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 281 毫秒
1.
Previous results have shown that the autoantibody eluted from the glomeruli of rats with active Heymann nephritis contain a population of antibodies not only to the putative autoantigen of the disease, gp330, but alos to plasminogen. Since gp330 has been shown to serve as a receptor for plasminogen, we have analyzed the effects of autoantibody on plasminogen-binding to gp330 and activation of plasminogen to plasmin by urokinase. Autoantibody does not inhibit the binding of plasminogen to gp330. The change in the conformation of plasminogen when its lysine-binding sites are occupied or after conversion to plasmin results in a significant decrease in autoantibody-binding. The most significant effect of autoantibody on this system is the inhibition of plasminogen activation to plasmin by urokinase. The binding of autoantibody to plasminogen acts as a competitive inhibitor of the reaction by apparently blocking access of urokinase to plasminogen's activation site. These results indicate that autoantibody obtained from the immune deposits in the glomeruli of rats with active Heyman nephritis does not inhibit the binding of plasminogen to gp330 but does significantly alter the urokinase catalyzed activation of plasminogen to plasmin.  相似文献   

2.
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   

3.
Previous results have demonstrated the binding of a 76- and 80-kDa serum protein to the Heymann nephritis autoantigen, gp330. This 76-kDa serum protein was purified by column chromatography and preparative sodium dodecyl sulfate-polyacrylamide gel electrophoresis. A rabbit polyclonal antibody for the serum protein was produced and used to screen a rat liver cDNA expression library. Sequence analysis of an isolated clone identified the serum protein as plasminogen. Plasminogen was isolated from rat serum by standard techniques, and the binding of plasminogen to gp330 was confirmed by Western analysis. Enzyme-linked immunosorbent assay results demonstrated a time-dependent, saturable, and specifically inhibitable binding of plasminogen to gp330. There was no significant difference in the binding of the two carbohydrate forms of plasminogen to gp330. Plasminogen binding to gp330 could be completely inhibited by the addition of exogenous gp330. This binding could also be partially inhibited by benzamidine but only slightly by the lysine analogue, epsilon-aminocaproic acid. However, even a combination of these two inhibitors could not completely block the binding of plasminogen to gp330 indicating that gp330 may be binding to plasminogen through some other unknown interactions. These results demonstrate that gp330 is a receptor site for plasminogen.  相似文献   

4.
Untreated Staphylococcus aureus cells, strain Cowan I, specifically bound 125I-Glu-plasminogen. The binding was inhibited by both unlabeled Glu-plasminogen and Glu-plasmin. The Lys form of plasminogen, which lacks the 8-kDa amino-terminal activation peptide, was approximately 100-fold more effective than the Glu form in competing with the binding of 125I-labeled Glu-plasminogen. This suggests an increase in binding affinity upon removal of the activation peptide. Fibronectin, fibrinogen and IgG, plasma components known to bind to the staphylococcal surface, did not significantly interfere with the binding. The competing activity in plasma was abolished by specifically absorbing plasminogen from the plasma sample. L-Lysine and a fragment of plasminogen containing three of the first five protein attachment domains present in the molecule (kringle structures) also competed with plasminogen for binding suggesting that the lysine-binding sites of plasminogen were involved in its interaction with staphylococci. Scatchard analysis revealed high- and low-affinity binding sites. Kd and the number of high-affinity binding sites were 1.7 nM and 780 binding sites/bacterial cell, respectively. 125I-Glu-plasminogen bound to staphylococcal surface was converted to plasmin by tissue-type plasminogen activator. The conversion took place also in the presence of plasma. If the conversion was carried out in the absence of low-molecular-mass plasmin inhibitors such as aprotinin, the bound Glu-plasmin was further converted to Lys-plasmin. The surface-bound plasmin was enzymically active, as judged by digestion of the synthetic substrate, S-2251. The plasminogen conversion shown by the present experiments not only leads to the surface-bound plasmin but seems to considerably increase the affinity of plasmin for its binding site. This may represent a physiologically relevant method for a bacterial cell to retain surface-bound active plasmin which is also protected from its soluble plasma inhibitors. This novel mechanism for staphylococci to adopt surface-bound proteolytic activity, without the interference of plasma components, may have some role in the tissue penetration and invasion of microbes during infection.  相似文献   

