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 31 毫秒
1.
An enzymatic method for hydrolyzing bovine milk proteins was developed. Purified milk proteins (alpha-lactalbumin, beta-lactoglobulin, and beta-casein) were hydrolyzed in 0.1 M Hepes buffer (pH 7.5) containing pronase E, aminopeptidase M, and prolidase at 37 degrees C for 20 h. Free glutamine and other amino acids were derivatized with phenylisothiocyanate and separated using a C18 Pico-Tag column. Amino acids were eluted from the column with an aqueous sodium acetate-acetonitrile gradient with detection at 254 nm. Glutamine recoveries from hydrolyzed alpha-lactalbumin, beta-lactoglobulin, and beta-casein were 78 +/- 4, 98 +/- 3, and 101 +/- 3% of the theoretical values, respectively. The recoveries of most amino acids were comparable with those obtained using acid hydrolysis, except for the recoveries of proline and acidic amino acids. These peptide bonds appeared to be resistant to enzymatic hydrolysis and also to inhibit the hydrolysis of adjacent amino acids. Free glutamine was found to be very stable (97% recovery) under the enzymatic hydrolysis conditions.  相似文献   

2.
A high-performance liquid chromatographic (HPLC) assay for a urinary hydroxyproline-containing peptide (hydroxyproline peptide, HypP) is described. This peptide represents about 50% of urinary hydroxyproline-containing peptides. Its concentration and total 4-hydroxyproline (Hyp) concentration evaluated in 325 urine samples have been shown to be closely correlated (r = 0.972; y = 0.499x − 1.5), which may indicate that the two markers provide the same information. The HypP assay, similar to Hyp assay, is carried out without hydrolysis of urine samples. After the blocking of primary amino acids by o-phthaldialdehyde (OPA) and derivatization of secondary amino acids by 9-fluorenylmethyl chloroformate (FMOC-Cl), the FMOC derivatives of HypP and 3,4-dehydroproline (internal standard) were separated on a strong anion-exchange column and detected fluorimetrically. HypP concentration was calculated by measurement of peak-area ratios of HypP and the hydroxyproline standard. The HypP/creatinine (mmol/mol) ratio in fasting urine samples from healthy adults was found to be 8.2 (S.D. = 1.6, n = 33) in 27–44-year-old premenopausal women and 6.9 (S.D. = 1.7, n = 21) in 28–49-year-old men.  相似文献   

3.
Trans-4-hydroxyproline (Hyp) in eukaryotic proteins arises from post-translational modification of proline residues. Because the modification enzyme is not present in prokaryotes, no natural means exists to incorporate Hyp into proteins synthesized in Escherichia coli. We show here that under appropriate culture conditions Hyp is incorporated co-translationally directly at proline codons in genes expressed in E. coli. The use of Hyp by E. coli protein synthesis machinery under typical culture conditions is not adequate to support protein synthesis; however, intracellular concentrations of Hyp sufficient to compensate for the poor use are achieved in media with hyperosmotic sodium chloride concentrations. Hyp incorporation was demonstrated in several recombinant proteins including human Type I collagen polypeptides. A fragment of the human collagen Type I (alpha1) polypeptide with global Hyp for Pro substitution forms a triple helix. Our results demonstrate a remarkable pliancy in the biosynthetic apparatus of bacteria that may be used more generally to incorporate novel amino acids into recombinant proteins.  相似文献   

4.
In the quantitation of amino acids by precolumn derivatization with phenylisothiocyanate, the yields of N'-phenylthiocarbamyl (PTC)-aspartate and PTC-glutamate from protein hydrolysates are often suboptimal, particularly in analyses following rapid hydrolysis at 160 degrees C. In this paper we show that these losses are due to the presence of materials extracted from the glass container during hydrolysis. In the presence of these extracts, the repeated drying and neutralization steps which precede phenylthiocarbamylation result in samples not fully solubilized by the presently used derivatizing mixtures. Thus the coupling yields for the acidic residues are highly variable. A coupling buffer with the composition 35% H2O, 30% acetonitrile, 25% pyridine, and 10% triethylamine (v/v/v/v) is an efficient solvent for all amino acids in hydrolysates and permits consistent, quantitative derivatization of all amino acids, including aspartate and glutamate.  相似文献   

