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 234 毫秒
1.
Vitronectin, found in the extracellular matrix and in circulating blood, has an important role in the control of plasminogen activation. It was shown to be the major protein substrate in human blood fluid for a protein kinase A (PKA) released from platelets upon their physiological stimulation with thrombin. Since vitronectin was shown to have only one PKA phosphorylation site, but to contain 2-3 mol covalently bound phosphate, it was reasonable to assume that other protein kinases might phosphorylate vitronectin at other sites in the protein. We have reported earlier that human serum contains at least three protein kinases, one of which was found to be cAMP independent and to phosphorylate a repertoire of plasma proteins that was very similar to that obtained upon phosphorylation of human plasma with protein kinase C (PKC). Since there are now several examples of proteins with extracellular functions that are phosphorylated by PKC, we undertook to study the phosphorylation of vitronectin by PKC. Here, we show that vitronectin is a substrate for PKC, and characterize the kinetic parameters of this phosphorylation (Km approximately tenfold lower than the concentration of vitronectin in blood), indicating that, from the biochemical point of view, this phosphorylation can occur at the locus of a hemostatic event. We also identify Ser362 as the major PKC phosphorylation site in vitronectin, and confirm this localization by means of synthetic peptides derived from the cluster of basic amino acids in vitronectin surrounding Ser362. We show that the PKC phosphorylation at Ser362 alters the functional properties of vitronectin, attenuating its cleavage by plasmin at Arg361-Ser362. This phosphorylation has the potential to regulate plasmin production from plasminogen by a feedback mechanism involving the above-mentioned plasmin cleavage, a loosening of the vitronectin grip on inhibitor 1 of plasminogen activators, and a subsequent latency of this regulatory inhibitor.  相似文献   

2.
During wound healing, migrating cells increase expression of both the vitronectin receptor (VNR) integrins and plasminogen activators. Here we report that vitronectin significantly enhances the migration of smooth muscle cells (SMCs), and that the specific VNR alpha V beta 3 is required for cell motility. We also show that the alpha V beta 3 attachment site on vitronectin overlaps with the binding site for plasminogen activator inhibitor (PAI)-1, and that the active conformation of PAI-1 blocks SMC migration. This effect requires high-affinity binding to vitronectin, and is not dependent on the ability of PAI-1 to inhibit plasminogen activators. Formation of a complex between PAI-1 and plasminogen activators results in loss of PAI-1 affinity for vitronectin and restores cell migration. These data demonstrate a direct link between plasminogen activators and integrin-mediated cell migration, and show that PAI-1 can control cell-matrix interactions by regulating the accessibility of specific cell-attachment sites. This indicates that the localization of plasminogen activators at sites of focal contact does not initiate a proteolytic cascade leading to generalized matrix destruction, but instead is required to expose cryptic cell-attachment sites necessary for SMC migration.  相似文献   

3.
Plasminogen activator inhibitor type 1 (PAI-1) is the rapid physiologic inhibitor of tissue-type plasminogen activator and urokinase-type plasminogen activator (uPA). In plasma and the extracellular matrix, PAI-1 is associated with the adhesive glycoprotein vitronectin. In order to characterize the PAI-1 structural domain responsible for binding to vitronectin, the segment of the PAI-1 cDNA encoding amino acids 13-147 (nucleotides 248-650) was randomly mutagenized and subcloned into a bacterial expression vector containing the mature PAI-1 coding sequence. Recombinant PAI-1 mutants were expressed in Escherichia coli and bacterial lysates assayed in duplicate for uPA inhibitory activity and vitronectin binding. Of 190 clones screened, six consistently demonstrated decreased vitronectin binding relative to uPA inhibitory activity. DNA sequence analysis of four of these clones identified 10 unique missense mutations, all located between base pairs 298 and 641, with each clone containing between one and four substitutions. Each substitution was expressed independently by site-directed mutagenesis and again analyzed for uPA inhibitory activity and vitronectin binding. Five point mutations that selectively disrupt vitronectin binding were identified. All 5 residues are located on the exterior of the PAI-1 structure. These findings appear to define a complex binding surface that bridges alpha-helices C and E to beta-strand 1A and includes amino acids 55, 109, 110, 116, and 123. These results suggest that vitronectin binding may stabilize the active conformation of PAI-1 by restricting the movement of beta-sheet A and thereby preventing insertion of the reactive center loop.  相似文献   

