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.
Calcium-dependent proteases: an enzyme system active at cellular membranes?   总被引:8,自引:0,他引:8  
Proteases having a neutral pH optimum and an absolute requirement for calcium ion are found in virtually all mammalian cells. Association of calcium-dependent proteases and a specific inhibitor protein with biological membranes seems to be an important regulatory feature of this proteolytic system, and it is likely that membranes are preferred sites for calcium-dependent protease action. Several recent hypotheses for the physiological function of calcium-dependent proteolysis are consistent with a membrane-associated protease action. Calcium-dependent proteases may participate in cell membrane fusion: the proteolysis of membrane proteins, which is required for the efficient fusion of erythrocytes, may be catalyzed by these enzymes. There is also evidence for the involvement of calcium-dependent proteolysis in postsynaptic membrane remodeling in the hippocampus after long-term potentiation. Although the relationship of the proteolysis to synaptic function is not known, it could have important physiological or pathophysiological consequences. Finally, it has recently been suggested that calcium-dependent proteolysis may be a physiologically significant mechanism for activating membrane-associated protein kinase C after exposure of some cell types to phorbol esters or other mitogens. Further pursuit of these hypotheses may reveal a novel role for intracellular calcium-regulated proteolysis in membrane-associated cell functions.  相似文献   

2.
Four intracellular proteases partially purified from liver preferentially degraded the oxidatively modified (catalytically inactive) form of glutamine synthetase. One of the proteases was cathepsin D which is of lysosomal origin; the other three proteases were present in the cytosol. Two of these were calcium-dependent proteases with different calcium requirements. The low-calcium-requiring type (calpain I) accounted for most of the calcium-dependent activity of both mouse and rat liver. The calcium-independent cytosolic protease, referred to as the alkaline protease, has a molecular weight of 300,000 determined by gel filtration. Native glutamine synthetase was not significantly degraded by the cytosolic proteases at physiological pH, but oxidative modification of the enzyme caused a dramatic increase in its susceptibility to attack by these proteases. In contrast, trypsin and papain did degrade the native enzyme and the degradation of modified glutamine synthetase was only 2- to 4-fold more rapid. Adenylylation of glutamine synthetase had little effect on its susceptibility to proteolysis. Although major structural modifications such as dissociation, relaxation, and denaturation also increased the rate of degradation, the oxidative modification is a specific type of covalent modification which could occur in vivo. Oxidative modification can be catalyzed by a variety of mixed function oxidase systems present within cells and causes inactivation of a number of enzymes. Moreover, the presence of cytosolic proteases which recognize the oxidized form of glutamine synthetase suggests that oxidative modification may be involved in intracellular protein turnover.  相似文献   

3.
Three thermostable proteases, designated S, N, and B, are extracellular enzymes produced by Bacillus stearothermophilus strain TLS33. They were purified by lysine affinity chromatography, strong anion exchange Q HyperD chromatography, and Ultrogel AcA44 gel filtration. The molecular masses of the enzymes determined by SDS-PAGE and zymography were approximately 36, 53, and 71 kDa, respectively. Thermostable protease S bound strongly to the lysine affinity column and could be purified by this single step. The optimum pH values of proteases S, N, and B were shown to be 8.5, 7.5, and 7.0, respectively. The maximum activities for the enzymes were at 70, 85, and 90 degrees C, respectively. Proteases S, N, and B at pH 7.0 in the presence of 5 mM CaCl(2) retained half their activities after 30 min at 72, 78, and 90 degrees C, respectively. All three thermostable proteases were strongly inhibited by the metal chelators EDTA and 1,10-phenanthroline, and the proteolytic activities were restored by addition of ZnCl(2). They can thus be classified as Zn(2+) metalloproteases. The cleavage specificities of proteases S, N, and B on a 30-residue synthetic peptide from pro-BPN' subtilisin were Tyr-Ile, Phe-Lys, and Gly-Phe, respectively.  相似文献   

4.
Protease which was found in the culture fluid of Pseudomonas sp. No. 548 was fractionated into four components with protease activity by a two step chromatography using DEAE-cellulose. Each protease was further purified by gel filtration on Sephadex G-100 and/or G-75. The protease of Ia was obtained in crystalline form and was shown to be homogeneous by analysis with electrophoresis, while the other three enzymes were also highly purified. The enzymatic properties of the proteases were investigated. All of the four enzymes were inactivated by ethylene diamine tetraacetate. Proteases Ia, Ib, and IIb were inactivated by diisopropylfluorophosphate. The optimum activity of protease Ia was shown to be at pH 10.0, and that of the other enzymes were at pH 7.0 to 8.0. The proteases of Ia, Ib, and IIb were stabilized by calcium ion. The effect of temperature, pH, and metal ions on the activity of the enzyme were also investigated.  相似文献   

