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 797 毫秒
1.
Egg whites of three species of tortoise and turtle have been compared by gel chromatography for inhibitory activity against proteases. The egg white of Geomda trijuga trijuga Schariggar contains trypsin/subtilisin inhibitor while the egg white of Caretta caretta Linn. contains both trypsin and chymotrypsin inhibitors. No protease inhibitory activity has been detected in the egg white of Trionyx gangeticus Cuvier. An acidic trypsin/subtilisin inhibitor has been purified to homogeneity from the egg white of tortoise (G. trijuga trijuga). It is a single polypeptide chain of 100 amino acid residues, having a molecular weight of 11 700. It contains six disulphide bonds and is devoid of methionine and carbohydrate moiety. Its isoelectric point is at pH 5.95 and is stable at 100°C for 4 h at neutral pH. The inhibitor inhibits both trypsin and subtilisin by forming enzyme-inhibitor complexes at a molar ratio close to unity. Their dissociation contants are 7.2·10?9 M for bovine trypsin adn 5.5·10?7 M for subtilisin. Chemical modification of amino groups with trinitrobenzene sulfonate has reduced its inhibitory activities against both trypsin and subtilisin, but the loss of its trypsin inhibitory activity is faster than of its subtilisin inhibitory activity. It has independent binding sites for inhibition of trypsin and subtilisin.  相似文献   

2.
A serine protease inhibitor with a molecular mass of 6106±2Da (designated as InhVJ) was isolated from the tropical anemone Radianthus macrodactylus by a combination of liquid chromatography methods. The molecule of InhVJ consists of 57 amino acid residues, has three disulfide bonds, and contains no Met or Trp residues. The N-terminal amino acid sequence of the inhibitor (19 aa residues) was established. It was shown that this fragment has a high degree of homology with the N-terminal amino acid sequences of serine protease inhibitors from other anemone species, reptiles, and mammals. The spatial organization of the inhibitor at the levels of tertiary and secondary structures was studied by the methods of UV and CD spectroscopy. The specific and molar absorption coefficients of InhVJ were determined. The percentage of canonical secondary structure elements in the polypeptide was calculated. The inhibitor has a highly ordered tertiary structure and belongs to mixed α/β-or α + β polypeptides. It was established that InhVJ is highly specific toward trypsin (K i 2.49 × 10?9 M) and α-chymotrypsin (K i 2.17 × 10?8 M) and does not inhibit other proteases, such as thrombin, kallikrein, and papain. The inhibitor InhVJ was assigned to the family of the Kunitz inhibitor according to its physicochemical properties.  相似文献   

3.
The interaction of the inhibitor VJ (InhVJ), isolated from sea anemone R. macrodactylus, with different proteases was investigated using the method of biosensor analysis. The following enzymes were tested: serine proteases (trypsin, α-chymotrypsin, plasmin, thrombin, kallikrein), cysteina protease (papain) and aspartic protease (pepsin). In the rage of the concentrations studied (10–400 nM) inhibitor VJ interacted only with trypsin and α-chymotrypsin. The intermolecular complexes formation between inhibitor VJ and each of these enzymes was characterized by the following kinetic and thermodynamics parameters: KD = 7.38 × 10?8 M and 9.93 × 10?7 M for pairs InhVJ/trypsin and InhVJ/α-chymotrypsin, respectively.  相似文献   

4.
A proteinase inhibitor resembling Bowman-Birk family inhibitors has been purified from the seeds of cultivar HA-3 of Dolichos lablab perpureus L. The protein was apparently homogeneous as judged by SDS–PAGE, PAGE, IEF, and immunodiffusion. The inhibitor had 12 mole% 1/2-cystine and a few aromatic amino acids, and lacks tryptophan. Field bean proteinase inhibitor (FBPI) exhibited a pI of 4.3 and an M r of 18,500 Da. CD spectral studies showed random coiled secondary structure. Conformational changes were detected in the FBPI–trypsin/chymotrypsin complexes by difference spectral studies. Apparent K a values of complexes of inhibitor with trypsin and chymotrypsin were 2.1 × 107 M?1 and 3.1 × 107 M?1, respectively. The binary and ternary complexes of FBPI with trypsin and chymotrypsin have been isolated indicating 1:1 stoichiometry with independent sites for cognate enzymes. Amino acid modification studies showed lysine and tyrosine at the reactive sites of FBPI for trypsin and chymotrypsin, respectively.  相似文献   