5.
We performed a comparative analysis of the genome sequences of three anaerobic halophilic fermentative bacteria belonging to the order Halanaerobiales: Halanaerobium praevalens, the alkaliphilic "Halanaerobium hydrogeniformans", and the thermophilic Halothermothrix orenii to assess the amino acid composition of their proteins. Members of the Halanaerobiales were earlier shown to accumulate KCl rather than organic compatible solutes for osmotic balance, and therefore the presence of a dominantly acidic proteome was predicted. Past reports indeed showed a large excess of acidic over basic amino acids in whole-cell hydrolysates of selected members of the order. However, the genomic analysis did not show unusually high contents of acidic amino acids or low contents of basic amino acids. The apparent excess of acidic amino acids in these anaerobic halophiles reported earlier is due to the high content in their proteins of glutamine and asparagine, which yield glutamate and aspartate upon acid hydrolysis. It is thus suggested that the proteins of the Halanaerobiales, which are active in the presence of high intracellular KCl concentrations, do not possess the typical acidic signature of the 'halophilic' proteins of the Archaea of the order Halobacteriales or of the extremely halophilic bacterium Salinibacter.  相似文献   

6.
A procedure has been developed which can detect the hydroxyproline isomers trans-4-hydroxyproline (Hyp), trans-3-hydroxyproline, cis-4-hydroxyproline, and cis-3-hydroxyproline present in hydrolysates of collagens. The method involves hydrolyzing collagen, and reacting the primary amino acids with o-phthaladehyde (OPA) and the hydroxyprolines and proline with 7-chloro-4-nitrobenzo-2-oxa-1,3-diazole (NBD-Cl) which combines specifically with secondary amino acids. The proline and hydroxyprolines are then separated by thin-layer chromatography and quantified by using a scanning spectrofluorometer. The method was used to show that both trans-4-L-hydroxyproline and trans-3-L-hydroxyproline were epimerized as a function of hydrolysis time to the cis isomers. An appreciable amount of trans-3-Hyp was degraded. Hydrolysis with 6 N HCl in the presence of 6% trichloroacetic acid gave greater epimerization than the 6 N HCl alone. Alkaline hydrolysis in 0.2 M Ba(OH)2 caused more epimerization of trans-4-Hyp and trans-3-Hyp compared with acid hydrolysis but less degradation, so that alkaline hydrolysis is proposed for the evaluation of trans-3-Hyp, provided that the total of the cis and trans isomers be considered in this case.  相似文献   

7.
The sequence specificity of human skin fibroblast collagenase has been investigated by measuring the rate of hydrolysis of 16 synthetic octapeptides covering the P4 through P4' subsites of the substrate. The choice of peptides was patterned after potential collagenase cleavage sites (those containing either the Gly-Leu-Ala or Gly-Ile-Ala sequences) found in types I, II, and III collagens. The initial rate of hydrolysis of the P1-P1' bond of each peptide has been measured by quantitating the concentration of amino groups produced upon cleavage after reaction with fluorescamine. The reactions have been carried out under first-order conditions ([S] much less than KM) and kcat/KM values have been calculated from the initial rates. The amino acids in subsites P3 (Pro, Ala, Leu, or Asn), P2 (Gln, Leu, Hyp, Arg, Asp, or Val), P1' (Ile or Leu), and P4' (Gln, Thr, His, Ala, or Pro) all influence the hydrolysis rates. However, the differences in the relative rates observed for these octapeptides cannot in themselves explain why fibroblast collagenase hydrolyzes only the Gly-Leu and Gly-Ile bonds found at the cleavage site of native collagens. This supports the notion that the local structure of collagen is important in determining the location of the mammalian collagenase cleavage site.  相似文献   

8.
Some properties and the hydrolysing ability of two novel enzyme preparations, a proteolytic preparation "C" from Acremonium chrysogenum (Cephalosporium acremonium) and a peptidase preparation (the producer from the family Pseudomanadaceae), are described. The preparations can be used for obtaining protein hydrolysates with different ratios of free amino acids and peptides. The protein hydrolysis with the preparation "C" enables one to obtain hydrolysates containing 13-18% of free amino acids. The further treatment of the hydrolysates with the peptidase preparation results either in complete hydrolysis of the remained peptide fractions or in obtainment of solutions containing from 60 to 85% of free amino acids and low-molecular weight peptides.  相似文献   