5.
Extracts from white croaker skeletal muscle showed two alkaline proteases and a trypsin inhibitor when they were chromatographed in DEAE-Sephacel. The activity against azocasein was maximal at pH 8.5 and 9.1 for proteases I and II, respectively. Both enzymes showed optimum activity at 60° C. The molecular masses were found to be 132 kDa for protease 1,363 kDa for protease II, and 65 kDa for the inhibitor. Protease I showed the characteristics of a trypsin-like enzyme, and protease II those of a SH-enzyme. These proteins may play important roles in mechanisms of cellular proteolysis.  相似文献   

6.
In this study, we investigated the application of cellulase and protease purified from rumen bacteria as detergent additives. Cellulase and protease were purified from the rumen cellulytic bacteria Fibrobacter succinogenes S85, and Prevotella ruminicola 23, respectively. An inhibitor test indicated that the purified protease belongs to the category of serine proteases and metalloproteases. Both the enzymes were effective at a high temperature (50 degrees C) and neutral pH (pH 7-8), but the protease activity increased with the increase in temperature and pH. The purified protease was treated with ten types of surfactants/detergents; it was found to retain over 60% of its activity in the presence of anionic and nonionic detergents. The cellulose plus protease combination was still effective after treatment with Triton X-100 and Tween 80, but the residual activity was low after treatment with Tween 20 than that after treatment with other nonionic detergents. Washing tests indicated that enzyme addition produced no significant improvement in the removal of grass stains, but individual enzyme addition in surfactants/detergents, especially in nonionic detergents, could improve the washing performance of the detergents by improving its ability to remove blood stains. This suggested that the surfactant/detergent class, enzyme properties, and the mixing ratio of ingredients should be considered simultaneously to enhance the washing performance.  相似文献   

7.
The isolation and partial characterization of the acid proteases A1 and A2 (EC3.4.23.6) from Aspergillus oryzae grown on solid bran culture are described. The purified preparations were essentially homogeneous by several criteria including sedimentation analysis and polyacrylamide gel electrophoresis. The physiochemical properties of the proteases A1 and A2 were as follows (in the order: A1, A2): molecular weight: 63 000 & 32 000; sedimentation coefficient s20, w: 3.93 and 3.16 S; diffusion constant D20, w, 5.63 - 10(-7) and 8.61 - 10(-7) CM2/S, partial specific volume, v: 0.73 ml/g for both; nitrogen content: 16.30 and 13.42%; E1% 1 cm at 280 nm: 5.9 and 11.1. The two enzymes had the same pH optima in the acid pH range, and both activated bovine pancreatic trypsinogen. The enzymes were essentially of the same amino acid composition and immunologically cross-reacted with each other. The protease A2 contained little or no carbohydrate, whereas the protease A1 was glycoprotein, containing 49% carbohydrate comprising glucose, mannose, and galactose. These results suggest that the protein portion of acid protease A1 is the same as that of acid protease A2.  相似文献   

8.
Pseudomonas aeruginosa myovirus ϕKZ has a 270-kb genome within a T=27 icosahedral capsid that contains a large, unusual, and structurally well-defined protein cylindrical inner body (IB) spanning its interior. Proteolysis forms a pivotal stage in ϕKZ head and IB morphogenesis, with the protease gp175 cleaving at least 19 of 49 different head proteins, including the major capsid protein and five major structural IB proteins. Here we show that the purified mature form of gp175 is active and cleaves purified IB structural proteins gp93 and gp89. Expression vector synthesis and purification of the zymogen/precursor yielded an active, mature-length protease, showing independent C-terminal gp175 self-cleavage autoactivation. Mutation of either the predicted catalytic serine or histidine inactivated mature gp175, supporting its classification as a serine protease and representing the first such direct biochemical demonstration with purified protease and substrate proteins for any phage protease. These mutations also blocked self-cleavage of the precursor while allowing intermolecular gp175 processing. To confirm the cleavage specificity of gp175, we mutated three cleavage sites in gp93, which blocked proteolysis at these sites. The N-terminal propeptide of gp93 was shown to undergo more extensive proteolysis than previously identified. We found that proteolysis in gp93 progressed from the N to C terminus, while blocking cleavage sites slowed but did not eliminate downstream proteolysis. These findings were shown by informatics to be relevant to the head morphogenesis of numbers of other related IB-containing giant phages as well as to T4 and herpesviruses, which have homologous proteases.  相似文献   

