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 234 毫秒
1.
An extracellular alkaline proteinase produced by Candida lipolytica was purified through iso-propanol and ammonium sulfate precipitation, decolorization with DEAE-cellulose, gel filtration with Sephadex G–100 and ion-exchange chromatography on DEAE-Sephadex A–50. The optimum pH of its caseinolytic activity was 9.0, and this activity was completely inactivated with DFP but not with chelating reagents, PCMB, STI, TLCK, TPCK, or SSI. This enzyme also hydrolyzed salmin and synthetic esters, such as Bz. Arg. OEt, Bz. His. OMe, Tos. Lys. OMe or Ac. Tyr. OEt, and the optimum pH of its esterase activity was 8.0. The molecular weight of the enzyme was estimated to be about 30,000 by the gel filtration method. These facts indicated that this enzyme was distinguishable from other microbial alkaline proteinases so far studied.  相似文献   

2.
The substrate specificities of four Cl?-activated arginine aminopeptidases purified from the livers and inflammatory exudates of the rat, human fetal livers, and human erythrocytes were studied using peptides and N-l-aminoacyl-2-naphthylamides as substrates. With 2-naphthylamide substrates, these aminopeptidases showed similar substrate specificity; only the derivatives of Arg and Lys were measurably hydrolyzed. Di- and tripeptides with Arg or Lys as the N-terminal residue were readily split by the enzymes from the livers and inflammatory exudates of the rat and human fetal livers but oligopeptides were not hydrolyzed. Arg- and Lys-peptides were also hydrolyzed by the erythrocyte enzyme but this enzyme additionally split several other peptides, oligopeptides being hydrolyzed at internal bonds. The following properties were similar for all four arginine aminopeptidases: Dipeptides were preferred over tripeptides both in substrate binding and catalysis. The rat and human liver, rat exudate, and human erythrocyte enzymes revealed similar Km values for the best substrates, the values increasing in the following order: ArgPhe, ArgTrp, ArgLys < ArgVal, ArgGly, Arg-2-naphthylamide < ArgGlyGly. The kcat values were also similar for the four arginine aminopeptidases. Arg-2-naphthylamide was by far the most rapidly hydrolyzed substrate by all enzymes followed by ArgPhe and ArgTrp. With peptide substrates the highest Cl? activation (10–20%) was found with ArgPhe and ArgTrp. With Arg-2-naphthylamide, however, the activating effect of 0.2 m Cl? was severalfold. The hydrophobicity of the C-terminal residue of the substrate seemed to play an important role both in the Cl? effect and substrate catalysis. Substrate binding, however, also depended on the charged groups of the substrate. Evidently Arg-2-naphthylamide and the peptides were hydrolyzed at the same active center but the mechanisms involved in the hydrolyses of chromogenic substrates and peptides may be different. It was also concluded that the less specific Cl?-activated enzyme from human erythrocytes does not belong to the same group of Cl?-activated arginine aminopeptidases that show a narrow substrate specificity.  相似文献   

3.
The presence of three trypsinogens (Try-III, Try-I, and Try-II) in the mouse is demonstrated by DEAE-cellulose column chromatography. Two genetic variants of Try-I are detected, because the activity of Try-I is different between the Mol-A strain and seven other strains. The Prt-3 locus controls the activity of Try-I. The Prt-3a gene exists in CFO, BS, KR, BALB/cJ, C57BL/6J, CBA/J, and 129/Sv-S1-CP strains, whereas the Prt-3b gene is present only in the Mol-A strain. Each Try-I from the CFO or Mol-A strain was purified. The fact that Try-I activity of the Mol-A strain is much higher than those of other strains is because of a difference in the specific activity; the ratio of the Kcat (sec-1) value with Tos-Arg-OMe to that with Bz-Arg-OEt is different between the variants from the CFO and those from the Mol-A strains, being much higher in the Mol-A strain. Also, chicken ovomucoid inhibited Try-I activity of the CFO strain at a molar ratio of one ovomucoid to two trypsins; Try-I activity of the Mol-A strain was only 50% inhibited even with an excess of ovomucoid. There was no difference between genetic variants of Try-I in molecular weight, Km values with Bz-Arg-OEt or Tos-Arg-OMe, pH optimum profile, or inhibition by soybean trypsin inhibitor.  相似文献   

