首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
A novel single polypeptide endopeptidase of 24 kDa (24k-endopeptidase) was purified with a yield of 300-400 microg/L from conditioned medium of a bacterial strain which was identified as a new species in the genus Chryseobacterium Sp. on the basis of its 16S rDNA sequence and DNA:DNA hybridizations. The NH(2)-terminal amino acid sequence (Val-Ala-Thr-Pro-Asn-Leu-Glu-.) was not found in the availabe databases. The 24k-endopeptidase specifically hydrolyzed the Ser(441)-Val(442) peptide bond in human plasmin(ogen), with additional cleavage of the Lys(78)-Val(79) and Pro(447)-Val(448) peptide bonds, and a secondary cleavage at Lys(615)-Val(616). Thereby, plasminogen is converted into an angiostatin-like fragment containing kringles 1-4 (K1-4) and miniplasminogen (kringle 5 and the serine proteinase domain). The purified K1-4 fragment showed a comparable cytotoxicity toward endothelial cells as the elastase-derived K1-3 fragment (12.7% versus 10.6% at a concentration of 10 microg/mL). Plasminogen, bound to monocytoid THP-1 cells, was also cleaved by the 24k-endopeptidase, resulting in generation of an angiostatin-like fragment and in a decreased capacity to generate cell-associated plasmin following activation by urokinase. The 24k-endopeptidase was not efficiently neutralized by specific inhibitors against the serine, cysteine, aspartic, or matrix metalloproteinase classes of enzymes. In human plasma or serum, however, it induced only very limited plasminogen degradation, apparently due to neutralization of its activity by alpha(2)-macroglobulin. Interaction of this novel 24k-endopeptidase with plasminogen thus yields an angiostatin-like fragment and affects plasmin-mediated cellular proteolytic activity.  相似文献   

2.
Römisch J 《Biological chemistry》2002,383(7-8):1119-1124
Recently a novel serine protease in human plasma was described and was named PHBP, PHBSP or factor seven activating protease (FSAP), respectively, the latter according to the finding that it can support coagulation by factor VII activation. Later on FSAP was identified as a potent activator of single chain plasminogen activators, in particular of prourokinase, as well. The physiological role of FSAP is still speculative, but recent studies suggest a contribution to hemostasis. Due to its affinity to glycosaminoglycans a role in cell-associated or extracellular proteolytic events is also likely. The impact of a very recently uncovered frequent polymorphism impairing the prourokinase activation potential of FSAP needs to be investigated in more detail.  相似文献   

3.
Interactions of the developmentally regulated chondroitin sulfate proteoglycan NG2 with human plasminogen and kringle domain-containing plasminogen fragments have been analyzed by solid-phase immunoassays and by surface plasmon resonance. In immunoassays, the core protein of NG2 binds specifically and saturably to plasminogen, which consists of five kringle domains and a serine protease domain, and to angiostatin, which contains plasminogen kringle domains 1-3. Apparent dissociation constants for these interactions range from 12 to 75 nm. Additional evidence for NG2 interaction with kringle domains comes from its binding to plasminogen kringle domain 4 and to miniplasminogen (kringle domain 5 plus the protease domain) with apparent dissociation constants in the 18-71 nm range. Inhibition of plasminogen and angiostatin binding to NG2 by 6-aminohexanoic acid suggests that lysine binding sites are involved in kringle interaction with NG2. The interaction of NG2 with plasminogen and angiostatin has very interesting functional consequences. 1) Soluble NG2 significantly enhances the activation of plasminogen by urokinase type plasminogen activator. 2) The antagonistic effect of angiostatin on endothelial cell proliferation is inhibited by soluble NG2. Both of these effects of NG2 should make the proteoglycan a positive regulator of the cell migration and proliferation required for angiogenesis.  相似文献   