5.
A trypsin inhibitor, termed ovostatin, has been purified approximately 265-fold with 82% yield, from unfertilized eggs of the sea urchin Strongylocentrotus intermedius, using trypsin coupled Sepharose 4B as an affinity column for chromatography. The isolated ovostatin is homogeneous in sodium dodecyl sulfate/polyacrylamide gel electrophoresis, the estimated molecular weight being 20K–21.5K. Ovostatin inhibits preferentially trypsin-like endogenous protease purified from the eggs of the same species and bovine pancreatic trypsin and also bovine pancreatic chymotrypsin. Values of IC50 (amount causing 50% inhibition of enzymes) for trypsin-like protease purified from eggs of the same species, bovine pancreatic trypsin, and bovine pancreatic chymotrypsin, are 0.91 ± 0.13 μg/ml (4.55 ± 0.65 × 10?8 M), 3.0 ± 0.28 μg/ml (1.5 ± 0.14 × 10?7 M), and 4.8 ± 0.2 μg/ml (2.4 ± 0.1 × 10?7 M), respectively, in the experimental condition used. Kinetic studies indicate that ovostatin is a noncompetitive inhibitor of trypsin. The inhibitor is relatively heat labile. NaCl (0.025–0.01 M) enhances the inhibitor activity, whereas KCl is inhibitory. Ovostatin requires a low concentration of Ca2+ for activity. The activity is higher in unfertilized eggs than in fertilized eggs; total activity and specific activity in unfertilized eggs is about 1.67-fold and 1.85-fold higher than those in fertilized eggs, respectively. We believe that ovostatin may regulate the function of the cortical granule protease and other trypsin-like proteases that are activated in sea urchin eggs during fertilization.  相似文献   

6.
Abstract: Pridefine (AHR-1118) is a pyrrolidine derivative with clinically established antidepressant efficacy. Previous work from this laboratory indicates that pridefine is a reuptake blocker of catecholamines and serotonin with weak releasing activity. This study characterized the mode of amine uptake inhibition by pridefine as noncompetitive. The uptake experiments were performed utilizing ouabain instead of zero-degree controls to differentiate between the passive and active components of uptake. Furthermore, the passive component was resolved into diffusion and binding of substrate. Correction was made for the effects of ouabain on binding. Kinetic constants determined from Lineweaver-Burk plots were: Km= 3 × 10?7 M for NE, Km= 9 × 10?8 M for DA, and Km= 3 × 10?8 M for 5-HT. Dixon analyses of uptake at various pridefine concentrations indicated noncompetitive inhibition with Ki= 2.5 × 10?6 M for NE uptake, Ki= 2.0 × 10?6 M for DA uptake, and Ki= 1 × 10?5 M for 5-HT uptake. These constants compare well with IC50 values for the same transmitters: NE, IC50= 2.4 × 10?6 M; DA, IC50= 2.8 × 10?6 M; 5-HT, IC50= 1.0 × 10?5 M. The in vitro results indicate that pridefine is relatively specific as a catecholamine uptake blocker. It differs from tricyclic antidepressants which are reportedly competitive inhibitors of monoamine uptake. The possible mechanisms by which pridefine acts as a noncompetitive inhibitor are discussed.  相似文献   

7.
Protease inhibitors control major biological protease activities to maintain physiological homeostasis. Marine bacteria isolated from oligotrophic conditions could be taxonomically distinct, metabolically unique, and offers a wide variety of biochemicals. In the present investigation, marine sediments were screened for the potential bacteria that can produce trypsin inhibitors. A moderate halotolerant novel marine bacterial strain of Oceanimonas sp. BPMS22 was isolated, identified, and characterized. The effect of various process parameters like salt concentration, temperature, and pH was studied on the growth of the bacteria and production of trypsin inhibitor. Further, the trypsin inhibitor was purified to near homogeneity using anion exchange, size exclusion, and affinity chromatography. The purified trypsin inhibitor was found to competitively inhibit trypsin activity with an inhibition coefficient, Ki, of 3.44?±?0.13 μM and second-order association rate constant, kass, of 1.08?×?103 M?1 S?1. The proteinaceous trypsin inhibitor had a molecular weight of approximately 30 kDa. The purified trypsin inhibitor showed anticoagulant activity on the human blood samples.  相似文献   