4.
Mouse pancreatic proteases were analyzed by one- and two-dimensional electrophoresis. Active proteases that existed in the luminal fluid were separated into at least eight bands in 8% polyacrylamide gel. Pancreatic proteases activated by intestinal extract were separated into at least seven bands. The mobilities of these bands were exactly the same as those of proteases in the luminal fluid except for those of the most cathodal band. Two kinds of trypsin (Try-I group and Try-II) and one kind of chymotrypsin (Chy-I) were determined by specific and nonspecific protease staining. Try-I group and Try-II were derived from different trypsinogens (Try G-I group and Try G-II), whereas Chy-I was derived from a single chymotrypsinogen (Chy G). Although Try G-II was activated by both intestinal extract and by bovine trypsin, Try G-I group activated only by intestinal extract. Intestinal-activating factors were analyzed by two-dimensional electrophoresis. Mouse enterokinase (enteropeptidase EC 3.4.4.8), which can activate bovine trypsinogen, had a slow mobility. In the intestine of the mouse there are several activating factors in addition to enterokinase. Although it is unclear what intestinal-activating factors can activate Chy G, there is a factor that can convert chymotrypsinogen into chymotrypsin directly. These data suggest that intestinal-activating factors play an important role in the activating mechanisms of mouse pancreatic zymogens.  相似文献   

5.
The uptake of glycine-N15 into the trypsinogen and chymotrypsinogen of mouse pancreas is much higher than that into any ribonucleoprotein component of the pancreas that has so far been investigated.  相似文献   

6.
The capability of the cysteine proteases ficin, papain and clostripain to form peptide bonds in frozen aqueous solutions was investigated. Freezing the reaction mixture resulted in increased peptide yields in kinetically controlled coupling of Bz–Arg–OEt with various amino acid amides and dipeptides. Under these conditions, peptide yields increased up to 70% depending on the enzyme and the amino component used. Enzyme-catalysed peptide syntheses were carried out under optimized reaction conditions (temperature, amino component concentration and pH before freezing) using the condensation of Bz–Arg–OEt and H–Leu–NH2 as a model reaction.  相似文献   

7.
Isolation and characterization of a cDNA encoding rat cationic trypsinogen   总被引:4,自引:0,他引:4  
A cDNA encoding rat cationic trypsinogen has been isolated by immunoscreening from a rat pancreas cDNA library. The protein encoded by this cDNA is highly basic and contains all of the structural features observed in trypsinogens. The amino acid sequence of rat cationic trypsinogen is 75% and 77% homologous to the two anionic rat trypsinogens. The homology of rat cationic trypsinogen to these anionic trypsinogens is lower than its homology to other mammalian cationic trypsinogens, suggesting that anionic and cationic trypsins probably diverged prior to the divergence of rodents and ungulates. The most unusual feature of this trypsinogen is the presence of an activation peptide containing five aspartic acid residues, in contrast to all other reported trypsinogen activation peptides which contain four acidic amino acid residues. Comparisons of cationic and anionic trypsins reveal that the majority of the charge changes occur in the C-terminal portion of the protein, which forms the substrate binding site. Several regions of conserved charge differences between cationic and anionic trypsins have been identified in this region, which may influence the rate of hydrolysis of protein substrates.  相似文献   