4.
Matrix metalloproteinase (MMP)-3 inhibited human MDA-MB-231 breast cancer cell invasion through reconstituted basement membrane in vitro. Inhibition of invasion was dependent upon plasminogen and MMP-3 activation, was impaired by the peptide MMP-3 inhibitor Ac-Arg-Cys-Gly-Val-Pro-Asp-NH2 and was associated with: rapid MMP-3-mediated plasminogen degradation to microplasminogen and angiostatin-like fragments; the removal of single-chain urokinase plasminogen activator from MDA-MB-231 cell membranes; impaired membrane plasminogen association; reduced rate of tissue plasminogen activator (t-PA) and membrane-mediated plasminogen activation; and reduced laminin-degrading capacity. Purified human plasminogen lysine binding site-1 (kringles 1-3) exhibited a similar capacity to inhibit MDA-MB-231 invasion, impair t-PA and cell membrane-mediated plasminogen activation and impair laminin degradation by plasmin. Our data provide evidence that MMP-3 can inhibit breast tumour cell invasion in vitro by a mechanism involving plasminogen degradation to fragments that limit plasminogen activation and the degradation of laminin. This supports the hypothesis that MMP-3, under certain conditions, may protect against tumour invasion, which would help to explain why MMP-3 expression, associated with benign and early stage breast tumours, is frequently lost in advanced stage, aggressive, breast disease.  相似文献   

5.
Prourokinase-induced plasminogen activation is complex and involves three distinct reactions: (1) plasminogen activation by the intrinsic activity of prourokinase; (2) prourokinase activation by plasmin; (3) plasminogen activation by urokinase. To further understand some of the mechanisms involved, the effects of epsilon-aminocaproic acid (EACA), a lysine analogue, on these reactions were studied. At a low range of concentrations (10-50 microM), EACA significantly inhibited prourokinase-induced (Glu-/Lys-) plasminogen activation, prourokinase activation by Lys-plasmin, and (Glu-/Lys-) plasminogen activation by urokinase. However, no inhibition of plasminogen activation by Ala158-prourokinase (a plasmin-resistant mutant) occurred. Therefore, the overall inhibition of EACA on prourokinase-induced plasminogen activation was mainly due to inhibition of reactions 2 and 3, by blocking the high-affinity lysine binding interaction between plasmin and prourokinase, as well as between plasminogen and urokinase. These findings were consistent with kinetic studies which suggested that binding of kringle 1-4 of plasmin to the N-terminal region of prourokinase significantly promotes prourokinase activation, and that binding of kringle 1-4 of plasminogen to the C-terminal lysine158 of urokinase significantly promotes plasminogen activation. In conclusion, EACA was found to inhibit, rather than promote, prourokinase-induced plasminogen activation due to its blocking of the high-affinity lysine binding sites on plasmin(ogen).  相似文献   

6.
The complete amino acid sequence of porcine miniplasminogen (Mr 37 600), comprising 341 residues, was determined by automated Edman degradation in a liquid-phase or solid-phase sequenator. Selected fragments were produced by cleavage with 2-(2-nitrophenylsulfenyl)-3-methyl-3'-bromoindolenine (BNPS-skatole), cyanogen bromide, hydroxylamine, Staphylococcus aureus protease or trypsin or with combinations thereof and by activation with urokinase. The sequence obtained was compared with the known sequences of human and bovine miniplasminogen, indicating that the porcine molecule apparently contains the same structural and functional domains as the protein of the other two species. Porcine miniplasminogen has a sequence homology of 83% with human and of 79% with bovine miniplasminogen; 74% of the amino acids are identical in all three species. The results show a higher degree of evolutionary conservatism in the structurally and/or functionally vital regions of the molecule (active site residues, kringle 5).  相似文献   