8.
Comparative data on the properties of four thiol proteinase inhibitors, and of four serine proteinase inhibitors (two subtilisin and two trypsin inhibitors) isolated from seeds of Vigna are presented. They were similar in their molecular weights (5000–15,000) and dissociation constants (10?8–10?9m). The range of isoelectric points of the thiol proteinase inhibitors was 6.5 to 10.6, and of the serine proteinase inhibitors was 5.0 to 5.9. The amino acid compositions of one papain isoinhibitor, one of subtilisin, and one of trypsin are presented. Papain inhibitor A1 and subtilisin inhibitor 2a were low in cystine. All of the inhibitors were stable upon heating to 80 °C for 5 min at low pH. The subtilisin inhibitor did not bind to catalytically inactive subtilisin derivatives, whereas the papain inhibitor was stoichiometrically bound to the Hg or thioacetamide derivatives of papain. Incubation of the subtilisin inhibitor with catalytic amounts of subtilisin led to the formation of a modified form with the same inhibitor activity as the native inhibitor but with a different electrophoretic mobility. There was no indication of a similar modification of the papain inhibitor by papain. Separate sites are present on the trypsin-chymotrypsin inhibitors for trypsin and chymotrypsin. The papain inhibitors have the same binding sites for papain and ficin.  相似文献   

9.
A novel of the potato inhibitor I family of serine proteinase inactivating proteins has been isolated from seeds of grain amaranth (Amaranthus caudatus L.) and characterized. The mature form of the amaranth trypsin/subtilisin inhibitor (ATSI) with pI ≈ 8.3 and molecular mass 7887 Da contains 69 amino acids in a sequence showing 33–51% identity with members of the inhibitor I family from other plant families. A minor form with pI ≈ 7.8 and same inhibitory properties lacked the N-terminal dipeptide Ala-Arg. In accordance with the reactive-site bond Lys45-Asp46, which was identified by specific cleavage on a subtilisin column, ATSI is a potent inhibitor of trypsin (Ki ≈ 0.34 nM) and more weakly of plasmin (Ki ≈ 38 nM) and Factor XIIa (Ki ≈ 440 nM). However, ATSI also inactivates chymotrypsin (Ki ≈ 0.41 nM), cathepsin G (Ki ≈ 122 nM) and several alkaline microbial proteinases, including subtilisin NOVO (Ki ≈ 0.37 nM). Interestingly, ATSI contains a Trp residue instead of the highly conserved Arg in position 53 (P′B), which is assumed to play a central role in stabilization of the active-site loop during complex formation. ATSI was immediately inactivated by peptsin and hardly represents an antinutritional component in foods or feeds.  相似文献   

10.
Adenylyl imidodiphosphate (AMP-PNP), and analog of adenosine triphosphate (ATP), is a potent competitive inhibitor of mitochondrial ATPase activity. It inhibits both the soluble oligomycin-insensitive ATPase (Ki = 9.2 × 10?7 M) and the bound oligomycin-sensitive APTase (Ki = 1.3 × 10?6 M). ATPase activity of inside-out submitochondrial preparations are more sensitive to AMP-PNP in the presence of an uncoupler (Ki = 2.0 × 10?7 M). Mitochondrial ATP-dependent reactions (reversed electron transfer and potassium uptake) do not proceed if ATP is replaced with AMP-PNP; however, the analog does affect these systems. Oxidative phosphorylation of whole mitochondria and submitochondrial preparations were unaffected by AMP-PNP.  相似文献   

11.
A protein that inhibited the proteolytic activity of trypsin was isolated from amaranth leaves (Amaranthus cruentus) by affinity chromatography on trypsin-Sepharose. The inhibition was noncompetitive (withp-nitroanilide-N-α-benzoyl-DL-arginine as substrate) and had aK i, of 1.87 × 10−7 M. The protein caused a weaker inhibitory effect on chymotrypsin, had no effect on subtilisin, displayed a molecular weight of 8 kDa, and contained no cysteine residues.  相似文献   

12.
Following determination of trypsin inhibitory activity, a serine protease inhibitor was purified and characterized from frog Duttaphrynus melanostictus serum. It was identified as serum albumin, with molecular weight of 67 kDa (DmA-serum). Different from bovine serum albumin, DmA-serum potently inhibited trypsin with similar K i values around 1.6 × 10−7 M. No inhibitory effect on thrombin, chymotrypsin, elastase and subtilisin was observed under the assay conditions. The N-terminal amino acid is EAEPHSRI. Subsequently, a protein with same N-terminal amino acid was purified from skin, termed as DmA-skin. However, DmA-skin is distinct from DmA-serum by binding of a haem b (0.5 mol/mol protein), and with low trypsin inhibitory activity. Frog albumin is distributed in frog skin and exhibited trypsin inhibitory activity, suggesting that it plays important roles in skin physiological functions, like water economy, metabolite exchange and osmoregulation, etc.  相似文献   

