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.
We have shown that the urokinase (UK) kringle domain contains a high-affinity plasminogen activator inhibitor-1 (PAI-1) binding site, responsible for the 10-fold faster complex formation between UK and PAI-1 than between PAI-1 and low-molecular-weight urokinase (LMWUK). Complex formation between UK and PAI-1, but not between LMWUK and PAI-1, was suppressed 10-fold in the presence of peptide U-107 derived from the UK kringle domain. Peptide U-373 derived from the UK catalytic domain slowed complex formation between UK and PAI-1 and also LMWUK and PAI-1. Inactivation of tissue-type plasminogen activator (tPA) by PAI-1 was slowed 10-fold in the presence of peptides derived from the tPA finger and kringle-2 domains. DFP-inactivated (DIP) UK and both forms of DIP-tPA inhibited PAI-1 binding to U-107 and to U-373 whereas single-chain urokinase-type PA (scuPA) was unable to compete with either peptide for PAI-1 binding. These data suggest that the reversible PAI-1 binding site in the UK A-chain plays a role in the rapid association with PAI-1 as important as those that reside in the tPA A-chain and that reversible PAI-1 binding sites are expressed on the surface of UK upon conversion from scuPA, in contrast to tPA.  相似文献   

2.
The urokinase plasminogen activator (uPA) and tissue plasminogen activator (tPA) are very similar serine proteases with the same physiological function, the activation of plasminogen. An increased amount or activity of uPA but not tPA has been detected in human cancers. The PAs are weak proteolytic enzymes, but they activate plasminogen to plasmin, a strong proteolytic enzyme largely responsible for the malignant properties of cancers. It has been shown recently that the administration of uPA inhibitors can reduce tumor size. Inhibitors of uPA could therefore be used as anti-cancer and anti-angiogenesis agents. It has been found that amiloride competitively inhibits the catalytic activity of uPA but not tPA. Modification of this chemical could therefore produce a new class of uPA specific inhibitors and a new class of anti-cancer agents. The X-ray structure of the uPA complex with amiloride is not known. There are structural differences in the specificity pocket of uPA and tPA. However, the potential energy of binding amiloride is lower outside this cavity in the case of tPA. A region responsible for binding amiloride to tPA has been proposed as the loop B93-B101, reached in negatively charged amino acids present in tPA but not uPA.  相似文献   

3.
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   

4.
G Pohl  H J?rnvall  P Kok  P Wallén 《FEBS letters》1986,205(1):92-96
Tissue plasminogen activator was purified in high yield from pig heart by immunoaffinity chromatography and characterized by analysis of the glycosylation pattern and the N-terminal amino acid sequence. Comparisons with the human enzyme reveals residue exchanges in the A-chain at positions 3 (porcine Arg/human Gln) and 5 (Thr/Ile), and in the B-chain at positions 6 (Tyr/Phe), 10 (Thr/Ala) and 20 (Val/Ala). The glycosylation pattern for the porcine activator was determined by endoglycosidase treatment followed by gel electrophoresis. The A-chain contains a single high-mannose type of N-linked glycan structure and the B-chain contains a complex type of oligosaccharide. A similar but not identical pattern has been observed for the human activator, purified from melanoma cells.  相似文献   

5.
Interaction of plasminogen and fibrin in plasminogen activation   总被引:2,自引:0,他引:2  
Glu1-, Lys77-, miniplasminogens, kringle 1-3, kringle 1-5A, and kringle 1-5R were able to bind with fibrin, while microplasminogen and kringle 4 did not bind significantly. Kringle 1-5A, but not kringle 1-3, effectively inhibited the binding of Glu1-, Lys77-, and miniplasminogens with fibrin. Miniplasminogen also inhibited the binding of Glu1-plasminogen with fibrin. The binding of kringle 1-3 with fibrin was blocked by mini- or Glu1-plasminogen. It is therefore evident that there are two fibrin-binding domains in plasminogen and that the one in kringle 5 is of higher affinity than that in kringle 1-3. CNBr cleavage products of fibrinogen effectively enhanced the activation of Glu1-, Lys77-, or miniplasminogens, but not microplasminogen, by tissue-type plasminogen activator. Kringle 1-5, but not kringle 1-3, dose-dependently inhibited the enhancement by fibrinogen degradation products of Glu1-plasminogen activation by the activator. Lysine and epsilon-aminocaproic acid could inhibit the binding of plasminogens and plasminogen derivatives with fibrin and block the enhancement effect of fibrinogen degradation products on plasminogen activation. The data clearly illustrate that the binding of plasminogen with fibrin, mainly determined by kringle 5, is essential for effective activation by tissue-type plasminogen activator. However, the presence of kringle 1-4 in the plasminogen molecule is required for the full enhancing effect since the kcat/Km of miniplasminogen activation in the presence of fibrinogen degradation products was 8.2 microM-1 min-1 which is significantly less than 52.0 microM-1 min-1 of Glu1-plasminogen.  相似文献   