7.
The human [Glu1]-plasminogen carbohydrate isozymes, plasminogen type I (Pg 1) and plasminogen type II (Pg 2), were separated by chromatography and studied in cell binding experiments at 4 degrees C with primary cultures of rat hepatocytes and rat C6 glioma cells. In both cell systems, Pg 1 and Pg 2 bound to an equivalent number of receptors, apparently representing the same population of surface molecules. The affinity for Pg 2 was slightly higher. With hepatocytes, the KD for Pg 1 was 3.2 +/- 0.2 microM, and the KD for Pg 2 was 1.9 +/- 0.1 microM, as determined from Scatchard transformations of the binding isotherms. The Bmax was approximately the same for both isozymes. With C6 cells, the KD for Pg 1 was 2.2 +/- 0.1 microM vs. 1.5 +/- 0.2 microM for Pg 2. Again, the Bmax was similar with both isozymes. 125I-Pg 1 and 125I-Pg 2 were displaced from specific binding sites by either nonradiolabeled isozyme. The KI for Pg 2 was slightly lower than the KI for Pg 1 with hepatocytes (0.9 vs. 1.3 microM) and with C6 cells (0.6 vs. 1.1 microM). No displacement was detected with miniplasminogen at concentrations up to 5.0 microM. Activation of Pg 1 and Pg 2 by recombinant two-chain tissue-plasminogen activator (rt-PA) was enhanced by hepatocyte cultures. The enhancing effect was greater with Pg 2. Hepatocyte cultures did not affect the activation of miniplasminogen by rt-PA or the activation of plasminogen by streptokinase. Unlike the hepatocytes, C6 cells did not enhance the activation of plasminogen by rt-PA or streptokinase; however, plasmin generated in the presence of C6 cells reacted less readily with alpha 2-antiplasmin.  相似文献   

8.
The contribution of polymorphonuclear neutrophils (PMN) to host defense and natural immunity extends well beyond their traditional role as professional phagocytes. In this study, we demonstrate that upon stimulation with proinflammatory stimuli, human PMN release enzymatic activities that, in vitro, generate bioactive angiostatin fragments from purified plasminogen. We also provide evidence that these angiostatin-like fragments, comprising kringle domain 1 to kringle domain 3 (kringle 1-3) of plasminogen, are generated as a byproduct of the selective proteolytic activity of neutrophil-secreted elastase. Remarkably, affinity-purified angiostatin kringle 1-3 fragments generated by neutrophils inhibited basic fibroblast growth factor plus vascular endothelial growth factor-induced endothelial cell proliferation in vitro, and both vascular endothelial growth factor-induced angiogenesis in the matrigel plug assay and fibroblast growth factor-induced angiogenesis in the chick embryo chorioallantoic membrane assay, in vivo. These results represent the first demonstration that biologically active angiostatin-like fragments can be generated by inflammatory human neutrophils. Because angiostatin is a potent inhibitor of angiogenesis, tumor growth, and metastasis, the data suggest that activated PMN not only act as potent effectors of inflammation, but might also play a critical role in the inhibition of angiogenesis in inflammatory diseases and tumors, by generation of a potent anti-angiogenic molecule.  相似文献   

9.
Glu-plasminogen, kringle 1-5, kringle 1-3, and miniplasminogen exhibited strong binding to human umbilical vein endothelial cells (HUVEC). On the other hand, no significant binding was obtained with microplasminogen and kringle 4. Kringle 1-5 and miniplasminogen, which both contained kringle 5, specifically inhibited the binding of plasminogen to HUVEC while kringle 1-3 did not. The results implied plasminogen molecule contained at least two binding sites, with which it interacted HUVEC. The stronger binding site was located in kringle 5 and the weaker one was in kringle 1-3. Kringle 4 and the active site domain exhibited no significant binding to HUVEC. The interaction of plasminogen with HUVEC is mainly through binding site on kringle 5.  相似文献   