4.
Active PAI-1 (plasminogen activator inhibitor 1) is bound to vitronectin in plasma and in the extracellular matrix. In this study we aimed at identifying the PAI-1 binding site in vitronectin, which at present is a matter of dispute. Vitronectin was cleaved with trypsin and the fragments were tested for inhibitory effect on the PAI-1/vitronectin interaction using vitronectin-coated microtiter plates. Intact vitronectin and the tryptic digest of vitronectin both caused a 50% reduction in PAI-1 binding at a concentration of about 2 nmol/I. Gel-filtration on Sephadex G-50 superfine of the tryptic peptides resulted in one main peak of inhibitory activity. The elution volume, Kav, was 0.55 indicating (a) medium-size peptide(s). The peptide was further purified by reverse-phase HPLC. Structural analysis revealed that it constituted the 45 NH2-terminal amino-acid residues in vitronectin. The NH2-terminal vitronectin peptide caused a 50% decrease in PAI-1 binding to the vitronectin-coated microtiter plates at a concentration of about 13 nmol/l. Thus, the peptide is a little less effective in this respect than intact vitronectin. Reduced and S-carboxymethylated peptide had no effect on the interaction. The NH2-terminal vitronectin fragment increased the stability of active PAI-1 by about 60%, which is a little less than with intact vitronectin. The peptide also prevented PAI-1 from oxidation with chloramine T. The half-life was prolonged about 4-fold as compared to about 30-fold with intact vitronectin.  相似文献   

5.
Plasminogen activator inhibitor-1 (PAI-1), the primary inhibitor of tissue-type plasminogen activator and urokinase, is known to convert readily to a latent form by insertion of the reactive center loop into a central beta-sheet. Interaction with vitronectin stabilizes PAI-1 and decreases the rate of conversion to the latent form, but conformational effects of vitronectin on the reactive center loop of PAI-1 have not been documented. Mutant forms of PAI-1 were designed with a cysteine substitution at either position P1' or P9 of the reactive center loop. Labeling of the unique cysteine with a sulfhydryl-reactive fluorophore provides a probe that is sensitive to vitronectin binding. Results indicate that the scissile P1-P1' bond of PAI-1 is more solvent exposed upon interaction with vitronectin, whereas the N-terminal portion of the reactive loop does not experience a significant change in its environment. These results were complemented by labeling vitronectin with an arginine-specific coumarin probe which compromises heparin binding but does not interfere with PAI-1 binding to the protein. Dissociation constants of approximately 100 nM are calculated for the vitronectin/PAI-1 interaction from titrations using both fluorescent probes. Furthermore, experiments in which PAI-1 failed to compete with heparin for binding to vitronectin argue for separate binding sites for the two ligands on vitronectin.  相似文献   