5.
Human HT-1080 fibrosarcoma cells produce urokinase-type plasminogen activator (u-PA) and type 1 plasminogen activator inhibitor (PAI-1). We found that after incubation of monolayer cultures with purified native human plasminogen in serum-containing medium, bound plasmin activity could be eluted from the cells with tranexamic acid, an analogue of lysine. The bound plasmin was the result of plasminogen activation on the cell surface; plasmin activity was not taken up onto cells after deliberate addition of plasmin to the serum-containing medium. The cell surface plasmin formation was inhibited by an anticatalytic monoclonal antibody to u-PA, indicating that this enzyme was responsible for the activation. Preincubation of the cells with diisopropyl fluorophosphate-inhibited u-PA led to a decrease in surface-bound plasmin, indicating that a large part, if not all, of the cell surface plasminogen activation was catalyzed by surface-bound u-PA. In the absence of plasminogen, most of the cell surface u-PA was present in its single-chain proenzyme form, while addition of plasminogen led to formation of cell-bound two-chain u-PA. The latter reaction was catalyzed by cell-bound plasmin. Cell-bound u-PA was accessible to inhibition by endogenous PAI-1 and by added PAI-2, while the cell-bound plasmin was inaccessible to serum inhibitors, but accessible to added aprotinin and an anticatalytic monoclonal antibody. A model for cell surface plasminogen activation is proposed in which plasminogen binding to cells from serum medium is followed by plasminogen activation by trace amounts of bound active u-PA, to form bound plasmin, which in turn serves to produce more active u-PA from bound pro-u-PA. This exponential process is subject to regulation by endogenous PAI-1 and limited to the pericellular space.  相似文献   

6.
Floden AM  Watt JA  Brissette CA 《PloS one》2011,6(11):e27502
Borrelia burgdorferi is the causative agent of Lyme disease, the most commonly reported arthropod-borne disease in the United States. B. burgdorferi is a highly invasive bacterium, yet lacks extracellular protease activity. In order to aid in its dissemination, B. burgdorferi binds plasminogen, a component of the hosts' fibrinolytic system. Plasminogen bound to the surface of B. burgdorferi can then be activated to the protease plasmin, facilitating the bacterium's penetration of endothelial cell layers and degradation of extracellular matrix components. Enolases are highly conserved proteins with no sorting sequences or lipoprotein anchor sites, yet many bacteria have enolases bound to their outer surfaces. B. burgdorferi enolase is both a cytoplasmic and membrane associated protein. Enolases from other pathogenic bacteria are known to bind plasminogen. We confirmed the surface localization of B. burgdorferi enolase by in situ protease degradation assay and immunoelectron microscopy. We then demonstrated that B. burgdorferi enolase binds plasminogen in a dose-dependent manner. Lysine residues were critical for binding of plasminogen to enolase, as the lysine analog εaminocaproic acid significantly inhibited binding. Ionic interactions did not play a significant role in plasminogen binding by enolase, as excess NaCl had no effects on the interaction. Plasminogen bound to recombinant enolase could be converted to active plasmin. We conclude that B. burgdorferi enolase is a moonlighting cytoplasmic protein which also associates with the bacterial outer surface and facilitates binding to host plasminogen.  相似文献   

7.
Since the signal transduction mechanisms responsible for liver regeneration mediated by the plasminogen/plasmin system remain largely undetermined, we have investigated whether plasmin regulates the pro-apoptotic protein Bim(EL) in primary hepatocytes. Plasmin bound to hepatocytes in part via its lysine binding sites (LBS). Plasmin also triggered phosphorylation of ERK1/2 without cell detachment. The plasmin-induced phosphorylation of ERK1/2 was inhibited by the LBS inhibitor epsilon-aminocaproic acid (EACA), the serine protease inhibitor aprotinin, and the MEK inhibitor PD98059. DFP-inactivated plasmin failed to phosphorylate ERK1/2. Plasmin temporally decreased the starvation-induced expression of Bim(EL) and activation of caspase-3 via the ERK1/2 signaling pathway, resulting in an enhancement of cell survival. The amount of mRNA for Bim increased 1 day after the injection of CCl(4) in livers of plasminogen knockout (Plg-KO) and the wild-type (WT) mice. The increase in Bim(EL) protein persisted for at least 7 days post-injection in livers of Plg-KO mice, whereas WT mice showed an increase in Bim(EL) protein 1 day after the injection. Plg-KO and WT mice showed notable phosphorylation of ERK1/2 7 and 3 days after the injection of CCl(4), respectively. Our data suggest that the plasminogen/plasmin system could decrease Bim(EL) expression via the ERK1/2 signaling pathway during liver regeneration.  相似文献   