10.
Forty different monoclonal antibodies were produced from hybridomas that were raised against human Lp[a]. Of these, 14 strongly cross-reacted with plasminogen on ELISA screening assays while 16 clearly did not and 10 were only marginally cross-reactive. We took advantage of the homology between plasminogen and apo[a] to define the epitopes of 8 strongly cross-reacting monoclonal antibodies. We were able to subdivide these into four general categories based upon site competition assays (using both plasminogen and Lp[a]), and their reactivity with elastolytically derived plasminogen fragments. Group A monoclonal antibodies (F1 1E3, F2 3A3) recognized epitopes within the kringle 5 and protease domains (miniplasminogen) of plasminogen. The group B monoclonal antibody (F6 1A3) reacted solely with plasminogen kringle 4-like domains and appeared to recognize a limited number of sites on Lp[a]. Group C monoclonal antibodies (F6 1B5, F6 1G9) recognized a second, more frequently distributed site within these kringle 4-like domains. The final group, D, monoclonal antibodies (F6 2C3, F6 2G2, F6 3F4) reacted with a cluster of sites found associated with kringle 4-like domains but also reacted with the miniplasminogen domain. Interestingly, only the members of this group were able to interfere with the proteolytic activity of plasmin. Neither periodate treatment of Lp[a] nor incubation of Lp[a] with epsilon-aminocaproic acid affected the binding of any of our monoclonal antibodies.  相似文献   

11.
Affinity of plasminogen fragments for p-aminobenzamidine-Sepharose was investigated to localize the benzamidine-binding site(s) of the protein. i/ Of the elastase fragments of plasminogen only miniplasminogen (kringle 5 plus light chain) was bound to the column. Kringle 1+2+3 and kringle 4, which carry the lysine-binding sites, were not adsorbed, proving that the lysine-and benzamidine-binding sites are on different domains of the protein. ii/ Light chain was bound to the column even if the primary benzamidine-binding site was covalently blocked, indicating that the protease part of plasmin has a second benzamidine-binding site. iii/ Kringle 5 also binds to the affinity column: the presence of a binding site on kringle 5 raises the possibility that this structure may take part in the interactions of plasminogen with other proteins.  相似文献   

12.
Etscheid M  Beer N  Dodt J 《Cellular signalling》2005,17(12):1486-1494
The hyaluronan-binding protease (HABP) is a serine protease in human plasma which is structurally related to plasminogen activators, coagulation factor XII and hepathocyte growth factor activator. It can in vitro activate the coagulation factor FVII, kininogen and plasminogen activators. The present study was initiated to gain a more complete picture of the cell-associated activities of this fibrinolysis-related protease. Treatment of lung fibroblasts with HABP lead to a rapid activation of signalling pathways, including the mitogen-activated protein kinase (MAPK) pathway with c-Raf, MEK and ERK1/2. Additionally the activation of the PI3K/Akt pathway and of several translation-related proteins was found. Proliferation assays confirmed the assumption of a strong growth-stimulating effect of HABP on human lung and skin fibroblasts. Intracellular signalling and growth stimulation were strongly dependent on the proteolytic activity of HABP. Stimulation of signalling and proliferation by HABP involved the fibroblast growth factor receptor 1 (FGFR-1). HABP-stimulated proliferation of lung fibroblasts MRC-5 was accompanied by a significant intracellular increase in basic fibroblast growth factor (bFGF), the major ligand of FGFR-1; bFGF could however not be identified in the supernatant of HABP-treated cells. Though, the conditioned medium from HABP-treated cells showed a strong growth-promoting activity on quiescent fibroblasts, indicating the release of a yet unknown growth factor amplifying the initial growth stimulus. In a two-dimensional wound model HABP stimulated the invasion of fibroblasts into a scratch area, adding a strong pro-migratory activity to this plasma protease. In summary, HABP exhibits a significant growth factor-like activity on quiescent human lung and dermal fibroblasts. Our findings suggest that this fibrinolysis-related plasma protease may participate in physiologic or pathologic processes where cell proliferation and migration are pivotal, like tissue repair, vascular remodelling, wound healing or tumor development.  相似文献   