8.
An Mn2+-activated phosphoprotein phosphatase of Mr = 80,000 from rabbit muscle catalyzes the dephosphorylation of skeletal muscle proteins that are phosphorylated by either phosphorylase kinase or cAMP-dependent protein kinase. Phosphorylase or glycogen synthase labeled by phosphorylase kinase at seryl residues 14 or 7, respectively, are both dephosphorylated by the phosphatase. Phosphorylase a and glycogen synthase compete with one another for the phosphatase. The phosphatase discriminates between different sites labeled by the cAMP-dependent protein kinase: glycogen synthase phosphorylated either to 1.0 or 1.8 mol phosphate/mol, or phosphorylase kinase phosphorylated on its β-subunit serve as substrates for the phosphatase, but the phosphorylase kinase α-subunit, the phosphorylated phosphatase inhibitor 1, or casein do not. Histone fraction IIA, phosphorylated by the catalytic subunit, was a poor substrate even at a concentration of 100 μm. Phosphorylation of the α-subunit of phosphorylase kinase had no influence on the kinetics of dephosphorylation of the β-subunit. Thus, the Mr = 80,000 phosphatase meets the functional definition of a protein phosphatase 1 [Cohen, P. (1978) Curr. Top. Cell. Regul.14, 117–196]. Furthermore, from a comparison of the known phosphorylated sites of these proteins, it appears that the phosphatase discriminates between different sites present in the phosphoproteins tested on the basis of the Km values for the reactions. It displays a preferential activity toward proteins with a primary structure wherein basic residues are two positions amino-terminal from the phosphoserine, AgrLysX-YSer(P) or LysArgX-YSer(P), rather and one residue away, ArgArgX-Ser(P).  相似文献   

9.
A series of N-acetyl-l-phenylalanyl peptides of general formula Ac-Phe-(Gly)n-NH2 (n = 0–2) has been synthesized to study the effect of leaving group chain length on the efficiency of chymotrypsin Aα amidase and peptidase activities. The effect upon catalysis of hydrophobic side chains on the leaving group was investigated using similar substrates with one of the glycine residues selectively substituted by an alanine residue as in AcPheAlaNH2, AcPheAlaGlyNH2, and AcPheGlyAlaNH2. Values of kcat and Km have been obtained from kinetic measurements at pH 8.00 and 25 °C. The results are shown to be consistent with binding schemes postulated from published model building studies. The catalytic reactions were studied over a range of temperature (15–35 °C) and in each case the Arrhenius law was obeyed. It was thus possible to obtain meaningful values for the thermodynamic functions of activation for the acylation step of the catalytic reaction. The results are shown to confirm the findings of postulated binding schemes but indicate that conclusions drawn from kinetic measurements at a single temperature may sometimes be misleading.  相似文献   

10.
The kinetic parameters for the hydrolyses of different l-α-amino acid-β-naphthylamides by Bacillus subtilis aminopeptidase have been measured for the native enzyme and for the enzyme activated in 5 mm Co(NO3)2. In most cases Co2+ activation decreased Km(app) values and increased kcat values, in other cases km(app) and kcat values were increased; for the remainder of the substrates tested km(app) values and kcat values were decreased. In all cases tested the ratios of (kcatKm(app))CO2+/(kcatKm(app)nativ) were increased (2- to 108-fold). For the native enzyme the order of specificity toward the l-amino acid-β-naphthylamides was Arg > Met > Trp > Lys > Leu and for the Co2+ activated enzyme the order of specificity was Lys > Arg > Met > Trp > Leu. The native enzyme hydrolyzed Pro-β-naphthylamide, but not α-Glu-β-naphthylamide; Co2+ activation of the enzyme affected an appreciable rate of hydrolysis of the latter substrate.  相似文献   

11.
Lysosomal carboxypeptidase B has been purified from rabbit lung acetone powder by acid precipitation and ammonium sulfate fractionation followed by further purification on Sephadex G-100, DEAE-Sephadex, Organomercurial-Sepharose, preparative isoelectric focusing, Sephadex G-75, and carboxymethyl-Sephadex. This procedure resulted in a homogeneous preparation as determined by polyacrylamide gel electrophoresis at pH 4.5, 8.3 and with sodium dodecyl sulfate. This enzyme has a molecular weight of 52,000, is composed of two subunits of approximately equivalent molecular weight, and is a glycoprotein with a carbohydrate content estimated to be 10% by weight. The amino acid composition is also reported. The enzyme is active on two synthetic substrates, α-N-benyoyl-l-arginineamide and hippuryl-l-arginine. With these two substrates, respectively, lysosomal carboxypeptidase B has pH optima of 5.7 and 5.0, temperature optima of 40 and 50 °C, and Km values of 10 and 16 mm. With each substrate, the enzyme requires the presence of a reducing agent for maximal activity and is inhibited to the same extent with several inhibitors. The most potent inhibitors were leupeptin and antipain at low concentrations (1 μm). Iodoacetate and Ac-(Ala)3-Ala-chloro-methyl ketone also inhibited at higher concentrations (10 μm). However, compounds such as leucyl-chloromethyl ketone, bestatin, pepstatin, phenylmethylsulfonyl fluoride, soybean trypsin inhibitor, and α-1-antitrypsin did not inhibit. When tested with peptides as substrates, this proteinase exhibited strong carboxypeptidase activity on the tetrapeptide, ThrProArgLys, and on angiotensin I, AspArgValTyrIle HisProPheHisLeu, liberating Lys, and Leu, respectively. Substance P (containing 11 amino acids plus a C-terminal amide group) was virtually inactive as a substrate for this enzyme. However, with oxidized insulin B chain as substrate, lysosomal carboxypeptidase B exhibited significant carboxypeptidase and endopeptidase activities.  相似文献   