6.
Human vascular smooth muscle cells (VSMC) bind tissue plasminogen activator (tPA) specifically, saturably, and with relatively high affinity (K(d) 25 nM), and this binding potentiates the activation of cell-associated plasminogen (Ellis, V., and Whawell, S. A. (1997) Blood 90, 2312-2322). We have observed that this binding can be efficiently competed by DFP-inactivated tPA and S478A-tPA but not by tPA inactivated with H-D-Phe-Pro-Arg-chloromethyl ketone (PPACK). VSMC-bound tPA also exhibited a markedly reduced inhibition by PPACK, displaying biphasic kinetics with second-order rate constants of 7. 5 x 10(3) M(-1) s(-1) and 0.48 x 10(3) M(-1) s(-1), compared with 7. 2 x 10(3) M(-1) s(-1) in the solution phase. By contrast, tPA binding to fibrin was competed equally well by all forms of tPA, and its inhibition was unaltered. These effects were shown to extend to the physiological tPA inhibitor, plasminogen activator inhibitor 1. tPA.plasminogen activator inhibitor 1 complex did not compete tPA binding to VSMC, and the inhibition of bound tPA was reduced by 30-fold. The behavior of the various forms of tPA bound to VSMC correlated with conformational changes in tPA detected by CD spectroscopy. These data suggest that tPA binds to its specific high affinity site on VSMC by a novel mechanism involving the serine protease domain of tPA and distinct from its binding to fibrin. Furthermore, reciprocally linked conformational changes in tPA appear to have functionally significant effects on both the interaction of tPA with its VSMC binding site and the susceptibility of bound tPA to inhibition.  相似文献   

7.
Seven murine monoclonal antibodies produced against tissue plasminogen activator (tPA) were evaluated by means of enzyme-linked immunosorbent assays (ELISAs), and their effects on the enzymatic activities of tPA towards a synthetic substrate (S-2288) and plasminogen were investigated. One of the antibodies, TPA1-70, strongly inhibited the enzymatic activity of tPA in a fibrin agarose plate assay, while it did not affect the enzymatic activity towards the synthetic substrate or plasminogen. The antibody is directed to an epitope on the B-chain of tPA, which is necessary for the formation of a ternary complex of tPA, fibrin and plasminogen, but probably not to the active site. Another antibody, TPA2-14, partially inhibited the enzymatic activities of tPA towards the synthetic substrate and plasminogen, but it was not able to bind to the inactive tPA complexed with plasminogen activator inhibitor-1 (PAI-1). The antibody is directed to an epitope on the second kringle region, which is probably one of the PAI-1 binding sites. This property of the antibody enabled us to develop an ELISA for selective quantitation of free tPA in culture media conditioned with several human cell lines. The results indicate that tPA in these media exists either partially or almost entirely in a complex with PAI-1.  相似文献   

8.
The crystal structure of the fibrinolytic enzyme tissue plasminogen activator (tPA) shows that the bulky side chain of Y99 hinders access to the active site by partially occluding the S2 site and may be responsible for the low catalytic activity of tPA toward plasminogen. We have tested the role of Y99 by replacing it with Leu, the residue found in more proficient proteases like trypsin and thrombin. The Y99L replacement results in an increase in the k(cat)/Km for chromogenic substrates due to enhanced diffusion into the active site. The increase is modest (threefold) for substrates specific for tPA that carry Pro or Gly at P2, but reaches 80-fold for less specific substrates carrying Arg at P2. On the other hand, the Y99L mutation has no effect on the activity of tPA toward the natural substrate plasminogen, that carries Gly at P2, and reduces more than 10-fold the inhibition of tPA by plasminogen activator inhibitor-1 (PAI-1), that carries Ala at P2. We conclude that the steric hindrance provided by Y99 in the crystal structure affects mostly nonphysiological substrates with bulky residues at P2. In addition, residue Y99 plays an active role in the recognition of PAI-1, but not plasminogen. Mutations of Y99 could therefore afford a resistance to inhibition by PAI-1 without compromising the fibrinolytic potency of tPA, a result of potential therapeutic relevance.  相似文献   