13.
Factor VII-activating protease (FSAP) is a novel plasma-derived serine protease structurally homologous to tissue-type and urokinase-type plasminogen activators. We demonstrate that plasminogen activator inhibitor-1 (PAI-1), the predominant inhibitor of tissue-type and urokinase-type plasminogen activators in plasma and tissues, is an inhibitor of FSAP as well. We detected PAI-1.FSAP complexes in addition to high levels of extracellular RNA, an important FSAP cofactor, in bronchoalveolar lavage fluids from patients with acute respiratory distress syndrome. Hydrolytic activity of FSAP was inhibited by PAI-1 with a second-order inhibition rate constant (K(a)) of 3.38 +/- 1.12 x 10(5) m(-1).s(-1). Residue Arg(346) was a critical recognition element on PAI-1 for interaction with FSAP. RNA, but not DNA, fragments (>400 nucleotides in length) dramatically enhanced the reactivity of PAI-1 with FSAP, and 4 microg.ml(-1) RNA increased the K(a) to 1.61 +/- 0.94 x 10(6) m(-1).s(-1). RNA also stabilized the active conformation of PAI-1, increasing the half-life for spontaneous conversion of active to latent PAI-1 from 48.4 +/- 8 min to 114.6 +/- 5 min. In contrast, little effect of DNA on PAI-1 stability was apparent. Residues Arg(76) and Lys(80) in PAI-1 were key elements mediating binding of nucleic acids to PAI-1. FSAP-driven inhibition of vascular smooth muscle cell proliferation was antagonized by PAI-1, suggesting functional consequences for the FSAP-PAI-1 interaction. These data indicate that extracellular RNA and PAI-1 can regulate FSAP activity, thereby playing a potentially important role in hemostasis and cell functions under various pathophysiological conditions, such as acute respiratory distress syndrome.  相似文献   

14.
On the mechanism of fibrin-specific plasminogen activation by staphylokinase   总被引:10,自引:0,他引:10  
The mechanism of plasminogen activation by recombinant staphylokinase was studied both in the absence and in the presence of fibrin, in purified systems, and in human plasma. Staphylokinase, like streptokinase, forms a stoichiometric complex with plasminogen that activates plasminogen following Michaelis-Menten kinetics with Km = 7.0 microM and k2 = 1.5 s-1. In purified systems, alpha 2-antiplasmin inhibits the plasminogen-staphylokinase complex with k1(app) = 2.7 +/- 0.30 x 10(6) M-1 s-1 (mean +/- S.D., n = 12), but not the plasminogen-streptokinase complex. Addition of 6-aminohexanoic acid induces a concentration-dependent reduction of k1(app) to 2.0 +/- 0.17 x 10(4) M-1 s-1 (mean +/- S.D., n = 5) at concentrations greater than or equal to 30 mM, with a 50% reduction at a 6-aminohexanoic acid concentration of 60 microM. Staphylokinase does not bind to fibrin, and fibrin stimulates the initial rate of plasminogen activation by staphylokinase only 4-fold. Staphylokinase induces a dose-dependent lysis of a 0.12-ml 125I-fibrin-labeled human plasma clot submersed in 0.5 ml of citrated human plasma; 50% lysis in 2 h is obtained with 17 nM staphylokinase and is associated with only 5% plasma fibrinogen degradation. Corresponding values for streptokinase are 68 nM and more than 90% fibrinogen degradation. In the absence of a fibrin clot, 50% fibrinogen degradation in human plasma in 2 h requires 790 nM staphylokinase, but only 4.4 nM streptokinase. These results suggest the following mechanism for relatively fibrin-specific clot lysis with staphylokinase in a plasma milieu. In plasma in the absence of fibrin, the plasminogen-staphylokinase complex is rapidly neutralized by alpha 2-antiplasmin, thus preventing systemic plasminogen activation. In the presence of fibrin, the lysine-binding sites of the plasminogen-staphylokinase complex are occupied and inhibition by alpha 2-antiplasmin is retarded, thus allowing preferential plasminogen activation at the fibrin surface.  相似文献   