9.
Macrophages actively internalize macromolecules into endosomal vesicles containing proteases. The plant toxin, ricin A chain delivered into this pathway by receptor-mediated endocytosis, was found to be exquisitely sensitive to cleavage by these proteases. Proteolytic fragments of ricin A chain were generated within cells as early as 2-3 min after internalization. Toxin proteolysis was initiated in early endosomal vesicles, and transport to lysosomes was not required. As endosomes transit the cell, their lumenal pH drops from neutral to acidic. Previous studies in macrophages had suggested that endosomal proteolysis is dependent on vesicle acidification. Isolated endosomal vesicles containing ricin A chain catalyzed the cleavage of this protein in vitro; however, proteolysis was observed at both neutral and acidic pH. Experiments using isolated endosomes demonstrated that both cysteine and aspartyl proteases were responsible for the cleavage of ricin A chain. The cysteine protease, cathepsin B, catalyzed toxin proteolysis in endosomes between pH 4.5 and 7.0 while aspartyl protease activity was maximal below pH 5.5. Radiolabeling the lumenal contents of macrophage endosomes confirmed that both the cysteine protease, cathepsin B, and the aspartyl protease, cathepsin D, were present in these vesicles. These proteases were not present on the plasma membrane but were found in early endosomes indicating they are derived from an intracellular source. The presence of proteases with different pH optima in early endosomes suggests that processing in these vesicles may be regulated by changes in endosomal pH. This result represents an important difference in protein processing in endosomes versus lysosomes and provides new insights into the function of endosomal proteases.  相似文献   

10.
Two types of alkaline serine proteases were isolated from the culture filtrate of an alkalophilic actinomycete, Nocardiopsis dassonvillei OPC-210. The enzymes (protease I and protease II) were purified by acetone precipitation, DEAE-Sephadex A-50, CM-Sepharose CL-6B, Sephadex G-75 and phenyl-Toyopearl 650 M column chromatography. The purified enzymes showed a single band on sodium dodecyl sulphate polyacrylamide gel electrophoresis. The molecular weights of proteases I and II were 21,000 and 36,000, respectively. The pIs were 6.4 (protease I) and 3.8 (protease II). The optimum pH levels for the activity of two proteases were pH 10-12 (protease I) and pH 10.5 (protease II). The optimum temperture for the activity of protease I was 70 degrees C and that for protease II was 60 degrees C. Protease I was stable in the range of pH 4.0-8.0 up to 60 degrees C and protease II was stable in the range of pH 6.0-12.0 up to 50 degrees C.  相似文献   

11.
Victoria blight of Avena sativa (oat) is caused by the fungus Cochliobolus victoriae, which is pathogenic because of the production of the toxin victorin. The victorin-induced response in sensitive A. sativa has been characterized as a form of programmed cell death (PCD) and displays morphological and biochemical features similar to apoptosis, including chromatin condensation, DNA laddering, cell shrinkage, altered mitochondrial function, and ordered, substrate-specific proteolytic events. Victorin-induced proteolysis of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) is shown to be prevented by caspase-specific and general protease inhibitors. Evidence is presented for a signaling cascade leading to Rubisco proteolysis that involves multiple proteases. Furthermore, two proteases that are apparently involved in the Rubisco proteolytic cascade were purified and characterized. These proteases exhibit caspase specificity and display amino acid sequences homologous to plant subtilisin-like Ser proteases. The proteases are constitutively present in an active form and are relocalized to the extracellular fluid after induction of PCD by either victorin or heat shock. The role of the enzymes as processive proteases involved in a signal cascade during the PCD response is discussed.  相似文献   

12.
Proteases are an expanding class of drugs that hold great promise. The U.S. FDA (Food and Drug Administration) has approved 12 protease therapies, and a number of next generation or completely new proteases are in clinical development. Although they are a well-recognized class of targets for inhibitors, proteases themselves have not typically been considered as a drug class despite their application in the clinic over the last several decades; initially as plasma fractions and later as purified products. Although the predominant use of proteases has been in treating cardiovascular disease, they are also emerging as useful agents in the treatment of sepsis, digestive disorders, inflammation, cystic fibrosis, retinal disorders, psoriasis and other diseases. In the present review, we outline the history of proteases as therapeutics, provide an overview of their current clinical application, and describe several approaches to improve and expand their clinical application. Undoubtedly, our ability to harness proteolysis for disease treatment will increase with our understanding of protease biology and the molecular mechanisms responsible. New technologies for rationally engineering proteases, as well as improved delivery options, will expand greatly the potential applications of these enzymes. The recognition that proteases are, in fact, an established class of safe and efficacious drugs will stimulate investigation of additional therapeutic applications for these enzymes. Proteases therefore have a bright future as a distinct therapeutic class with diverse clinical applications.  相似文献   