9.
Urokinase-type plasminogen activator (uPA) plays a central role in tissue remodeling processes. Most of our understanding of the role of uPA in vivo is derived from studies using gene-targeted uPA-deficient mice. To enable in vivo studies on the specific interference with uPA functionality in mouse models, we have now developed murine monoclonal antibodies (mAbs) directed against murine uPA by immunization of uPA-deficient mice with the recombinant protein. Guided by enzyme-linked immunosorbent assay, Western blotting, surface plasmon resonance, and enzyme kinetic analyses, we have selected two highly potent and inhibitory anti-uPA mAbs (mU1 and mU3). Both mAbs recognize epitopes located on the B-chain of uPA that encompasses the catalytic site. In enzyme activity assays in vitro, mU1 blocked uPA-catalyzed plasminogen activation as well as plasmin-mediated pro-uPA activation, whereas mU3 only was directed against the first of these reactions. We additionally provide evidence that mU1, but not mU3, successfully targets uPA-dependent processes in vivo. Hence, systemic administration of mU1 (i) rescued mice treated with a uPA-activable anthrax protoxin and (ii) impaired uPA-mediated hepatic fibrinolysis in tissue-type plasminogen activator (tPA)-deficient mice, resulting in a phenotype mimicking that of uPA;tPA double deficient mice. Importantly, this is the first report demonstrating specific antagonist-directed targeting of mouse uPA at the enzyme activity level in a normal physiological process in vivo.  相似文献   

10.
The increased levels of extracellular DNA found in a number of disorders involving dysregulation of the fibrinolytic system may affect interactions between fibrinolytic enzymes and inhibitors. Double-stranded (ds) DNA and oligonucleotides bind tissue-(tPA) and urokinase (uPA)-type plasminogen activators, plasmin, and plasminogen with submicromolar affinity. The binding of enzymes to DNA was detected by EMSA, steady-state, and stopped-flow fluorimetry. The interaction of dsDNA/oligonucleotides with tPA and uPA includes a fast bimolecular step, followed by two monomolecular steps, likely indicating slow conformational changes in the enzyme. DNA (0.1-5.0 μg/ml), but not RNA, potentiates the activation of Glu- and Lys-plasminogen by tPA and uPA by 480- and 70-fold and 10.7- and 17-fold, respectively, via a template mechanism similar to that known for fibrin. However, unlike fibrin, dsDNA/oligonucleotides moderately affect the reaction between plasmin and α(2)-antiplasmin and accelerate the inactivation of tPA and two chain uPA by plasminogen activator inhibitor-1 (PAI-1), which is potentiated by vitronectin. dsDNA (0.1-1.0 μg/ml) does not affect the rate of fibrinolysis by plasmin but increases by 4-5-fold the rate of fibrinolysis by Glu-plasminogen/plasminogen activator. The presence of α(2)-antiplasmin abolishes the potentiation of fibrinolysis by dsDNA. At higher concentrations (1.0-20 μg/ml), dsDNA competes for plasmin with fibrin and decreases the rate of fibrinolysis. dsDNA/oligonucleotides incorporated into a fibrin film also inhibit fibrinolysis. Thus, extracellular DNA at physiological concentrations may potentiate fibrinolysis by stimulating fibrin-independent plasminogen activation. Conversely, DNA could inhibit fibrinolysis by increasing the susceptibility of fibrinolytic enzymes to serpins.  相似文献   