15.
Purification of alpha 2-plasmin inhibitor (alpha 2PI) from human plasma by affinity chromatography on plasminogen-Sepharose resulted in copurification of a contaminating protein with Mr 17,000 as judged by sodium dodecyl sulphate/polyacrylamide gel electrophoresis. This contaminating protein could not be removed from the purified alpha 2-PI preparation by several types of gel chromatography applied. The use of the kringle 1-3 part of plasminogen, K(1 + 2 + 3), bound to Sepharose for affinity chromatography, instead of plasminogen-Sepharose, resulted in an alpha 2PI preparation without this contaminant. The contaminating protein was found to interact specifically with the kringle 4 part of plasminogen (K4) and not with K(1 + 2 + 3) or miniplasminogen. The K4-binding protein was purified by ammonium sulphate precipitation, affinity chromatography on K4-Sepharose, ion-exchange chromatography and gel filtration on AcA 34. The relative molecular mass of the protein (Mr 68 000) was estimated by gel filtration. This suggests a tetrameric protein composed of four subunits (Mr 17,000), that are dissociated by 1% sodium dodecyl sulphate. Dissociation into subunits was also demonstrated by gel filtration in the presence of 6 M guanidine hydrochloride. A specific antibody was raised in rabbits against the purified protein and this antibody was shown not to react with any known fibrinolytic components. The pI of the K4-binding protein was found to be 5.8. The first three N-terminal amino acids were determined to be Glu-Pro-Pro. The concentration of the protein in plasma was estimated to be 0.20 +/- 0.03 microM (15 +/- 2 mg/l). The electrophoretic mobility of the K4-binding protein was shown by crossed immunoelectrophoresis to be influenced by the presence of Ca2+, EDTA and heparin. The protein was found to enhance plasminogen activation catalyzed by tissue-type plasminogen activator (t-PA) in the presence of poly(D-lysine). The protein appeared to be a novel plasma protein tentatively called 'tetranectin'.  相似文献   

16.
It was shown that activation of two native plasminogen and miniplasminogen forms by the tissue activator in the presence of fibrin obeys the Michaelis-Menten kinetics. The kinetic parameters of activation of both plasminogen native forms differ insignificantly. For miniplasminogen whose molecule contains no lysine-binding sites, a marked decrease of activation power was observed. The Km value of activator for miniplasminogen is 10 times that of plasminogen form I and 20 times that of plasminogen form II. The kcat/Km value of activator for miniplasminogen is 7 times less than that of plasminogen form I and by one order of magnitude more than that of plasminogen form II. These results testify to the importance of lysine-binding sites in the native plasminogen molecule during the activation of fibrinolysis by the major physiological activator.  相似文献   

17.
Reciprocal activation of prourokinase (pro-u-PA) and plasminogen is an important mechanism in the initiation and propagation of local fibrinolytic activity. We found that glucosyldiacylglycerol (GDG) enhanced the reciprocal activation by 1.5- to 2-fold at 0.7-16 microM, accompanying increased conversions of both zymogens to active two-chain forms. The reciprocal activation system consists of (i) plasminogen activation by pro-u-PA to form plasmin, (ii) pro-u-PA activation by the resulting plasmin to form two-chain u-PA (tcu-PA), and (iii) plasminogen activation by the resulting tcu-PA. Whereas GDG minimally affected steps (ii) and (iii) in isolated systems, it markedly enhanced step (i) in the absence of the conversion of pro-u-PA to tcu-PA. GDG significantly increased the intrinsic fluorescence of pro-u-PA (6.7%), but not that of tcu-PA or plasminogen. The large change in intrinsic fluorescence suggests that GDG selectively affects pro-u-PA to alter its conformation, and this mechanism may account for enhancement of its intrinsic plasminogen activator activity.  相似文献   