13.
Human immunodeficiency virus type 1 (HIV-1) protease optimally catalyzes in the pH range of 4-6 in contrast to nearly all of the other eukaryotic aspartic proteases, which catalyze best in the pH range of 2-4. A possible structural reason for the higher optimal pH of HIV-1 protease is the absence of a hydrogen bond to the carboxyl group of active-site Asp25, which is nearly universally present in others. To investigate this hypothesis, we have mutated residue 28 in HIV-1 protease from alanine to serine. Both the wild-type and the mutant A28S enzymes have been overexpressed in Escherichia coli using a chemically synthesized gene and purified for a comparative study in enzyme kinetics. The kcat and Km values were determined by a radiometric assay for the wild-type enzyme from pH 3.2 to 7.0, and for the mutant enzyme from pH 3.2 to 6.0. The low pK values of the active site of the free enzyme, pKe1, are 3.3 and 3.4 for the wild-type and mutant enzymes, respectively. The low pK values of the active site of the enzyme bound to substrate, pKes1, are 5.1 and 4.3 for the wild-type and mutant enzymes, respectively. The high pK values of the free enzyme, pKe2, are 6.8 and 5.6, and the corresponding ones for the substrate-bound enzyme, pKes2, are 6.9 and 6.0 for the wild-type and mutant enzymes, respectively. The lowering of pK values in mutant HIV-1 protease indicates that the hydroxyl group of Ser28 forms a new hydrogen bond to active-site Asp25 to increase its acidity.  相似文献   

14.
Six deep-sea proteolytic bacteria taken from Aleutian margin sediments were screened; one of them produced a cold-adapted neutral halophilic protease. These bacteria belong to Pseudoalteromonas spp., which were identified by the 16S rDNA sequence. Of the six proteases produced, two were neutral cold-adapted proteases that showed their optimal activity at pH 7–8 and at temperature close to 35°C, and the other four were alkaline proteases that showed their optimal activity at pH 9 and at temperature of 40–45°C. The neutral cold-adapted protease E1 showed its optimal activity at a sodium chloride concentration of 2 M, whereas the activity of the other five proteases decreased at elevated sodium chloride concentrations. Protease E1 was purified to electrophoretic homogeneity and its molecular mass was 34 kDa, as estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). The molecular weight of protease E1 was determined to be 32,411 Da by mass spectrometric analysis. Phenylmethyl sulfonylfluoride (PMSF) did not inhibit the activity of this protease, whereas it was partially inhibited by ethylenediaminetetra-acetic acid sodium salt (EDTA-Na). De novo amino acid sequencing proved protease E1 to be a novel protein.  相似文献   

15.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

16.
The relationship between a large molecular weight (9S) and a small molecular weight (4.5S, 60,000 molecular weight) species of phytochrome was examined to determine if the larger species was an aggregate of the smaller. Alterations of pH, salt concentration, or phytochrome concentration did not cause any observable formation of the large form from the small form. However, in partially purified phytochrome extracts from Secale cereale L. and Avena sativa L., the large form was converted to the small form over time at 4 C in the dark. This breakdown was inhibitable by the protease inhibitor phenylmethanesulfonyl fluoride. When highly purified large molecular weight rye phytochrome was incubated with a neutral protease isolated from etiolated oat shoots, the large phytochrome was converted to the small form without qualitative visible absorbancy changes. The effect of the oat protease could be mimicked by a wide variety of commercial endopeptidases, including trypsin. Examination of the trypsin-induced breakdown on sodium dodecyl sulfate-polyacrylamide gel electrophoresis revealed that as the size of the photoreversible unit changes from large to small, the size of its constituent polypeptide chains is reduced from 120,000 to 62,000 molecular weight. These experiments provide evidence that the endogenous breakdown observed in extracts is a result of contaminant protease and, consequently, that the small molecular weight species of phytochrome is an artifact due to proteolysis.  相似文献   