6.
The low density lipoprotein receptor-related protein (LRP), a multi-functional endocytic receptor, mediates the cellular internalization of tissue-type (t-PA) and urokinase-type (u-PA) plasminogen activator and their complexes with plasminogen activator inhibitor type 1 (PAI-1). LRP preferentially binds the complexed forms, exemplified by equilibrium dissociation constants (KD) that are at least an order of magnitude lower than those of the free components. To understand the molecular interactions, underlying the preference of the receptor for complexes rather than for the free components, we have performed a detailed analysis of the affinity and kinetics of the binding of PAI-1 and t-PA:PAI-1 complexes to the receptor, using surface plasmon resonance. To assess the involvement of the heparin-binding domain of PAI-1 for the interaction with LRP, we determined the equilibrium dissociation constants for the binding to LRP of a panel of PAI-1 mutants with single- and multiple amino-acid substitutions of the basic residues that constitute the heparin binding site of PAI-1 (K65, K69, R76, K80 and K88). The binding of these PAI-1 mutants was partially reduced with a 2 to 4 fold increase in KD values for single (K80, K88) and combined (K80, 88) substitution mutant proteins respectively. LRP binding of complexes, composed of t-PA with either wild type PAI-1 or any one of the single PAI-1 mutants indicated a major role of lysine 69 (K69) for the binding of t-PA:PAI-1 complexes to LRP (KD values of 6.1, 3.7. 75.4, 5.4, 12.5 and 8.1 nM for wild type, K65A, K69A, R76A, K80A and K88A complexes, respectively). Since the KD for the binding of free t-PA to LRP is 158 nM, we conclude that the PAI-1 moiety harbors the major determinant for t-PA:PAI-1 complex binding to LRP. The in vitro binding studies were extended by binding and clearance studies with COS-1 cells. Degradation of both 125I-t-PA:PAI-1 K69A and 125I-t-PA:PAI-1 K69A K80A K88A complexes after 2 h of incubation was reduced compared to the degradation of 125I-t-PA:PAI-1 complexes. We conclude that PAI-1 contains a cryptic binding site (lysine 69) for LRP, that is specifically expressed upon t-PA:PAI-1 complex formation.  相似文献   

7.
We have analysed the susceptibility of latent, active, reactive-centre-cleaved and plasminogen-activator-complexed type-1 plasminogen-activator inhibitor (PAI-1) to the non-target proteinases trypsin, endoproteinase Asp-N, proteinase K and subtilisin. This analysis has allowed us to detect conformational differences between the different forms of PAI-1 outside the reactive-centre loop and beta-sheet A. Proteinase-hypersensitive sites were clustered in three regions. Firstly, susceptibility was observed in the region around alpha-helix E, beta-strand 1A, and the flanking loops, which are believed to form flexible joints during movements of beta-sheet A. Secondly, hypersensitive sites were observed in the loop between alpha-helix I and beta-strand 5A. Thirdly, the gate region, encompassing beta-strands 3C and 4C, was highly susceptible to trypsin in latent PAI-1, but not in the other conformations. The digestion patterns differed among all four forms of PAI-1, indicating that each represents a unique conformation. The differential proteolytic susceptibility of the flexible-joint region may be coupled to the differential affinity to vitronectin, binding in the same region. The analysis also allowed detection of conformational differences between reactive-centre-cleaved forms produced under different solvent conditions. The digestion pattern of plasminogen-activator-complexed PAI-1 was different from that of active PAI-1, but indistinguishable from that of one of the reactive-centre-cleaved forms, as the complexed and this particular cleaved PAI-1 were completely resistant to all the non-target proteinases tested. This observation is in agreement with the notion that complex formation involves reactive-centre cleavage and a large degree of insertion of the reactive-centre loop into beta-sheet A. Our analysis has allowed the identification of some flexible regions that appear to be implicated in the conformational changes during the movements of beta-sheet A and during the inhibitory reaction of serpins with their target proteinases.  相似文献   

8.
Induction of the urokinase type plasminogen activator receptor (uPAR) promotes cell adhesion through its interaction with vitronectin (VN) in the extracellular matrix, and facilitates cell migration and invasion by localizing uPA to the cell surface. We provide evidence that this balance between cell adhesion and cell detachment is governed by PA inhibitor-1 (PAI-1). First, we demonstrate that uPAR and PAI-1 bind to the same site in VN (i.e., the amino-terminal somatomedin B domain; SMB), and that PAI-1 competes with uPAR for binding to SMB. Domain swapping and mutagenesis studies indicate that the uPAR-binding sequence is located within the central region of the SMB domain, a region previously shown to contain the PAI-1-binding motif. Second, we show that PAI-1 dissociates bound VN from uPAR and detaches U937 cells from their VN substratum. This PAI-1 mediated release of cells from VN appears to occur independently of its ability to function as a protease inhibitor, and may help to explain why high PAI-1 levels indicate a poor prognosis for many cancers. Finally, we show that uPA can rapidly reverse this effect of PAI-1. Taken together, these results suggest a dynamic regulatory role for PAI-1 and uPA in uPAR-mediated cell adhesion and release.  相似文献   

