation early in the folding process, but is also shown to speed up the search for productive folding events.  相似文献   

17.
18.
A ruminant mammary cell culture that accurately reproduces mammary function in vitro would be a valuable tool in studies of ruminant lactation, With this in mind, we have examined milk protein synthesis and secretion, milk protein mRNA abundance, and hormonal responsiveness in primary cultures of mammary acini from lecturing sheep. α- and β-casein protein synthesis, β-lactoglobulin synthesis, and α-casein, β-casein, and β-lactoglobulin secretion are maintained at high levels for 8 h in culture, but then decline to approximately 25% of maximal rates between 8 and 24 h in culture, whereas synthesis of other proteins remains unaltered. The relative abundance of α-S1-casein, β-lactoglobulin, and α-lactalbumin mRNAs similarly decline between 8 and 24 h in culture. Extracellular labeled α-casein is increased fourfold in the presence of fetal calf serum (FCS). In total, FCS alters the abundance of 47 of 68 secreted proteins detected by two-dimensional electrophoresis. However, FCS and lactogenic/galactopoietic hormones had no effect on the rate of decline of mammary function and did not promote any regaining of function when present for up to 9 days in culture. These results suggest that providing its limitations are recognized, this primary cell culture system may be useful in studying some aspects of ruminant mammary function in vitro.  相似文献   

19.
The effect of autoproteolysis of Lactococcus lactis lactocepin III on its specificity towards β-casein was investigated. β-Casein degradation was performed by using either an autolysin-defective derivative of L. lactis MG1363 carrying the proteinase genes of L. lactis SK11, which was unable to transport oligopeptides, or autoproteolyzed enzyme purified from L. lactis SK11. Comparison of the peptide pools by high-performance liquid chromatography analysis revealed significant differences. To analyze these differences in more detail, the peptides released by the cell-anchored proteinase were identified by on-line coupling of liquid chromatography to mass spectrometry. More than 100 oligopeptides were released from β-casein by the cell-anchored proteinase. Analysis of the cleavage sites indicated that the specificity of peptide bond cleavage by the cell-anchored proteinase differed significantly from that of the autoproteolyzed enzyme.  相似文献   

20.
An X-prolyl-dipeptidyl peptidase has been purified from Lactobacillus sakei by ammonium sulfate fractionation and five chromatographic steps, which included hydrophobic interaction, anion-exchange chromatography, and gel filtration chromatography. This procedure resulted in a recovery yield of 7% and an increase in specificity of 737-fold. The enzyme appeared to be a dimer with a subunit molecular mass of approximately 88 kDa. Optimal activity was shown at pH 7.5 and 55°C. The enzyme was inhibited by serine proteinase inhibitors and several divalent cations (Cu2+, Hg2+, and Zn2+). The enzyme almost exclusively hydrolyzed X-Pro from the N terminus of each peptide as well as fluorescent and colorimetric substrates; it also hydrolyzed X-Ala at the N terminus, albeit at lower rates. Km s for Gly-Pro- and Lys-Ala-7-amido-4-methylcoumarin were 29 and 88 μM, respectively; those for Gly-Pro- and Ala-Pro-p-nitroanilide were 192 and 50 μM, respectively. Among peptides, β-casomorphin 1-3 was hydrolyzed at the highest rates, while the relative hydrolysis of the other tested peptides was only 1 to 12%. The potential role of the purified enzyme in the proteolytic pathway by catalyzing the hydrolysis of peptide bonds involving proline is 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