8.
9.
It was demonstrated that plasminogen and the plasmin heavy chain form a complex with an immobilized fibrinogen fragment E. The E-fragment interacts, in its turn, with the immobilized heavy chain; this interaction is provided for by the lysin binding sites of the plasminogen molecule. The plasmin light chain having no lysin binding sites is specifically absorbed on the immobilized fragment D, whereas the D-fragment--on the immobilized light chain. The elution is caused by arginine or benzamidine; 6-aminohexanoic acid does not affect this interaction. It is assumed that the interaction of plasminogen and plasmin with fibrin is provided for not only by the lysine binding but also by the benzamidine binding sites of the plasminogen molecule.  相似文献   

10.
Vitronectin (VN) has been implicated as a major matrix-associated regulator component of plasminogen activation by serving as a potent stabilizing cofactor of plasminogen activator inhibitor-1 (PAI-1). The direct binding of heparin, plasminogen as well as PAI-1 in its latent and active form to immobilized VN was studied in the absence or presence of competitors. Monoclonal antibodies against the carboxyl-terminal portion of VN inhibited both PAI-1 and plasminogen binding, whereas heparin, heparan sulfate with a high degree of sulfation, or dextran sulfate interfered with PAI-1 binding (KD = 20 nM) only. Utilizing synthetic peptides encompassing overlapping sequences of the heparin-binding domain of VN, adjacent heparin and PAI-1-binding sites were localized within the sequence 348-370 of VN. Although a number of other serine protease inhibitors which do not form binary complexes with VN contain a reactive-site Ser at their P1'-position, a reactive-site P1' mutant of PAI-1 (Met----Ser) showed comparable if not increased binding to VN. Binding of Lys-plasminogen and active-site-blocked plasmin was at least 10-fold higher in affinity (KD = 85-100 nM) compared to Glu-plasminogen (KD approximately 1 microM) and could be inhibited by lysine analogs but not by glycosaminoglycans or PAI-1, indicating that heteropolar plasmin(ogen) binding of VN occurs to an adjacent segment upstream to the heparin and PAI-1-binding sites. This contention was further supported in binding studies with plasmin-modified VN which lost both heparin and PAI-1 binding but exhibited 2-3-fold higher capacity to bind plasminogen. The essential plasmin(ogen)-binding site was mapped by ligand blot analysis to the carboxyl-terminal portion of proteolytically trimmed VN (M(r) = 61,000). Moreover, treatment of the extracellular matrix of human umbilical vein endothelial cells with plasmin resulted in partial degradation of matrix-associated VN and concomitant release of PAI-1, but increased the ability of the matrix by about 2-fold to bind plasminogen. These results are indicative of differential interactions of VN with components of the plasminogen activation system, whereby plasmin itself may provoke the switch of VN from an anti-fibrinolytic into a pro-fibrinolytic cofactor. This process reflects a novel role for the adhesive protein and its degradation product(s) in the possible feedback regulation of localized plasmin formation at extracellular sites.  相似文献   

11.
We have found that tissue plasminogen activator catalyzes the binding of plasminogen (Pg) to immunoglobulin G (IgG) immobilized on a surface. This enhancement is due to the formation of plasmin, since plasmin treatment of immobilized IgG produced a 20-fold increase in Pg binding. Pg binding is lysine site dependent and reversible. The augmentation of Pg binding by plasmin is specific as other proteases produced significantly less or no effect. Immobilized plasmin-treated IgG also specifically binds Pg in plasma. IgG-immobilized Pg is activated by tissue plasminogen activator, and a significant portion of the plasmin formed remains bound to the IgG. The Pg reactive species in a plasmin-treated IgG digest was identified as the Fab fragment by chromatography utilizing the immobilized high affinity lysine-binding site of plasminogen. Specificity of the interaction was further demonstrated by immunoblot-ligand analysis which demonstrated that the plasmin-derived Fab fragment bound Pg whereas papain-derived Fab or plasmin-derived Fc fragments did not. These data suggest that Pg binds to the new COOH-terminal lysine residue of the plasmin-derived Fab. Pg also binds to an immobilized immune complex following plasmin treatment. These findings indicate that surface-bound IgG localizes plasminogen thus extending the spectrum of activity of the plasmin system to immunologic reactions.  相似文献   