9.
Very-low-density lipoprotein receptor (VLDLR) and alpha2-macroglobulin receptor/low-density-lipoprotein-receptor-related protein (alpha2MR/LRP) are multifunctional endocytosis receptors of the low-density lipoprotein receptor family. Both have been shown to mediate endocytosis and degradation of complex between plasminogen activators and type-1 plasminogen-activator inhibitor (PAI-1) by cultured cells. We have now studied the specificity of binding and endocytosis by VLDLR and alpha2MR/LRP among a variety of serine proteinase/serpin complexes, including various combinations of the serine proteinases urokinase-type and tissue-type plasminogen activators, plasmin, thrombin, human leukocyte elastase, cathepsin G, and plasma kallikrein with the serpins PAI-1, horse leukocyte elastase inhibitor, protein C inhibitor, C1-inhibitor, alpha2-antiplasmin, alpha1-proteinase inhibitor, alpha1-antichymotrypsin, protease nexin-1, heparin cofactor II, and antithrombin III. Binding was estimated with radiolabelled ligands in ligand blotting analysis and microtiter well assays. Endocytosis was estimated by measuring receptor-associated protein (RAP)-sensitive degradation of radiolabelled complexes by Chinese hamster ovary cells transfected with VLDLR cDNA and by COS-1 cells, which have a high endogenous expression of alpha2MR/LRP. We found that the receptors bind with high affinity to some, but not all, combinations of plasminogen activators and thrombin with PAI-1, protease nexin-1, protein C inhibitor, and antithrombin III, while complexes of many serine proteinases with their primary inhibitor, i.e. plasmin/alpha2-antiplasmin complex, do not bind, or bind with a very low affinity. Both the serine proteinase and the serpin moieties contribute to the binding specificity. The binding specificities of VLDLR and alpha2MR/LRP are overlapping, but not identical. The results suggest that VLDLR and alpha2MR/LRP have different biological functions by having different binding specificities as well as by being expressed by different cell types.  相似文献   

10.
The multifunctionality of adhesion receptor ligands as well as the promiscuous nature of vascular integrins and nonintegrin-dependent adhesive interactions allow ligand-receptor binding of variable strength. The cooperation with pericellular proteolysis cascades is required for vascular remodelling during angiogenesis, atherogenesis or inflammatory processes. In particular, integrin-dependent cell contact, spreading and (trans-)migration can be modulated by ECM-associated PAI-1 and uPA-receptor driven reactions that are intimately linked to the invasive potential of cells. Recently, mechanisms of molecular crosstalk between these receptor systems have been recognized: (a) uPA-receptor may directly interact with beta 1- and beta 2-integrins on circulating blood cells; (b) av beta 3-integrin-directly binds to a matrix metalloproteinase; (c) uPA and PAI-1 balance the high affinity binding of vitronectin to uPA-receptor; (d) vitronectin-dependent cell adhesion and migration involving alpha v-integrins or uPA-receptor are blocked by active PAI-1 independent of its role as protease inhibitor. These results are compatible with vascular injury studies in uPA- and PAI-1 knock-out mice and provide new targets for the treatment of diseases associated with imbalanced vascular remodelling.  相似文献   

11.
Plasminogen activator inhibitor-1 (PAI-1), an important risk factor for thrombotic diseases, is a member of the superfamily of serine proteinase inhibitors. To define structural rearrangements occurring during interaction between PAI-1 and its target proteinases we have raised monoclonal antibodies against the PAI-1/t-PA complex. Thirteen out of 401 monoclonal antibodies reacted preferentially with the PAI-1/t-PA complex as compared to free PAI-1 or free t-PA. Detailed characterization revealed the presence of two non-overlapping neoantigenic epitopes in the PAI-1/t-PA complex. Both neoantigenic epitopes were also exposed after complex formation between PAI-1 and either urokinase-type plasminogen activator, plasmin or thrombin as well as after cleavage of the reactive site loop of non-inhibitory substrate type PAI-1 variants. Thus, we have identified two neoantigenic epitopes, localized entirely in PAI-1, and commonly exposed after complex formation of active PAI-1 with various proteinases or after cleavage of substrate PAI-1. These monoclonal antibodies should facilitate further studies on the mechanism of interaction between various PAI-1 forms and its target proteinases.  相似文献   