13.
Z-Ala-Pro-Phe-glyoxal (where Z is benzyloxycarbonyl) has been shown to be a competitive inhibitor of subtilisin with a Ki=2.3±0.2 μM at pH 7.0 and 25 °C. Using Z-Ala-Pro-[2-13C]Phe-glyoxal we have detected a signal at 107.3 ppm by 13C NMR, which we assign to the tetrahedral adduct formed between the hydroxy group of serine-195 and the 13C-enriched keto-carbon of the inhibitor. The chemical shift of this signal is pH independent from pH 4.2 to 7.0 and we conclude that the oxyanion pKa<3. This is the first observation of oxyanion formation in a reversible subtilisin–inhibitor complex. The inhibitor is bound as a hemiketal which is in slow exchange with the free inhibitor. Inhibitor binding depends on a pKa of ~6.5 in the free enzyme and on a pKa<3.0 when the inhibitor is bound to subtilisin. Protonation of the oxyanion promotes the disassociation of the inhibitor. We show that oxyanion formation cannot be rate limiting during catalysis and that subtilisin stabilises the oxyanion by at least 45.1 kJ mol?1. We conclude that if the energy required for oxyanion stabilisation is utilised as binding energy in drug design it should make a significant contribution to inhibitor potency.  相似文献   

14.
Some properties of a preparation of an enzyme, lunularic acid decarboxylase, from the liverwort Conocephalum conicum are described. The enzyme is normally bound and could be solubilized with Triton X-100; at least some of the bound decarboxylase activity appears to be associated with chloroplasts. For lunularic acid the enzyme has Km 8.7 × 10?5 M (pH 7.8 and 30°). Some substrate analogues have been tested but no other substrate was found. Pinosylvic acid is a competitive inhibitor for the enzyme, Ki 1.2 × 10?4 M (pH 7.8 and 30°). No product inhibition was observed. Lunularic acid decarboxylase activity has also been observed with a cell-free system from Lunularia cruciata.  相似文献   

15.
A trypsin inhibitor was isolated from grains of two row barley (cv. Proctor). The purified protein was identical with the corresponding inhibitor of a six row barley (cv. Pirkka); both proteins showed, a Pi of 7.4. The N-terminal amino acid was phenylalanine and an arginine residue was involved in the active site. Effects of substrate concentration showed that the inhibition was noncompetitive with a Ki of about 0.9 × 10?7M. An enzyme-inhibitor complex was demonstrated by disc electrophoresis.  相似文献   

16.
New trypsin inhibitors Z-Lys-COCHO and Z-Lys-H have been synthesised. Ki values for Z-Lys-COCHO, Z-Lys-COOH, Z-Lys-H and Z-Arg-COOH have been determined. The glyoxal group (–COCHO) of Z-Lys-COCHO increases binding ~300 fold compared to Z-Lys-H. The α-carboxylate of Z-Lys-COOH has no significant effect on inhibitor binding. Z-Arg-COOH is shown to bind ~2 times more tightly than Z-Lys-COOH. Both Z-Lys-13COCHO and Z-Lys-CO13CHO have been synthesized. Using Z-Lys-13COCHO we have observed a signal at 107.4 ppm by 13C NMR which is assigned to a terahedral adduct formed between the hydroxyl group of the catalytic serine residue and the 13C-enriched keto-carbon of the inhibitor glyoxal group. Z-Lys-CO13CHO has been used to show that in this tetrahedral adduct the glyoxal aldehyde carbon is not hydrated and has a chemical shift of 205.3 ppm. Hemiketal stabilization is similar for trypsin, chymotrypsin and subtilisin Carlsberg. For trypsin hemiketal formation is optimal at pH 7.2 but decreases at pHs 5.0 and 10.3. The effective molarity of the active site serine hydroxyl group of trypsin is shown to be 25300 M. At pH 10.3 the free glyoxal inhibitor rapidly (t1/2=0.15 h) forms a Schiff base while at pH 7 Schiff base formation is much slower (t1/2=23 h). Subsequently a free enol species is formed which breaks down to form an alcohol product. These reactions are prevented in the presence of trypsin and when the inhibitor is bound to trypsin it undergoes an internal Cannizzaro reaction via a C2 to C1 alkyl shift producing an α-hydroxycarboxylic acid.  相似文献   