12.
Lin LF  Houng A  Reed GL 《Biochemistry》2000,39(16):4740-4745
Lysine side chains induce conformational changes in plasminogen (Pg) that regulate the process of fibrinolysis or blood clot dissolution. A lysine side-chain mimic, epsilon amino caproic acid (EACA), enhances the activation of Pg by urinary-type and tissue-type Pg activators but inhibits Pg activation induced by streptokinase (SK). Our studies of the mechanism of this inhibition revealed that EACA (IC(50) 10 microM) also potently blocked amidolytic activity by SK and Pg at doses nearly 10000-fold lower than that required to inhibit the amidolytic activity of plasmin. Different Pg fragments were used to assess the role of the kringles in mediating the inhibitory effects of EACA: mini-Pg which lacks kringles 1-4 of Glu-Pg and micro-Pg which lacks all kringles and contains only the catalytic domain. SK bound with similar affinities to Glu-Pg (K(A) = 2.3 x 10(9) M(-1)) and to mini-Pg (K(A) = 3.8 x 10(9) M(-)(1)) but with significantly lower affinity to micro-Pg (K(A) = 6 x 10(7) M(-)(1)). EACA potently inhibited the binding of Glu-Pg to SK (K(i) = 5.7 microM), but was less potent (K(i) = 81.1 microM) for inhibiting the binding of mini-Pg to SK and had no significant inhibitory effects on the binding of micro-Pg and SK. In assays simulating substrate binding, EACA also potently inhibited the binding of Glu-Pg to the SK-Glu-Pg activator complex, but had negligible effects on micro-Pg binding. Taken together, these studies indicate that EACA inhibits Pg activation by blocking activator complex formation and substrate binding, through a kringle-dependent mechanism. Thus, in addition to interactions between SK and the protease domain, interactions between SK and the kringle domain(s) play a key role in Pg activation.  相似文献   

13.
L A Miles  E F Plow 《Biochemistry》1986,25(22):6926-6933
An antibody population that reacted with the high-affinity lysine binding site of human plasminogen was elicited by immunizing rabbits with an elastase degradation product containing kringles 1-3 (EDP I). This antibody was immunopurified by affinity chromatography on plasminogen-Sepharose and elution with 0.2 M 6-aminohexanoic acid. The eluted antibodies bound [125I]EDP I, [125I]Glu-plasminogen, and [125I]Lys-plasminogen in radioimmunoassays, and binding of each ligand was at least 99% inhibited by 0.2 M 6-aminohexanoic acid. The concentrations for 50% inhibition of [125I]EDP I binding by tranexamic acid, 6-aminohexanoic acid, and lysine were 2.6, 46, and 1730 microM, respectively. Similar values were obtained with plasminogen and suggested that an unoccupied high-affinity lysine binding site was required for antibody recognition. The antiserum reacted exclusively with plasminogen derivatives containing the EDP I region (EDP I, Glu-plasminogen, Lys-plasminogen, and the plasmin heavy chain) and did not react with those lacking an EDP I region [miniplasminogen, the plasmin light chain or EDP II (kringle 4)] or with tissue plasminogen activator or prothrombin, which also contain kringles. By immunoblotting analyses, a chymotryptic degradation product of Mr 20,000 was derived from EDP I that retained reactivity with the antibody. The high-affinity lysine binding site was equally available to the antibody probe in Glu- and Lys-plasminogen and also appeared to be unoccupied in the plasmin-alpha 2-antiplasmin complex. alpha 2-Antiplasmin inhibited the binding of radiolabeled EDP I, Glu-plasminogen, or Lys-plasminogen by the antiserum, suggesting that the recognized site is involved in the noncovalent interaction of the inhibitor with plasminogen.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