12.
13.
Plasminogen-activator inhibitor type I (PAI-1), the primary inhibitor of urinary-type plasminogen activator, is thought to play an important role in the control of stroma invasion by both endothelial and tumor cells. Using an in vitro angiogenesis model of capillary extension through a preformed monolayer, in conjunction with in situ hybridization analysis, we showed that PAI-1 mRNA is specifically induced in cells juxtaposed next to elongating sprouts. To further establish that PAI-1 expression is induced as a consequence of a direct contact with endothelial cells, coculture experiments were performed. PAI-1 mRNA was induced exclusively in fibroblasts (L-cells) contacting endothelial cell (LE-II) colonies. Reporter gene constructs driven by a PAI-1 promoter and stably transfected into L-cells were used to establish that both mouse and rat PAI-1 promoters mediate apposition-dependent regulation. This mode of PAI-1 regulation is not mediated by plasmin, as an identical spatial pattern of expression was detected in cocultures treated with plasmin inhibitors. Because endothelial cells may establish direct contacts with fibroblasts only during angiogenesis, we propose that focal induction of PAI-1 at the site of heterotypic cell contacts provides a mechanism to negate excessive pericellular proteolysis associated with endothelial cell invasion.  相似文献   

14.
BACKGROUND: We examined the relationships among vitronectin (VN), plasminogen activator inhibitor-1 (PAI-1), and transforming growth factor beta 1 (TGF-beta 1) in liver diseases to evaluate the presence of plasmin cascade in human hepatic fibrosis. METHODS: Blood and liver tissues were obtained from 57 patients with liver disease. Plasma VN, PAI-1 antigen, and PAI-1 activity levels were evaluated. Biopsied liver specimens were observed by light and electron microscopy after immunohistochemical staining. Morphometric analysis was performed on these specimens. RESULTS: Plasma VN and PAI-1 activity levels decreased significantly with the progression of hepatic fibrosis and were particularly marked in the liver cirrhosis group. Plasma PAI-1 antigen level increased significantly. The immunolocalization of the active form of TGF-beta became more intense with the progression of hepatic fibrosis, whereas that of the dual-stained positive areas of PAI-1 and VN (PAI-1.VN) decreased. There was a positive correlation between TGF-beta and PAI-1, whereas there was a negative correlation between TGF-beta and PAI-1.VN. Immunoelectron microscopy showed the localization of PAI-1-VN in the extracellular space around the sinusoidal cells or surface of aggregating platelets, TGF-beta mainly in Ito cells, and VN in hepatocytes near the focal necrotic area or fibrous septa. CONCLUSIONS: These findings suggest that VN and PAI-1 are related to the active form of TGF-beta and that it is possible that the plasmin cascade is present in the human liver.  相似文献   

15.
Serpin inhibitors are believed to form an acyl enzyme intermediate with their target proteinases which is stabilized through insertion of the enzyme-linked part of the reactive center loop (RCL) as strand 4 in beta-sheet A of the inhibitor. To test critically the role and timing of these steps in the reaction of the plasminogen activator inhibitor PAI-1, we blocked the vacant position 4 in beta-sheet A of this serpin with an octapeptide. The peptide-blocked PAI-1 was a substrate for both tissue-type plasminogen activator (tPA) and trypsin and was hydrolyzed at the scissile bond. The reactivity of the peptide-blocked substrate PAI-1 was compared to that of the unmodified inhibitor by rapid acid quenching as well as photometric techniques. With trypsin as target, the limiting rate constants for enzyme acylation were essentially the same with inhibitor and substrate PAI-1 (21-23 s-1), as were also the associated apparent second-order rate constants (2.8-2.9 microM-1 s-1). With tPA, inhibitor and substrate PAI-1 reacted identically to form a tightly bound Michaelis complex (Kd approximately Km approximately 20 nM). The limiting rate constant for acylation of tPA, however, was 57 times faster with inhibitor PAI-1 (3.3 s-1) than with the substrate form (0.059 s-1), resulting in a 5-fold difference in the corresponding second-order rate constants (13 vs 2.5 microM-1 s-1). We attribute the ability of tPA to discriminate between the two PAI-1 forms to exosite bonds that cannot occur with trypsin. The exosite bonds retain specifically the distal part of the PAI-1 RCL in the substrate pocket, which favors a reversal of the acylation step. Acylation of tPA becomes effective only by separating the products of the acylation step. With substrate PAI-1, this depends on passive displacement of bonds, whereas with inhibitor PAI-1, separation is accomplished by loop insertion that pulls tPA from its docking site on PAI-1, resulting in faster acylation than with substrate PAI-1.  相似文献   