17.
The binding mechanism of Streptomyces subtilisin inhibitor and subtilisin BPN′ was studied kinetically with the stopped-flow method by monitoring the protein fluorescence increase due to complex formation. In the lower concentration range of proteins, the reaction followed the second-order kinetics. The pH dependence of the apparent second-order rate constant, kon, suggested the involvement of the two ionizable groups of pKa of 7.8 and 10 in the binding. The activation parameters were calculated from the temperature dependence of the apparent second-order rate constants. The value of the apparent activation energy (EA = 39.7 kJ · mol?1, 9.50 kcal · mol?1) and insensitivity of kon to the viscosity of the medium suggest that the binding is not a simple diffusion-controlled bimolecular association. Further studies with a much broader range of protein concentrations have revealed that the reaction tends to approach first-order kinetics as the inhibitor concentration increases. The binding reaction is, therefore, reconcilable with a two-step mechanism, in which a fast bimolecular association is followed by a slow unimolecular isomerization step; the dissociation constant of the first step, KL, is estimated to be 1.2 × 10?4m and the rate constant of the second step, k+2, to be 770 s?1. It was also found that the increase of tryptophan fluorescence due to the complex formation occurs solely in the rate-determining unimolecular process.  相似文献   

18.
The interaction of alkylguanidines and decahydrohistrionicotoxin with the membrane-bound and solubilized muscarinic acetylcholine receptor (mAcChR) from porcine atria was described. Alkylguanidines with alkyl chain lengths from one to ten carbons displaced l-[3H]quinuclidinyl benzilate (l-[3H]QNB) competitively from a single class of sites for the membrane-bound mAcChR. From a plot of ?ln Ki versus alkyl carbon chain number, a value of ?(473 ± 30) cal/mol was estimated as the energetic contribution per methylene group to the total binding energy. The binding of alkylguanidines to the digitonin/cholate solubilized mAcChR was complex in nature resulting in titration curves that did not obey the law of mass action for simple competitive inhibition at higher alkyl carbon numbers and a sigmoidal plot of ?ln Ki versus carbon number. Decahydrohistrionicotoxin bound in a competitive manner versus l-[3H]QNB to both the membrane-bound (Ki = (6.9 ± 1.4) × 10?6 M) and the solubilized (Ki = (1.5 ± 0.3) × 10?5 M) preparations.  相似文献   

19.
Streptomyces lividans 1326 carries inducible mercury resistance genes on the chromosome, which are arranged in two divergently transcribed operons. Expression of the genes is negatively regulated by the repressor MerR, which binds in the intercistronic region between the two operons. The merR gene was expressed in E. coli using a T7 RNA polymerase/promoter expression system, and MerR was purified to around 95% homogeneity by ammonium sulfate precipitation, gel filtration and affinity chromatography. Gel filtration showed that the native MerR is a dimer with a molecular mass of 31?kDa. Two DNA binding sites were identified in the intercistronic mer promoter region by footprinting experiments. No evidence for cooperativity in the binding of MerR to the adjacent operator sequences was observed in gel mobility shift assays. The dissociation constants (KD) for binding of MerR were: binding site I, 8.5?×?10?9?M; binding site II, 1.2?×?10?8?M; and for the complete promoter/operator region 1?×?10?8?M. The half-life of the MerR-DNA complex was 19.4?min and 18.8?min for binding site I and binding site II, respectively. The KD value for binding of mercury(II)chloride to MerR, again determined by mobility shift assay, was 1.1?×?10?7?M.  相似文献   

20.
Phosphoramidon, N-(α-l-rhamnopyranosyloxyhydroxyphosphinyl)-l-leucyl-l-tryptophan, and its analog, N-phosphoryl-l-leucyl-l-tryptophan, inhibited thermolysin in a competitive manner and Ki values were calculated to be 2.8 × 10?8 and 2.0 × 10?9m, respectively. The l-rhamnose moiety in phosphoramidon was suggested to be not involved in inhibition of thermolysin. A phosphoramidon analog containing histidine instead of tryptophan showed weaker inhibition. Spectrophotometric titration based on difference ultraviolet absorption spectra of the enzyme-inhibitor complex showed equimolar binding of the inhibitor to the enzyme.  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