9.
Substrate specificity of beta-collagenase from Clostridium histolyticum   总被引:2,自引:0,他引:2  
The substrate specificity of beta-collagenase from Clostridium histolyticum has been investigated by measuring the rate of hydrolysis of more than 50 tri-, tetra-, penta-, and hexapeptides covering the P3 to P3' subsites of the substrate. The choice of peptides was patterned after sequences found in the alpha 1 and alpha 2 chains of type I collagen. Each peptide contained either a 2-furanacryloyl (FA) or cinnamoyl (CN) group in subsite P2 or the 4-nitrophenylalanine (Nph) residue in subsite P1. Hydrolysis of the P1-P1' bond produces an absorbance change in these chromophoric peptides that has been used to quantitate the rates of their hydrolysis under first order conditions ([S] much less than KM) from kcat/KM values have been obtained. The identity of the amino acids in all six subsites (P3-P3') markedly influences the hydrolysis rates. In general, the best substrates have Gly in subsites P3 and P1', Pro or Ala in subsite P2', and Hyp, Arg, or Ala in subsite P3'. This corresponds well with the frequency of occurrence of these residues in the Gly-X-Y triplets of collagen. In contrast, the most rapidly hydrolyzed substrates do not have residues from collagen-like sequences in subsites P2 and P1. For example, CN-Nph-Gly-Pro-Ala is the best known substrate for beta-collagenase with a kcat/KM value of 4.4 X 10(7) M-1 min-1, in spite of the fact that there is neither Pro nor Ala in P2 or Hyp nor Ala in P1. These results indicate that the previously established rules for the substrate specificity of the enzyme require modification.  相似文献   

10.
AIMS: The aim of this study was to develop a method for microbial degradation of indigenous keratin wastes and to compare it with a method of alkaline hydrolysis. METHODS AND RESULTS: Native sheep skin and wool were chosen as a model mixture of collagen and keratin wastes discarded by the leather and fur industries. Suitable conditions were found for hydrolysis of this mixture by four newly isolated thermoactinomycete strains. Another set of experiments was carried out using alkaline hydrolysis of keratin wastes. It was shown that microbial hydrolysates contained predominantly low molecular peptides and amino acids, including essential ones, while the alkaline hydrolysis produced predominantly peptides of higher molecular weight. CONCLUSION: A simple and a low-cost method was proposed for rapid and effective biodegradation of keratin wastes using Thermoactinomyces strains. SIGNIFICANCE AND IMPACT OF THE STUDY: The proposed method could find application in agriculture for preparing mixtures containing valuable peptides and amino acids.  相似文献   