16.
Type 1 plasminogen activator inhibitor (PAI-1) is known to be synthesized in an active conformation but it is rapidly converted into an inactive conformation (t1/2 1 h) upon incubation at 37 degrees C. This study was initiated to investigate the mechanism that account for the presence of active PAI-1 in anucleated platelets that have a mean life span of 9-12 days in the circulation. Stabilization experiments with a functional immunoassay indicated that the activity of PAI-1 in both platelets and in isolated alpha-granules was prolonged in comparison to the rapid inactivation of this molecule in their lysates (t1/2 1 h). Although combined ligand blot/immunoblot analysis revealed that vitronectin was the major PAI-1 binding protein in platelets, vitronectin/PAI-1 complexes were not detected in alpha-granules using a two-site immunoassay. Co-incubation of alpha-granules with a number of agents that disrupt pH gradients (e.g. ionophores) had no effect on the stability of PAI-1 activity, whereas incubation of alpha-granules with the calcium ionophore A23187 reduced the half-life of PAI-1 to the levels observed for PAI-1 in solution. Addition of calcium ions to intact alpha-granules was an effective means of neutralizing the ionophore's effect on PAI-1 activity. Fractionation of alpha-granule proteins on molecular sieving columns using conditions known to be present within storage granules (e.g. a high calcium concentration) revealed the presence of PAI-1 in fractions with a molecular mass of > 10(6) daltons. Immunoabsorption of PAI-1 from these column fractions followed by negative staining revealed 25-nm diameter complexes of alpha-granule proteins under the electron microscope. PAI-1 activity associated with these complexes was prolonged in the presence of calcium ions and these high Mr complexes were shown to be composed of a defined set of proteins that can be dissociated from PAI-1 by chelation of calcium ions. These data indicate that PAI-1 is stabilized by its packaging with other alpha-granule proteins in a calcium-dependent manner.  相似文献   

17.
Thrombin binds to fibrin at two classes of non-substrate sites, one of high affinity and the other of low affinity. We investigated the location of these thrombin binding sites by assessing the binding of thrombin to fibrin lacking or containing gamma' chains, which are fibrinogen gamma chain variants that contain a highly anionic carboxyl-terminal sequence. We found the high affinity thrombin binding site to be located exclusively in D domains on gamma' chains (Ka, 4.9 x 10(6) M-1; n, 1.05 per gamma' chain), whereas the low affinity thrombin binding site was in the fibrin E domain (Ka, 0.29 x 10(6) M-1; n, 1.69 per molecule). The amino-terminal beta15-42 fibrin sequence is an important constituent of low affinity binding, since thrombin binding at this site is greatly diminished in fibrin molecules lacking this sequence. The tyrosine-sulfated, thrombin exosite-binding hirudin peptide, S-Hir53-64 (hirugen), inhibited both low and high affinity thrombin binding to fibrin (IC50 1.4 and 3.0 microM respectively). The presence of the high affinity gamma' chain site on fibrinogen molecules did not inhibit fibrinogen conversion to fibrin as assessed by thrombin time measurements, and thrombin exosite binding to fibrin at either site did not inhibit its catalytic activity toward a small thrombin substrate, S-2238. We infer from these findings that there are two low affinity non-substrate thrombin binding sites, one in each half of the dimeric fibrin E domain, and that they may represent a residual aspect of thrombin binding and cleavage of its substrate fibrinogen. The high affinity thrombin binding site on gamma' chains is a constitutive feature of fibrin as well as fibrinogen.  相似文献   