12.
The sequence of 96 amino acid residues from the COOH-terminus of the active subunit of cholera toxin, A1, has been determined as PheAsnValAsnAspVal LeuGlyAlaTyrAlaProHisProAsxGluGlu GluValSerAlaLeuGlyGly IleProTyrSerGluIleTyrGlyTrpTyrArg ValHisPheGlyValLeuAsp GluGluLeuHisArgGlyTyrArgAspArgTyr TyrSerAsnLeuAspIleAla ProAlaAlaAspGlyTyrGlyLeuAlaGlyPhe ProProGluHisArgAlaTrp ArgGluGluProTrpIleHisHisAlaPro ProGlyCysGlyAsnAlaProArg(OH). This is the largest fragment obtained by BrCN cleavage of the subunit A1 (Mr 23,000), and has previously been indicated to contain the active site for the adenylate cyclase-stimulating activity. Unequivocal identification of the COOH-terminal structure was achieved by separation and analysis of the terminal peptide after the specific chemical cleavage at the only cysteine residue in A1 polypeptide. The site of self ADP-ribosylation in the A1 subunit [C. Y. Lai, Q.-C. Xia, and P. T. Salotra (1983) Biochem. Biophys. Res. Commun.116, 341–348] has now been identified as Arg-50 of this peptide, 46 residues removed from the COOH-terminus. The cysteine that forms disulfide bridge to A2 subunit in the holotoxin is at position 91.  相似文献   

13.
Digestion of the native pig kidney fructose 1,6-bisphosphatase tetramer with subtilisin cleaves each of the 35,000-molecular-weight subunits to yield two major fragments: the S-subunit (Mr ca. 29,000), and the S-peptide (Mr 6,500). The following amino acid sequence has been determined for the S peptide: AcThrAspGlnAlaAlaPheAspThrAsnIle Val ThrLeuThrArgPheValMetGluGlnGlyArgLysAla ArgGlyThrGlyGlu MetThrGlnLeuLeuAsnSerLeuCysThrAlaValLys AlaIleSerThrAla z.sbnd;ValArgLysAlaGlyIleAlaHisLeuTyrGlyIleAla. Comparison of this sequence with that of the NH2-terminal 60 residues of the enzyme from rabbit liver (El-Dorry et al., 1977, Arch. Biochem. Biophys.182, 763) reveals strong homology with 52 identical positions and absolute identity in sequence from residues 26 to 60.Although subtilisin cleavage of fructose 1,6-bisphosphatase results in diminished sensitivity of the enzyme to AMP inhibition, we have found no AMP inhibition-related amino acid residues in the sequenced S-peptide. The loss of AMP sensitivity that occurs upon pyridoxal-P modification of the enzyme does not result in the modification of lysyl residues in the S-peptide. Neither photoaffinity labeling of fructose 1,6-bisphosphatase with 8-azido-AMP nor modification of the cysteinyl residue proximal to the AMP allosteric site resulted in the modification of residues located in the NH2-terminal 60-amino acid peptide.  相似文献   