11.
The amidolytic plasmin activity of a mixture of tissue plasminogen activator (tPA) and plasminogen is enhanced by heparin at therapeutic concentrations. Heparin also increases the activity in mixtures of urokinase-type plasminogen activator (uPA) and plasminogen but has no effect on streptokinase or plasmin. Direct analyses of plasminogen activation by polyacrylamide gel electrophoresis demonstrate that heparin increases the activation of plasminogen by both tPA and uPA. Binding studies show that heparin binds to various components of the fibrinolytic system, with tight binding demonstrable with tPA, uPA, and Lys-plasminogen. The stimulation of tPA activity by fibrin, however, is diminished by heparin. The ability of heparin to promote plasmin generation is destroyed by incubation of the heparin with heparinase, whereas incubation with chondroitinase ABC or AC has no effect. Also, stimulation of plasmin formation is not observed with dextran sulfate or chondroitin sulfate A, B, or C. Analyses of heparin fractions after separation on columns of antithrombin III-Sepharose suggest that both the high-affinity and the low-affinity fractions, which have dramatically different anticoagulant activity, have similar activity toward the fibrinolytic components.  相似文献   

12.
Tissue-type plasminogen activator (tPA) regulates fibrin clot lysis by stimulating the conversion of plasminogen into the active protease plasmin. Fibrin is required for efficient tPA-mediated plasmin generation and thereby stimulates its own proteolysis. Several fibrin regions can bind to tPA, but the structural basis for this interaction is unknown. Amyloid beta (Abeta) is a peptide aggregate that is associated with neurotoxicity in brains afflicted with Alzheimer's disease. Like fibrin, it stimulates tPA-mediated plasmin formation. Intermolecular stacking of peptide backbones in beta sheet conformation underlies cross-beta structure in amyloid peptides. We show here that fibrin-derived peptides adopt cross-beta structure and form amyloid fibers. This correlates with tPA binding and stimulation of tPA-mediated plasminogen activation. Prototype amyloid peptides, including Abeta and islet amyloid polypeptide (IAPP) (associated with pancreatic beta cell toxicity in type II diabetes), have no sequence similarity to the fibrin peptides but also bind to tPA and can substitute for fibrin in plasminogen activation by tPA. Moreover, the induction of cross-beta structure in an otherwise globular protein (endostatin) endows it with tPA-activating potential. Our results classify tPA as a multiligand receptor and show that cross-beta structure is the common denominator in tPA binding ligands.  相似文献   

13.
1. Possible interactions between fibrin(ogen) and heparin in the control of plasminogen activation were studied in model systems using the thrombolytic agents tissue-type plasminogen activator (t-PA), urokinase and streptokinase.plasminogen activator complex and the substrates Glu- and Lys-plasminogen. 2. Both t-PA and urokinase activities were promoted by heparin and by pentosan polysulphate, but not by chondroitin sulphate or hyaluronic acid. The effect was on Km. 3. In the presence of soluble fibrin (and its mimic, CNBr-digested fibrinogen) the effect of heparin on t-PA was attenuated, although not abolished. In studies using a monoclonal antibody and 6-aminohexanoic acid, it was found that heparin and fibrin did not seem to share a binding site on t-PA. 4. The activity of t-PA B-chain was unaffected by heparin, so the binding site is located on the A-chain of t-PA (and urokinase). 5. Fibrin potentiated the activity of heparin on urokinase. The activity of streptokinase.plasminogen was unaffected by heparin whether or not fibrin was present. 6. If these influences of heparin and fibrin also occur in vivo, then, in the presence of heparin, the relative fibrin enhancement of t-PA will be diminished and the likelihood of systemic activation by t-PA is increased.  相似文献   

14.
A new spectrophotometric solid-phase fibrin-tissue plasminogen activator activity assay (SOFIA-tPA), specific for the quantitation of tissue plasminogen activators, is described. The method is based on (1) the high-affinity binding (Kp = 1.4 +/- 2 nM) of tPA to a solid-phase fibrin network constructed by thrombin proteolysis of fibrinogen covalently coupled to polyglutaraldehyde-activated polyvinyl chloride microtiter plates, and (2) the subsequent development of PA activity by the fibrin-tPA complex and its measurement with a coupled assay using a chromogenic substrate highly selective for plasmin. Conditions were chosen such that the rate of para-nitroaniline release from the substrate is directly proportional to the concentration of tPA. The support is able to isolate tPA from the bulk of proteins present in any biological fluid allowing the assay to specifically detect tPA activity (range: 0.01 to 50 IU/ml) even in the presence of other activators, proteases, and inhibitors. Since the assay is done in a well-defined reaction mixture (the fibrin-tPA complex, plasminogen, and the synthetic substrate), kinetics studies using pure or crude tPA can be performed. Standard curves (rate measurement and endpoint methods) were made using the international standard (preparation 83/517) for tPA.  相似文献   