18.
Angiostatin is a potent inhibitor of tumor angiogenesis and the growth of metastatic foci. Recent studies have indicated that neoplastic cells can generate angiostatin directly or in cooperation with tumor-associated macrophages. In studies reported here, we determined whether angiostatin is generated in mice under non-neoplastic settings. Utilizing murine RAW264.7 macrophages and thioglycollate-elicited peritoneal macrophages, we demonstrate that angiostatin-like fragments are generated as a byproduct of the proteolytic regulation of membrane-bound plasmin. Plasmin proteolysis and subsequent loss in membrane-bound plasmin activity requires active plasmin but was unaffected by inhibitors of metalloproteinases. Lysine binding fragments of plasmin, isolated from macrophage-conditioned media utilizing affinity chromatography, appeared as a major (48 kDa) and two minor bands (42 and 50 kDa) in SDS-polyacrylamide gel electrophoresis and were immunoreactive with anti-kringle 1-3 IgG. Each peptide begins with Lys77 and contains the entire sequence of angiostatin. The affinity isolated plasmin fragments inhibited bFGF-induced endothelial cell proliferation. Lavage fluid recovered from the peritoneal cavities of mice previously injected with thioglycollate contained angiostatin-like plasmin fragments similar to those generated in vitro. This is the first demonstration that angiostatin-like plasmin fragments are generated in a non-neoplastic inflammatory setting. Thus, in addition to regulating pericellular plasmin activity, proteolysis of plasmin generates inactive kringle-containing fragments expressing angiostatic properties.  相似文献   

19.
DSEF-1 protein selectively binds to a G-rich auxiliary sequence element which influences the efficiency of processing of the SV40 late polyadenylation signal. We have obtained cDNA clones of DSEF-1 using sequence information from tryptic peptides isolated from DSEF-1 protein purified from HeLa cells. DSEF-1 protein contains three RNA-binding motifs and is a member of the hnRNP H family of RNA-binding proteins. Recombinant DSEF-1 protein stimulated the efficiency of cleavage and polyadenylation in an AAUAAA-dependent manner in in vitro reconstitution assays. DSEF-1 protein was shown to be able to interact with several poly(A) signals that lacked a G-rich binding site using a less stringent, low ionic strength gel band shift assay. Recombinant DSEF-1 protein specifically stimulated the processing of all of the poly(A) signals tested that contained a high affinity G-rich or low affinity binding site. DSEF-1 specifically increased the level of cross-linking of the 64 kDa protein of CstF to polyadenylation substrate RNAs. These observations suggest that DSEF-1 is an auxiliary factor that assists in the assembly of the general 3'-end processing factors onto the core elements of the polyadenylation signal.  相似文献   

20.
Plasminogen activator inhibitor-1 (PAI-1) is a specific inhibitor of plasminogen activators and may be the principal regulator of plasminogen activation in vivo. Abnormal expression of PAI-1 has been reported in various types of human disorders and in animal models for the diseases in relation to thrombosis. For example, plasma PAI-1 activity was elevated in patients with endotoxemia, and a dramatic induction of PAI-1 mRNA was observed in tissues of endotoxin-treated animals, resulting in tissue microthrombosis. It has also been demonstrated that PAI-1 expression levels are increased in the kidneys of mice with glomerulonephritis, in the adipose tissue of obese subjects or mice, and in human atherosclerotic arteries. This PAI-1 induction may be relevant to pathological processes in these diseases because PAI-1 not only regulated fibrin dissolution in vivo but also inhibited degradation of extracellular matrix by reducing plasmin generation. The responsible cells for abnormal expression of PAI-1 have been identified in each tissue under pathological conditions and PAI-1 synthesis appears to be regulated in a tissue-specific manner. These observations suggest that PAI-1 could play an important role in the progression of tissue pathologies in a variety of human diseases by controlling the rate of plasmin formation.  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