17.
Cell extracts from Pyrococcus furiosus were found to contain five proteases, two of which (S66 and S102) are resistant to sodium dodecyl sulfate (SDS) denaturation. Cell extracts incubated at 98 degrees C in the presence of 1% SDS for 24 h exhibited substantial cellular proteolysis such that only four proteins could be visualized by amido black-Coomassie brilliant blue staining of SDS-polyacrylamide gels. The SDS-treated extract retained 19% of the initial proteolytic activity as represented by two proteases, S66 (66 kilodaltons [kDa]) and S102 (102 kDa). Immunoblot analysis with guinea pig sera containing antibodies against protease S66 indicated that S66 is related neither to S102 nor to the other proteases. The results of this analysis also suggest that S66 might be the hydrolysis product of a 200-kDa precursor which does not have proteolytic activity. The 24-h SDS-treated extract showed unusually thermostable proteolytic activity; the measured half-life at 98 degrees C was found to be 33 h. Proteases S66 and S102 were also resistant to denaturation by 8 M urea, 80 mM dithiothreitol, and 5% beta-mercaptoethanol. Purified protease S66 was inhibited by phenylmethylsulfonyl fluoride and diisopropyl fluorophosphate but not by EDTA, ethylene glycol-bis(beta-aminoethyl ether)-N,N,N',N'-tetraacetic acid, or iodoacetic acid. These results indicate that S66 is a serine protease. Amino acid ester hydrolysis studies showed that protease S66 was hydrolytically active towards N-benzoyl-L-arginine ethyl ester.  相似文献   

18.
The profile of sedimentation on a 4-20% (w/v) linear sucrose gradient of the digestive juice of the mollusk Archachatina ventricosa revealed the presence of at least four specific proteases. A first peak, corresponding to a sedimentation coefficient of 3.9 S, contained two endoproteases that could be assayed, one with Leu-pNA and the other with Met-pNA. Their activity was maximal at pH 8.0 and increased in the presence of Ca(2+) ions. Both enzymes were inhibited by the chelating agent 1,10-phenanthroline but their thermal inactivation kinetics were different. A second protease peak was observed at 6.8 S and corresponded to a metallo-endoprotease that hydrolyzed with a maximal activity at pH 8.0 only the amide bonds of peptide substrates having a threonine residue at the P1' position. A last protease peak identified at 9.0 S contained a protease that preferentially acted on tripeptides, such as Val-Pro-Leu (diprotin B) and Thr-Val-Leu, releasing the C-terminal residue. Unlike the proteases identified in the two other peaks, its activity was maximal at acid pH (5.0) and was inhibited by the serine protease inhibitors. Together these results show the potential of A. ventricosa as a source of specific proteases.  相似文献   

19.
Summary The question posed in the present study was: are there hydrolytic enzymes, including proteases, present in the extracellular matrix of the glomerular basement membrane? If these enzymes are present they may play a role in the catabolism of the glomerular basement membrane (GBM) and removal of macromolecular debris resulting from ultrafiltration. Enzymes, acid phosphatase - the marker for lysosomal enzymes - β-galactosidase, β-glucuronidase and acid protease (using albumin as substrate) were biochemically assayed in purified basement membrane preparations. It was found that all enzymes were present in significant amounts in the basement membrane. Compared to other enzymes, acid protease activity was present in much higher amounts. The pH optima of these enzymes were variable but all had significant activity at neutral pH. A method was developed to localize the marker enzyme, acid phosphatase, ultrastructurally in the basement membrane in order to substantiate the biochemical findings. Activity was shown by the presence of dense deposits of lead phosphate. Staining for acid phosphatase could also be shown on isolated, purified basement membrane. The demonstration of acid hydrolases in the GBM matrix argues for their role in (i) the extracellular turnover of basement membrane macromolecules, and (ii) clearance of debris of ultrafiltration which tend to clog the membrane pores.  相似文献   

20.
T sujibo , H., M iyamoto , K., H asegawa , T. & I namori , Y. 1990. Purification and characterization of two types of alkaline serine proteases produced by an alkalophilic actinomycete. Journal of Applied Bacteriology 69 , 520–529.
Two types of alkaline serine proteases were isolated from the culture filtrate of an alkalophilic actinomycete, Nocardiopsis dassonvillei OPC-210. The enzymes (protease I and protease II) were purified by acetone precipitation, DEAE-Sephadex A-50, CM-Sepharose CL-6B, Sephadex G-75 and phenyl-Toyopearl 650 M column chromatography. The purified enzymes showed a single band on sodium dodecyl sulphate polyacrylamide gel electrophoresis. The molecular weights of proteases I and II were 21000 and 36000, respectively. The pIs were 6.4 (protease I) and 3.8 (protease II). The optimum pH levels for the activity of two proteases were pH 10–12 (protease I) and pH 10.5 (protease II). The optimum temperature for the activity of protease I was 70°C and that for protease II was 60°C. Protease I was stable in the range of pH 4.0–8.0 up to 60°C and protease II was stable in the range of pH 6.0–12.0 up to 50°C.  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