15.
We have demonstrated that tissue plasminogen activator (tPA) binds specifically to human vascular smooth muscle cells (VSMC) in a functionally relevant manner, both increasing plasminogen activation and decreasing tPA inhibition (Ellis, V., and Whawell, S. A. (1997) Blood 90, 2312-2322; Werner, F., Razzaq, T. M., and Ellis, V. (1999) J. Biol. Chem. 274, 21555-21561). To further understand this system we have now identified and characterized the protein responsible for this binding. Rat VSMC were surface-labeled with 125I, and cell lysates were subjected to an affinity chromatography scheme based on the previously identified tPA binding characteristics. A single radiolabeled protein of 63 kDa bound specifically and was eluted at low pH. This protein was isolated from large scale preparations of VSMC and unambiguously identified as the rat homologue of the human type-II transmembrane protein p63 (CKAP4) by matrix-assisted laser desorption ionization and nano-electrospray tandem mass spectrometry of tryptic fragments. In confirmation of this, a monoclonal antibody raised against authentic human p63 recognized the isolated protein in Western blotting. Immunofluorescence microscopy demonstrated that p63 was located principally in the endoplasmic reticulum but was also detected in significant quantities on the surface of human VSMC. In support of the hypothesis that p63 is the functional tPA binding site on VSMC, an anti-p63 monoclonal antibody was found to block tPA binding. Furthermore, heterologous expression of an N-terminally truncated mutant of p63, which targets exclusively to the plasma membrane, led to an increase in tPA-catalyzed plasminogen activation. Therefore, p63 on the surface of VSMC may contribute to the functional regulation of the plasminogen activation system in the vessel wall.  相似文献   

16.
W E Van Nostrand  M Porter 《Biochemistry》1999,38(35):11570-11576
Cerebrovascular amyloid beta-protein (A beta) deposition, a key pathological feature of Alzheimer's disease and hereditary cerebral hemorrhage with amyloidosis Dutch-type, can lead to intracerebral hemorrhage; however, the mechanism for this remains unclear. Assembled A beta is a potent stimulator of tissue-type plasminogen activator (tPA) in vitro. Herein, we investigated the stimulation of tPA by freshly solubilized A beta 1-40. The rate of tPA stimulation by A beta 1-40 increased dramatically over time, suggesting that A beta may be altered during the course of the reaction. SDS-PAGE analysis showed that A beta 1-40 was cleaved during the course of the reaction. Subsequent studies showed that it was plasmin, the product of tPA activation of plasminogen, that specifically cleaved A beta 1-40 in the amino terminal region between Arg5 and His6. Plasmin effectively cleaved a chromogenic substrate corresponding to this cleavage site in A beta. Circular dichroism spectral analysis showed that A beta 6-40 adopted a strong beta-sheet secondary structure. This truncated A beta 6-40 peptide was a potent stimulator of tPA in vitro. Our results indicate that beta-sheet secondary structure of A beta, which can be promoted by plasmin cleavage, stimulates tPA activity. These findings suggest that pathologic interactions between A beta, tPA, and plasmin in the cerebral vessel wall could result in excessive proteolysis contributing to intracerebral hemorrhages.  相似文献   

17.
Tissue plasminogen activator (tPA) is necessary for hippocampal long-term potentiation. Synaptically released zinc also contributes to long-term potentiation, especially in the hippocampal CA3 region. Using cortical cultures, we examined whether zinc increased the concentration and/or activity of tPA. Two hours after a 10-min exposure to 300 μM zinc, expression of tPA and its substrate, plasminogen, were significantly increased, as was the proteolytic activity of tPA. In contrast, increasing extracellular or intracellular calcium levels did not affect the expression or secretion of tPA. Changing zinc influx or chelating intracellular zinc also failed to alter tPA/plasminogen induction by zinc, indicating that zinc acts extracellularly. Zinc-mediated extracellular activation of matrix metalloproteinase (MMP) underlies the up-regulation of brain-derived neurotrophic factor (BDNF) and tropomyosin receptor kinase (Trk) signaling. Consistent with these findings, co-treatment with a neutralizing antibody against BDNF or specific inhibitors of MMPs or Trk largely reversed tPA/plasminogen induction by zinc. Treatment of cortical cultures with p-aminophenylmercuric acetate, an MMP activator, MMP-2, or BDNF alone induced tPA/plasminogen expression. BDNF mRNA and protein expression was also increased by zinc and mediated by MMPs. Thus, an extracellular zinc-dependent, MMP- and BDNF-mediated synaptic mechanism may regulate the levels and activity of tPA.  相似文献   