11.
A series of Fmoc‐Phe(4‐aza‐C60)‐OH of fullerene amino acid derived peptides have been prepared by solid phase peptide synthesis, in which the terminal amino acid, Phe(4‐aza‐C60)‐OH, is derived from the dipolar addition to C60 of the Fmoc‐Nα‐protected azido amino acids derived from phenylalanine: Fmoc‐Phe(4‐aza‐C60)‐Lys3‐OH ( 1 ), Fmoc‐Phe(4‐aza‐C60)‐Pro‐Hyp‐Lys‐OH ( 2 ), and Fmoc‐Phe(4‐aza‐C60)‐Hyp‐Hyp‐Lys‐OH ( 3 ). The inhibition constant of our fullerene aspartic protease PRIs utilized FRET‐based assay to evaluate the enzyme kinetics of HIV‐1 PR at various concentrations of inhibitors. Simulation of the docking of the peptide Fmoc‐Phe‐Pro‐Hyp‐Lys‐OH overestimated the inhibition, while the amino acid PRIs were well estimated. The experimental results show that C60‐based amino acids are a good base structure in the design of protease inhibitors and that their inhibition can be improved upon by the addition of designer peptide sequences. Copyright © 2015 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
A standard procedure for the identification of the N-terminal amino acid in N alpha-acylated proteins has been developed. After exhaustive proteolysis, the amino acids with blocked alpha-amino groups are separated from positively charged, free amino acids by ion exchange chromatography and subjected to digestion with acylase I. Amino acid analysis before and after the acylase treatment identifies the blocked N-terminal amino acid. A survey of acylamino acid substrates showed that acylase will liberate all the common amino acids except Asp, Cys or Pro from their N-acetyl-and N-butyryl derivatives, and will also catalyze the hydrolysis of N-formyl-Met and N-myristyl-Val. Thus, the procedure cannot identify acylated Asp, Cys or Pro, nor, because of the ion exchange step, N alpha-acyl-derivatives of Arg, Lys or His. Whenever the protease treatment releases free acylamino acids, the remaining amino acids should be detected. When applied to several proteins, the procedure confirmed known N-terminal acylamino acids and identified acyl-Ser in enolases from chum and coho salmon muscle and in pyruvate kinase from rabbit muscle, and acyl-Thr in phosphofructokinase from rabbit muscle. The protease-acylase assay has been used to identify blocked peptides from CNBr- or protease-treated proteins. When such peptides were treated with 1 N HCl at 110 degrees for 10 min, sufficient yields of deacylated, mostly intact, peptide were obtained to permit direct automatic sequencing. The N-terminal sequences of rabbit muscle and coho salmon enolase were determined in this way and are compared to each other and to the sequence of yeast enolase.  相似文献   

13.
A simple and fast reversed-phase high-performance liquid chromatographic method has been developed for the complete separation of 35 dimethylaminoazobenzene sulfonyl (DABS)-amino acids and by-products. This method allows simultaneous determination of primary and secondary amino acids which can be present in protein and peptide hydrolysates and also detects the presence of cysteic acid, S-sulfocysteine, hydroxyproline, taurine, norleucine, cystine, and delta-hydroxylysine. The precolumn derivatization of amino acids with dimethylaminoazobenzene sulfonyl chloride (DABS-Cl) is simple and quick (10 min at 70 degrees C) and allows the complete reaction of primary and secondary amino acids. The separation of the compounds under investigation is achieved in 25 min using a reversed-phase 3-microns Supelcosil LC-18 column at room temperature. The versatility of the proposed method is documented by amino acid determination on protein samples obtained using different hydrolysis techniques (HCl, methane-sulfonic acid, and NaOH), with attention given to the detection of tryptophan in protein samples with high sugar concentration. Furthermore, we have reported the experimental conditions necessary to apply this method to the amino acid analysis of very low amount of proteins (1 to 5 micrograms) electroeluted from a stained band after sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The stability of DABS-derivatives, the short time of analysis, the high reproducibility and sensitivity of the system, and the complete resolution of all compounds of interest make this method suitable for routine analysis. Furthermore, we have also developed a fast reversed-phase high-performance liquid chromatographic method for the complete separation of dimethylaminoazobenzene thiohydantoin (DABTH)-amino acids. The separation of the compounds under investigation is obtained, at room temperature, in less than 18 min using a reversed-phase Supelcosil LC-18 DB column, 3-micron particles, and also allows the complete separation of DABTH-Ile, DABTH-Leu, and DABTH-Norleu. The short time of analysis, together with the high reproducibility of the system and its sensitivity at picomole levels, make this method very suitable for the identification of DABTH-amino acids released during microsequencing studies of proteins and peptides with the dimethylaminoazobenzene isothiocyanate reagent. In addition, we have shown that it is possible to obtain complete separation of DABTH-amino acids also under isocratic conditions.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
A systematic investigation of the optimal temperature and hydrolysis time for amino acid analysis has been carried out under various conditions. It is found that some simplification and increase in speed relative to the conventional protocol of employing vacuum-sealed tubes and 110 C/24-72 hour hydrolysis can be achieved without loss of accuracy and performance in amino acid analyses of proteins and peptides. The effects of hydrolysis temperature and heating time on the recoveries of various labile and hydrophobic amino acids are exemplified in the hydrolysis of oxidized ribonuclease A, lysozyme and lens crystallin. The method provides a rapid processing of multiple samples within hours instead of days with the potential for the total automation of amino acid analysis starting from the preparation of protein hydrolysates.  相似文献   