14.
Rajamohan G  Dikshit KL 《FEBS letters》2000,474(2-3):151-158
Staphylokinase (SAK) forms an inactive 1:1 complex with plasminogen (PG), which requires both the conversion of PG to plasmin (Pm) to expose an active site in PG-SAK activator complex and the amino-terminal processing of SAK to expose the positively charged (Lys-11) amino-terminus after removal of the 10 N-terminal amino acid residues from the full length protein. The mechanism by which the N-terminal segment of SAK affects its PG activation capability was investigated by generating SAK mutants, blocked in the native amino-terminal processing site of SAK, and carrying an alteration in the placement of the positively charged amino acid residue, Lys-11, and further studying their interaction with PG, Pm, miniplasmin and kringle structures. A ternary complex formation between PG-SAK PG was observed when an immobilized PG-SAK binary complex interacted with free radiolabelled PG in a sandwich binding experiment. Formation of this ternary complex was inhibited by a lysine analog, 6-aminocaproic acid (EACA), in a concentration dependent manner, suggesting the involvement of lysine binding site(s) in this process. In contrast, EACA did not significantly affect the formation of binary complex formed by native SAK or its mutant derivatives. Furthermore, the binary (activator) complex formed between PG and SAK mutant, PRM3, lacking the N-terminal lysine 11, exhibited 3-4-fold reduced binding with PG, Pm or miniplasmin substrate during ternary complex formation as compared to native SAK. Additionally, activator complex formed with PRM3 failed to activate miniplasminogen and exhibited highly diminished activation of substrate PG. Protein binding studies indicated that it has 3-5-fold reduction in ternary complex formation with miniplasmin but not with the kringle structure. In aggregate, these observations provide experimental evidence for the participation of the N-terminal region of SAK in accession and processing of substrate by the SAK-Pm activator complex to potentiate the PG activation by enhancing and/or stabilizing the interaction of free PG.  相似文献   

15.
We recently discovered several nonlysine-analog conformational modulators for plasminogen. These include SMTP-6, thioplabin B and complestatin that are low molecular mass compounds of microbial origin. Unlike lysine-analog modulators, which increase plasminogen activation but inhibit its binding to fibrin, the nonlysine-analog modulators enhance both activation and fibrin binding of plasminogen. Here we show that some nonlysine-analog modulators promote autoproteolytic generation of plasmin(ogen) derivatives with its catalytic domain undergoing extensive fragmentation (PMDs), which have angiostatin-like anti-endothelial activity. The enhancement of urokinase-catalyzed plasminogen activation by SMTP-6 was followed by rapid inactivation of plasmin due to its degradation mainly in the catalytic domain, yielding PMD with a molecular mass ranging from 68 to 77 kDa. PMD generation was observed when plasmin alone was treated with SMTP-6 and was inhibited by the plasmin inhibitor aprotinin, indicating an autoproteolytic mechanism in PMD generation. Thioplabin B and complestatin, two other nonlysine-analog modulators, were also active in producing similar PMDs, whereas the lysine analog 6-aminohexanoic acid was inactive while it enhanced plasminogen activation. Peptide sequencing and mass spectrometric analyses suggested that plasmin fragmentation was due to cleavage at Lys615-Val616, Lys651-Leu652, Lys661-Val662, Lys698-Glu699, Lys708-Val709 and several other sites mostly in the catalytic domain. PMD was inhibitory to proliferation, migration and tube formation of endothelial cells at concentrations of 0.3-10 microg.mL(-1). These results suggest a possible application of nonlysine-analog modulators in the treatment of cancer through the enhancement of endogenous plasmin(ogen) fragment formation.  相似文献   

16.
Streptococcus canis is an important zoonotic pathogen capable of causing serious invasive diseases in domestic animals and humans. In the present paper we report the binding of human plasminogen to S. canis and the recruitment of proteolytically active plasmin on its surface. The binding receptor for plasminogen was identified as a novel M-like protein designated SCM (S. canis M-like protein). SPR (surface plasmon resonance) analyses, radioactive dot-blot analyses and heterologous expression on the surface of Streptococcus gordonii confirmed the plasminogen-binding capability of SCM. The binding domain was located within the N-terminus of SCM, which specifically bound to the C-terminal part of plasminogen (mini-plasminogen) comprising kringle domain 5 and the catalytic domain. In the presence of urokinase, SCM mediated plasminogen activation on the bacterial surface that was inhibited by serine protease inhibitors and lysine amino acid analogues. Surface-bound plasmin effectively degraded purified fibrinogen as well as fibrin clots, resulting in the dissolution of fibrin thrombi. Electron microscopic illustration and time-lapse imaging demonstrated bacterial transmigration through fibrinous thrombi. The present study has led, for the first time, to the identification of SCM as a novel receptor for (mini)-plasminogen mediating the fibrinolytic activity of S. canis.  相似文献   