18.
The mammalian protease plasminogen can be activated by bacterial activators, the three-domain (alpha, beta, gamma) streptokinases and the one-domain (alpha) staphylokinases. These activators act as plasmin(ogen) cofactors, and the resulting complexes initiate proteolytic activity of host plasminogen which facilitates bacterial colonization of the host organism. We have investigated the kinetic mechanism of the plasminogen activation mediated by a novel two-domain (alpha, beta) streptokinase isolated from Streptococcus uberis (Sk(U)) with specificity toward bovine plasminogen. The interaction between Sk(U) and plasminogen occurred in two steps: (1) rapid association of the proteins and (2) slow transition to the active complex Sk(U)-PgA. The complex Sk(U)-PgA converted plasminogen to plasmin with the following parameters: K(m) < or = 1.5 microM and k(cat) = 0.55 s(-)(1). The ability of proteolytic fragments of Sk(U) to activate plasminogen was investigated. Only two C-terminal segments (97-261 and 123-261), which both contain the beta-domain (126-261), were shown to be active. They initiated plasminogen activation in complex with plasmin, but not with plasminogen, and thereby exhibited functional similarity to the staphylokinase. The fusion protein His(6)-Sk(U) (i.e., Sk(U) with a small N-terminal tag) acted exclusively in complex with plasmin as well. These observations demonstrate that (1) the N-terminal alpha-domain, including a native N-terminus, was necessary for "virgin" activation of the associated plasminogen in the Sk(U)-PgA complex and (2) the C-terminal beta-domain of Sk(U) is important for recognition of the substrate in the Sk(U)-PgA complex.  相似文献   

19.
Plasminogen activator inhibitor-1 (PAI-1) is a serpin protease inhibitor that binds plasminogen activators (uPA and tPA) at a reactive center loop located at the carboxyl-terminal amino acid residues 320-351. The loop is stretched across the top of the active PAI-1 protein maintaining the molecule in a rigid conformation. In the latent PAI-1 conformation, the reactive center loop is inserted into one of the beta sheets, thus making the reactive center loop unavailable for interaction with the plasminogen activators. We truncated porcine PAI-1 at the amino and carboxyl termini to eliminate the reactive center loop, part of a heparin binding site, and a vitronectin binding site. The region we maintained corresponds to amino acids 80-265 of mature human PAI-1 containing binding sites for vitronectin, heparin (partial), uPA, tPA, fibrin, thrombin, and the helix F region. The interaction of "inactive" PAI-1, rPAI-1(23), with plasminogen and uPA induces the formation of a proteolytic protein with angiostatin properties. Increasing amounts of rPAI-1(23) inhibit the proteolytic angiostatin fragment. Endothelial cells exposed to exogenous rPAI-1(23) exhibit reduced proliferation, reduced tube formation, and 47% apoptotic cells within 48 h. Transfected endothelial cells secreting rPAI-1(23) have a 30% reduction in proliferation, vastly reduced tube formation, and a 50% reduction in cell migration in the presence of VEGF. These two studies show that rPAI-1(23) interactions with uPA and plasminogen can inhibit plasmin by two mechanisms. In one mechanism, rPAI-1(23) cleaves plasmin to form a proteolytic angiostatin-like protein. In a second mechanism, rPAI-1(23) can bind uPA and/or plasminogen to reduce the number of uPA and plasminogen interactions, hence reducing the amount of plasmin that is produced.  相似文献   

20.
The reciprocal activation of plasminogen and prourokinase (pro-u-PA) is an important mechanism in the initiation and propagation of local fibrinolytic activity. We have found that a bacterial lipopeptide compound, surfactin C (3-20 microM), enhances the activation of pro-u-PA in the presence of plasminogen. This effect accompanied increased conversions of both pro-u-PA and plasminogen to their two-chain forms. Surfactin C also elevated the rate of plasminogen activation by two-chain urokinase (tcu-PA) while not affecting plasmin-catalyzed pro-u-PA activation and amidolytic activities of tcu-PA and plasmin. The intrinsic fluorescence of plasminogen was increased, and molecular elution time of plasminogen in size-exclusion chromatography was shortened in the presence of surfactin C. These results suggested that surfactin C induced a relaxation of plasminogen conformation, thus leading to enhancement of u-PA-catalyzed plasminogen activation, which in turn caused feedback pro-u-PA activation. Surfactin C was active in enhancing [125I]fibrin degradation both by pro-u-PA/plasminogen and tcu-PA/plasminogen systems. In a rat pulmonary embolism model, surfactin C (1 mg/kg, i.v.) elevated 125I plasma clot lysis when injected in combination with pro-u-PA. The present results provide first evidence that pharmacological relaxation of plasminogen conformation leads to enhanced fibrinolysis in vivo.  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