15.
Methyl-, ethyl- and propyl-esters of beta-lactoglobulin, alpha-lactalbumin and beta-casein were prepared and then hydrolyzed with trypsin in various conditions. Resulting hydrolysates were analysed by SDS electrophoresis and RP-HPLC. The degree of hydrolysis of esterified samples was generally lower than those of the non-modified proteins. The highest degrees of hydrolysis were obtained at pH 7--8 with native and esterified protein samples. beta-Lactoglobulin propyl ester and beta-casein methyl ester yielded the lowest degrees of hydrolysis. Ethyl- and propyl-esters of beta-casein showed high resistance towards tryptic attack, even after 20 h of hydrolysis. SDS electrophoretic patterns of tryptic hydrolysates of native proteins showed bands corresponding to low molecular weights. Tryptic hydrolysates of esterified proteins showed bands with higher sizes. RP-HPLC profiles of tryptic hydrolysates of esterified samples showed peaks with longer elution times than those obtained with native proteins, indicating the presence of more hydrophobic peptide populations. A peptic pre-treatment improved tryptic action on esterified proteins. It resulted in a better resolution of RP-HPLC profiles and in a complete disappearance of the protein after 20 h tryptic hydrolysis.  相似文献   

16.
To perform hydrolysis with the enzyme complex from the hepatopancreas of the Kamchatka crab, a protein mixture was isolated from soybean meal by extraction at alkaline pH values. Extractable low-molecular impurities were removed by ultrafiltration and precipitation of proteins with alcohol. The amino acid composition of the obtained protein extract turned out to be similar to the composition of the fish meal traditionally used in the production of fish feeds. Analysis of the products of fermentolysis by DDS-electrophoresis, HPLC, and mass spectrometry showed a high degree of hydrolysis of soybean proteins. Depending on the time of fermentolysis, the hydrolysates contained up to 60% (18 h of hydrolysis) of free amino acids (the fraction of the weight of the hydrolyzed protein mixture) and short peptides (2–20 amino acid residues).  相似文献   

17.
High molecular weight material recovered from the culture filtrate of cell suspension cultured Pyrus communis was composed of 81% carbohydrate, 13% protein and 5% inorganic material. This material was separated into three fractions (one neutral (Fraction A) and two acidic (Fractions B and C)), by anion-exchange chromatography on DEAE-Sepharose CL-6B using a gradient of imidazole-HCl at pH 7.0. The monosaccharide and linkage composition of each fraction was determined after carboxyl reduction of uronic acid residues. From the combined results of the carbohydrate analyses, we conclude that the high molecular weight extracellular material consists of three major and two minor polysaccharides: a (fucogalacto)xyloglucan (36%) in the unbound neutral Fraction A; a type II arabinogalactan (as an arabinogalactan-protein, 29%) and an acidic (glucurono)arabinoxylan (2%) in Fraction B; and a galacturonan (33%) and a trace of heteromannan in Fraction C. The main amino acids in the proteins were Glx, Thr, Ser, Hyp/Pro and Gly. Further separation of Fraction B by solvent partition, SDS-PAGE and analysis by LC-MS/MS identified the major proteins as two chitanases, two thaumatin-like proteins, a beta-1,3-glucanase, an extracellular dermal glycoprotein and a pathogenesis-related protein.  相似文献   

18.
The use of low grade starting material for the generation of peptides with bioactivity properties is of interest. The proteins from the potato starch industry byproduct is a promising source, as several health benefits may be associated with their hydrolysates. The efficiency of selected proteases (Novo Pro‐D, Alcalase, Flavourzyme, and Papain), exhibiting different substrate specificities and cleavage action modes, to hydrolyze potato protein isolate (patatin and protease inhibitors) was investigated. Novo Pro‐D resulted in the lowest degree of hydrolysis, whereas Alcalase, Flavourzyme, and Papain exhibited a high catalytic efficiency for the hydrolysis of potato proteins. The degree of hydrolysis behaved in a concentration dependent manner. However, the end‐product profile (peptides and free amino acids) was dependent not only on the protease specificity, its cleavage action mode (endo/exo) and the availability of the intermediate substrates but also on the contribution of the protease inhibitors to the reaction kinetics through their inhibitory effects. Indeed, the dependence of the exo‐activity on the catalytic efficiency of the endo‐action of protease was shown to be significant. Papain generated more unique peptide sequences with homology assessment matching several potato proteins when compared with Flavourzyme. This can be attributed to the high exo‐peptidase activity of Flavourzyme resulting in the generation of shorter peptides which were difficult to match. Flavourzyme produced more peptides originated from patatin fraction, whereas Papain resulted in the release of more peptides corresponding to the protease inhibitor fractions. © 2016 American Institute of Chemical Engineers Biotechnol. Prog., 32:420–429, 2016  相似文献   