17.
V Gurewich 《Enzyme》1988,40(2-3):97-108
Single chain urokinase (SC-UK) is a precursor of 55 kd two-chain UK (TC-UK). Treatment with catalytic proportions of plasmin or kallikrein converts SC-UK to TC-UK as a consequence of cleavage of its Lys158-Ile159 peptide bond. This plasmin-mediated activation of SC-UK induces a positive feedback secondary reaction and complicates measurement of its activity against its natural substrate, Glu-plasminogen. The fibrin-selective effect of pro-UK-induced clot lysis is not related to fibrin binding. Rather, a conformational change in Glu-plasminogen, conferred when it binds to certain carboxy-terminal lysine residues on fibrin, has been implicated in this mechanism. This is complementary to t-PA. Fibrin-bound t-PA was found to exclusively activate plasminogen bound to certain internal lysine residues. Their complementariness is believed to explain their synergism in fibrinolysis.  相似文献   

18.
Active-site-blocked, fluorescent derivatives of tPA (Activase) and a variant (delta FEIX) which lacks the finger and epidermal growth factor-like domains and possesses Asn to Gln and Val to Met mutations at residues 117 and 245, respectively, were prepared. The binding of these to fibrin was studied by adding them at systematically varying concentrations to fibrinogen, at a fixed concentration, inducing clotting with thrombin, separating free and bound tPA or delta FEIX by centrifugation, and measuring the concentration of unbound material by extrinsic fluorescence. Similar studies were performed with Glu and Lys-plasminogen, using intrinsic fluorescence. epsilon-amino caproic acid (EACA) was utilized to distinguish kringle-dependent from finger-dependent binding. In the absence of EACA, delta FEIX-bound fibrin through a single class of sites with Kd = 0.69 microM and n = 1.34 delta FEIX/fibrin. The binding of delta FEIX was completely inhibited by EACA and 50% displacement occurred at [EACA] = 300 microM. Fibrin-bound tPA was only partially displaced with EACA. In the presence of 30 mM EACA, tPA binding reflected a single class of sites with Kd = 0.26 microM and n = 0.60 tPA/fibrin. In the absence of EACA, tPA binding was complex, typified by downwardly curved Scatchard plots, and was consistent with interactions of the two classes of sites, characterized by Kd = 0.13 microM, n = 0.60 and Kd = 0.61 microM, n = 1.23. These were attributed to finger and kringle-dependent interactions, respectively. Under the experimental conditions employed, Glu-plasminogen exhibited no binding to fibrin, whereas Lys-plasminogen bound to a single class of sites with Kd = 0.25 microM and n = 1.02 plasminogen/fibrin. This binding was completely inhibited by EACA and 50% displacement occurred at [EACA] = 28 microM. Competition experiments indicated that Lys-plasminogen does not displace either tPA or delta FEIX from fibrin. From these results the conclusions are drawn that tPA can interact with intact fibrin by two different and independent modes, involving, respectively, the finger and kringle 2 domains, and neither of these modes are competitive with the kringle-dependent binding of Lys-plasminogen.  相似文献   

19.
20.
Binding and activation of plasminogen on the platelet surface   总被引:18,自引:0,他引:18  
A mechanism by which platelets might participate in fibrinolysis by binding plasminogen and influencing its activation has been examined. Binding of radioiodinated human Glu-plasminogen to washed human platelets was time-dependent and was enhanced 3-9-fold by stimulation of platelets with thrombin but not with ADP. The interaction with both stimulated and unstimulated cells was specific, saturable, divalent ion-independent, and reversible. The platelet-bound ligand had the molecular weight of plasminogen, and no conversion to plasmin was detected. Scatchard analyses provided evidence for a single class of plasminogen-binding sites on both stimulated and unstimulated cells. The Kd for thrombin-stimulated platelets was 2.6 +/- 1.3 microM, and 190,000 +/- 45,000 molecules were bound per cell, whereas unstimulated platelets bound 37,000 +/- 10,500 molecules/cell with a Kd of 1.9 +/- 0.15 microM. Plasminogen binding was inhibited in a dose-dependent manner by omega-aminocarboxylic acids at concentrations consistent with a requirement for an unoccupied high affinity lysine-binding site for plasminogen binding to the cells. When platelet-bound plasminogen was incubated with tissue plasminogen activator, urokinase, or streptokinase, gel analysis established that plasmin was preferentially associated with the platelet relative to the supernatant. Plasminogen and plasmin interacted with thrombin-stimulated platelets with similar binding characteristics, and there was no evidence for a binding site for plasmin which did not also bind plasminogen. Therefore, the results suggest that plasminogen activation is enhanced on the cell surface. In sum, these results indicate that platelets bind plasminogen at physiologic zymogen concentrations and this interaction may serve to localize and promote plasminogen activation.  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