14.
ADP-ribosyl cyclase and NAD+ glycohydrolase (CD38, E.C.3.2.2.5) efficiently catalyze the exchange of the nicotinamidyl moiety of NAD+, nicotinamide adenine dinucleotide phosphate (NADP+) or nicotinamide mononucleotide (NMN+) with an alternative base. 4′-Pyridinyl drugs (amrinone, milrinone, dismerinone and pinacidil) were efficient alternative substrates (kcat/KM = 0.9-10 μM−1 s−1) in the exchange reaction with ADP-ribosyl cyclase. When CD38 was used as a catalyst the kcat/KM values for the exchange reaction were reduced two or more orders of magnitude (0.015-0.15 μM−1 s−1). The products of this reaction were novel dinucleotides. The values of the equilibrium constants for dinucleotide formation were determined for several drugs. These enzymes also efficiently catalyze the formation of novel mononucleotides in an exchange reaction with NMN+, kcat/KM = 0.05-0.4 μM−1 s−1. The kcat/KM values for the exchange reaction with NMN+ were generally similar (0.04-0.12 μM−1 s−1) with CD38 and ADP-ribosyl cyclase as catalysts. Several novel heterocyclic alternative substrates were identified as 2-isoquinolines, 1,6-naphthyridines and tricyclic bases. The kcat/KM values for the exchange reaction with these substrates varied over five orders of magnitude and approached the limit of diffusion with 1,6-naphthyridines. The exchange reaction could be used to synthesize novel mononucleotides or to identify novel reversible inhibitors of CD38.  相似文献   

15.
Hong SH  Lim YR  Kim YS  Oh DK 《Biochimie》2012,94(9):1926-1934
A recombinant thermostable l-fucose isomerase from Dictyoglomus turgidum was purified with a specific activity of 93 U/mg by heat treatment and His-trap affinity chromatography. The native enzyme existed as a 410 kDa hexamer. The maximum activity for l-fucose isomerization was observed at pH 7.0 and 80 °C with a half-life of 5 h in the presence of 1 mM Mn2+ that was present one molecular per monomer. The isomerization activity of the enzyme with aldose substrates was highest for l-fucose (with a kcat of 15,500 min−1 and a Km of 72 mM), followed by d-arabinose, d-altrose, and l-galactose. The 15 putative active-site residues within 5 Å of the substrate l-fucose in the homology model were individually replaced with other amino acids. The analysis of metal-binding capacities of these alanine-substituted variants revealed that Glu349, Asp373, and His539 were metal-binding residues, and His539 was the most influential residue for metal binding. The activities of all variants at 349 and 373 positions except for a dramatically decreased kcat of D373A were completely abolished, suggesting that Glu349 and Asp373 were catalytic residues. Alanine substitutions at Val131, Met197, Ile199, Gln314, Ser405, Tyr451, and Asn538 resulted in substantial increases in Km, suggesting that these amino acids are substrate-binding residues. Alanine substitutions at Arg30, Trp102, Asn404, Phe452, and Trp510 resulted in decreases in kcat, but had little effect on Km.  相似文献   

16.
Summary Enteropeptidase (enterokinase EC 3.4.21.9), catalyzing trypsinogen activation, exhibits unique properties for high efficiency hydrolysis of the polypeptide chain after the N-terminal tetraaspartyl-lysyl sequence. This makes it a convenient tool for the processing of fusion proteins containing this sequence. We found the enteropeptidase-catalysing degradation of some bioactive peptides: cattle hemoglobin beta-chain fragments Hb (2–8) (LTAEEKA) and Hb (1–9) (MLTAEEKAA), human angiotensin II (DRVYIHPF) (AT). Model peptides with truncated linker WDDRG and WDDKG also were shown to be susceptible to enteropeptidase action. Kinetic parameters of enteropeptidase hydrolysis for these substrates were determined.K m values for all substrates with truncated linker (≈10−3 M) are an order of magnitude higher than corresponding values for typical enteropeptidase artificial peptide or fusion protein substrates with full enteropeptidase linker-DDDDK-(K m ≈10−4 M).k cat values for AT, Hb (2–8), WDDRG and WDDKG are ≈30–40 min−1. But one additional amino acid residue at both N-and C-terminus of Hb (2–8) results in a drastic increase of hydrolysis efficiency:k cat value for Hb (1–9) is 1510 min−1. Recent study demonstrates the possibility of undesirable cleavage of target peptides or proteins containing the above-mentioned truncated linker sequences; further, the ability of enteropeptidase to hydrolyse specifically several biologically active peptidesin vitro along with its unique natural substrate trypsinogen was demonstrated.  相似文献   