19.
Structures of (Pro‐Pro‐Gly)4‐Xaa‐Yaa‐Gly‐(Pro‐Pro‐Gly)4 (ppg9‐XYG) where (Xaa, Yaa) = (Pro, Hyp), (Hyp, Pro) or (Hyp, Hyp) were analyzed at high resolution using synchrotron radiation. Molecular and crystal structures of these peptides are very similar to those of the (Pro‐Pro‐Gly)9 peptide. The results obtained in this study, together with those obtained from related compounds, indicated the puckering propensity of the Hyp in the X position: (1) Hyp(X) residues involved in the Hyp(X):Pro(Y) stacking pairs prefer the down‐puckering conformation, as in ppg9‐OPG, and ppg9‐OOG; (2) Hyp(X) residues involved in the Hyp(X):Hyp(Y) stacking pairs prefer the up‐puckering conformation if there is no specific reason to adopt the down‐puckering conformation. Water molecules in these peptide crystals are classified into two groups, the 1st and 2nd hydration waters. Water molecules in the 1st hydration group have direct hydrogen bonds with peptide oxygen atoms, whereas those in the 2nd hydration group do not. Compared with globular proteins, the number of water molecules in the 2nd hydration shell of the ppg9‐XYG peptides is very large, likely due to the unique rod‐like molecular structure of collagen model peptides. In the collagen helix, the amino acid residues in the X and Y positions must protrude outside of the triple helix, which forces even the hydrophobic side chains, such as Pro, to be exposed to the surrounding water molecules. Therefore, most of the waters in the 2nd hydration shell are covering hydrophobic Pro side chains by forming clathrate structures. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 361–372, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

20.
This paper presents stable carboxypeptidase A (CPA)-glyoxyl derivatives, to be used in the controlled hydrolysis of proteins. They were produced after immobilizing-stabilizing CPA on cross-linked 6% agarose beads, activated with low and high concentrations of aldehyde groups, and different immobilization times. The CPA-glyoxyl derivatives were compared to other agarose derivatives, prepared using glutaraldehyde as activation reactant. The most stabilized CPA-glyoxyl derivative was produced using 48 h of immobilization time and high activation grade of the support. This derivative was approximately 260-fold more stable than the soluble enzyme and presented approximately 42% of the activity of the soluble enzyme for the hydrolysis of long-chain peptides (e.g., cheese whey proteins previously hydrolyzed with immobilized trypsin and chymotrypsin) and of the small substrate N-benzoylglycyl-l-phenylalanine (hippuryl-l-Phe). These results were much better than those achieved using the conventional support, glutaraldehyde-agarose. Amino acid analysis of the products of the acid hydrolysis of CPA (both soluble and immobilized) showed that approximately four lysine residues were linked on the glyoxyl agarose beads, suggesting the existence of an intense multipoint covalent attachment between the enzyme and the support. The maximum temperature of hydrolysis was increased from 50 degrees C (soluble enzyme) to 70 degrees C (most stable CPA-glyoxyl derivative). The most stable CPA-glyoxyl derivative could be efficiently used in the hydrolysis of long-chain peptides at high temperature (e.g., 60 degrees C), being able to release 2-fold more aromatic amino acids (Tyr, Phe, and Trp) than the soluble enzyme, under the same operational conditions. This new CPA derivative greatly increased the feasibility of using this protease in the production of protein hydrolysates that must be free of aromatic amino acids.  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