18.
The vampire bat salivary plasminogen activator (BatPA) is virtually inactive toward Glu-plasminogen in the absence of a fibrin-like cofactor, unlike human tissue-type plasminogen activator (tPA) (the kcat/Km values were 4 and 470 M-1 s-1, respectively). In the presence of fibrin II, tPA and BatPA activated Glu-plasminogen with comparable catalytic efficiencies (158,000 and 174,000 M-1 s-1, respectively). BatPA's cofactor requirement was partially satisfied by polymeric fibrin I (54,000 M-1 s-1), but monomeric fibrin I was virtually ineffective (970 M-1 s-1). By comparison, a variety of monomeric and polymeric fibrin-like species markedly enhanced tPA-mediated activation of Glu-plasminogen. Fragment X polymer was 2-fold better but 9-fold worse as cofactor for tPA and BatPA, respectively, relative to fibrin II. Fibrinogen, devoid of plasminogen, was a 10-fold better cofactor for tPA than fibrinogen rigorously depleted of plasminogen, Factor XIII, and fibronectin; the enhanced stimulatory effect of the less-purified fibrinogen was apparently due to the presence of Factor XIII. By contrast, the two fibrinogen preparations were equally poor cofactors of BatPA-mediated activation of Glu-plasminogen. BatPA possessed only 23 and 4% of the catalytic efficiencies of tPA and two-chain tPA, respectively, in hydrolyzing the chromogenic substrate Spectrozyme tPA. However in the presence of fibrin II, BatPA and tPA exhibited similar kcat/Km values for the hydrolysis of Spectrozyme tPA. Our data revealed that BatPA, unlike tPA, displayed a strict and fastidious requirement for polymeric fibrin I or II. Consequently, BatPA may preferentially promote plasmin generation during a narrow temporal window of fibrin formation and dissolution.  相似文献   

19.
Human single-chain tissue-type plasminogen activator (tPA) was activated by the culture media of mouse B16 melanoma and Lewis lung carcinoma cells. This activation was due to the proteolytic conversion of single-chain tPA to two-chain tPA. Typical serine proteinase inhibitors, such as diisopropylfluorophosphate and aprotinin, strongly inhibited the proteolytic activation, suggesting that the enyzme responsible for this activation is a serine proteinase. Through a process of gel filtration, the molecular weight of the putative tPA-activating enzyme was estimated to be approximately 35 kDa. Expression of the tPA mRNA was demonstrated by Northern blot analysis both for the melanoma and carcinoma cells. Zymographic experiments showed that the culture medium of the melanoma cells contained active two-chain tPA. These results indicate that a common serine proteinase may be involved in the proteolytic activation of single-chain tPA in these cancer cells.  相似文献   

20.
Pituitary adenylate cyclase-activating polypeptide (PACAP) and tissue plasminogen activator (tPA) play important roles in neuronal migration and survival. However, a direct link between the neurotrophic effects of PACAP and tPA has never been investigated. In this study, we show that, in PC12 cells, PACAP induced a 9.85-fold increase in tPA gene expression through activation of the protein kinase A- and protein kinase C-dependent signaling pathways. In immature cerebellar granule neurons (CGN), PACAP stimulated tPA mRNA expression and release of proteolytically active tPA. Immunocytochemical labeling revealed the presence of tPA in the cytoplasm and processes of cultured CGN. The inhibitory effect of PACAP on CGN motility was not affected by the tPA substrate plasminogen or the tPA inhibitor plasminogen activator inhibitor-1. In contrast, plasminogen activator inhibitor-1 significantly reduced the stimulatory effect of PACAP on CGN survival. Altogether, these data indicate that tPA gene expression is activated by PACAP in both tumoral and normal neuronal cells. The present study also demonstrates that PACAP stimulates the release of tPA which promotes CGN survival by a mechanism dependent of its proteolytic activity.  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