17.
Liu YL  Chiang YH  Liu GY  Hung HC 《PloS one》2011,6(6):e21314
Peptidylarginine deiminase 4 (PAD4) is a homodimeric enzyme that catalyzes Ca2+-dependent protein citrullination, which results in the conversion of arginine to citrulline. This paper demonstrates the functional role of dimerization in the regulation of PAD4 activity. To address this question, we created a series of dimer interface mutants of PAD4. The residues Arg8, Tyr237, Asp273, Glu281, Tyr435, Arg544 and Asp547, which are located at the dimer interface, were mutated to disturb the dimer organization of PAD4. Sedimentation velocity experiments were performed to investigate the changes in the quaternary structures and the dissociation constants (K d) between wild-type and mutant PAD4 monomers and dimers. The kinetic data indicated that disrupting the dimer interface of the enzyme decreases its enzymatic activity and calcium-binding cooperativity. The K d values of some PAD4 mutants were much higher than that of the wild-type (WT) protein (0.45 µM) and were concomitant with lower k cat values than that of WT (13.4 s−1). The K d values of the monomeric PAD4 mutants ranged from 16.8 to 45.6 µM, and the k cat values of the monomeric mutants ranged from 3.3 to 7.3 s−1. The k cat values of these interface mutants decreased as the K d values increased, which suggests that the dissociation of dimers to monomers considerably influences the activity of the enzyme. Although dissociation of the enzyme reduces the activity of the enzyme, monomeric PAD4 is still active but does not display cooperative calcium binding. The ionic interaction between Arg8 and Asp547 and the Tyr435-mediated hydrophobic interaction are determinants of PAD4 dimer formation.  相似文献   

18.
Comparative substrate analysis of full-length bovine enteropeptidase and trypsin, bovine and human enteropeptidase light chains was performed using model N-terminal dodecapeptides corresponding to wild-type human trypsinogen and pancreatitis-associated mutant trypsinogens K23R and D22G. The substitution of Lys residue by Arg at P1 leads to 2-fold increase in the efficiency of enteropeptidase hydrolysis; the absence of the negatively charged residue at P2 reduces the efficiency of such hydrolysis by two orders of magnitude. The difference in efficiency of peptide chain hydrolysis after Lys/Arg residues by enteropeptidase compared to trypsin is equal to the difference in hydrolysis by serine proteases of different primary specificity of their specific substrates.  相似文献   

19.
PheMetArgPheNH2 (FMRFamide), injected at < μmol/kg intravenously in the anesthetized rat, produces sharp elevations of blood pressure and changes in respiration. The effects were dependent on the carboxyterminal ArgPhe (RF) configuration and were stereospecific for these two amino acids. A related peptide with RF carboxyterminus, γ1-melanotropic stimulating hormone, also had potent blood pressure stimulating activity. The mechanisms underlying the pressor effect of FMRFamide have not yet been established but this pressor action was not significantly attenuated by standard pharmacologic antagonists or prevented by removal of the adrenal or pituitary gland.  相似文献   

20.
The rates of deuterium exchange reactions of malondialdehyde (MDA) and deuterated malondialdehyde (MDAd) have been studied as a function of acidity and the content of dimethyl sulfoxide (DMSO) in binary mixtures with D2O . MDA incorporates deuterium from D2O solutions in a first-order reaction with a rate constant (kobs) that depends on the acid concentration. From this dependence, a catalytic constant, kcat, can be derived (kcatMDA = 2.25 × 105M?s?1). Similar kinetic behavior was found for MDAd in H2O solutions, and in this case, kcatMDA = 1.56 × 105M?1s?1. Results from reactions of MDA and MDAd in identical H2OD2O mixtures show that primary and secondary isotope effects are small (kH/kD = 1.13) and that solvent isotope effects cause most of the differences found between reactions in D2O and H2O. Reactions in binary DMSOd6D2O mixtures show a six-fold rate increase as the proportion of DMSOd6 increases from 50% to 90%. These results also illustrate the relatively high reactivity of MDA at pH values well above its pKa and the importance of medium composition on its reaction rate.  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