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.
Cultures of neurons from neonatal rat superior cervical, dorsal root, and trigeminal ganglia were grown in the absence of nonneuronal cells in serum-free defined medium. Proteins metabolically labeled with radioactive amino acids and spontaneously released into the culture medium were studied using two-dimensional gel electrophoresis and photofluorography. All three populations of neurons released 12-15 major proteins into the culture medium. Four proteins were released selectively by sympathetic neurons and two proteins were consistently released by both populations of sensory neurons but not by sympathetic neurons. Enzymatic activities are associated with at least two of the released proteins. One is a calcium-dependent metalloprotease, and the other a plasminogen activator. The calcium-dependent metalloprotease has a MW of 62 kDa, requires millimolar calcium for maximum activity, and has a restricted substrate specificity. It degraded native and denatured collagen more readily than casein, albumin, or fibronectin and denatured collagen (gelatin) was a better substrate than native collagen. The plasminogen activator released by neurons has a MW of 51 kDa and is converted to an active 32 kDa form. Its physiochemical properties are similar to urokinase and it was precipitated by a rabbit antiserum produced against human urokinase. A large fraction of both proteases was released by distal processes and/or growth cones suggesting that these proteases could be involved in growth cone functions.  相似文献   

2.
Two-chain 70 000-dalton plasminogen activator of tissue origin displays only weak activity toward plasminogen in a two-component system. The rate of activation is enhanced a minimum of 50-fold by the presence of fibrin clots or denatured proteins. The stimulation must depend on both chemical determinants and spatial configuration, since native proteins, including fibrinogen, lack significant stimulatory activity. These studies employed chemical modifications of four stimulatory proteins (fibrin, denatured fibrinogen, denatured IgG and denatured ovalbumin) to identify a critical role for lysine residues. Arginine, aspartic acid, cysteine, cystine, glutamic acid, histidine, methionine, tyrosine and tryptophan were found not to be essential. The critical spatial determinant(s) remain(s) unknown.  相似文献   

3.
Plasminogen activator inhibitor-1 is the main physiological regulator of tissue-type plasminogen activator in normal plasma. In addition to its critical function in fibrinolysis, plasminogen activator inhibitor-1 has been implicated in roles in other physiological and pathophysiological processes. To investigate structure-function aspects of mouse plasminogen activator inhibitor-1, the recombinant protein was expressed in Escherichia coli and purified. Five variant recombinant murine proteins (R76E, Q123K, R346A, R101A, and Q123K/R101A) were also generated using site-directed mutagenesis. The variant (R346A) was found to be defective in its inhibitory activity against tissue plasminogen activator relative to its wild-type counterpart. Enzyme-linked immunosorbent assay and surface plasmon resonance experiments demonstrated reduced vitronectin-binding affinity of the (Q123K) variant (K(D) = 1800 nm) relative to the wild-type protein (K(D) = 5.4 nm). Kinetic analyses indicated that the (Q123K) variant had a slower association (k(on) = 2.92 x 10(4) m(-1) s(-1)) to, and a faster dissociation from, vitronectin (k(off) = 5.3 x 10(-2) s(-1)), (wild-type k(on) = 1.03 x 10(6) m(-1) s(-1) and k(off) = 5.27 x 10(-3) s(-1)). The Q123K/R101A variant demonstrated an even lower vitronectin-binding ability. Low density lipoprotein receptor-related protein binding was decreased for the (R76E) variant. It was also demonstrated that the plasminogen activator inhibitor-1/vitronectin complex decreased the interaction of plasminogen activator inhibitor-1 with low density lipoprotein receptor-related protein. These results indicate that the complex interactions traditionally associated with different plasminogen activator inhibitor-1 functions apply to the murine system, thus showing a commonality of subtle functions among different species and evolutionary conservation of this protein. Further, this study provides additional evidence that the human hemostasis system can be studied effectively in the mouse, which is a great asset for investigations with gene-altered mice.  相似文献   

4.
The kinetics of tissue plasminogen activator (t-PA) and DSPAalpha1-catalyzed plasminogen activation using untreated and TAFIa-treated fibrin degradation products (FDPs), ranging in weight average molecular weight (M(w)) from 0.48 x 10(6) to 4.94 x 10(6) g/mol, were modeled according to the steady-state template model. The FDPs served as effective cofactors for both activators. The intrinsic catalytic efficiencies of both t-PA (17.4 x 10(5) m(-1) s(-1)) and DSPAalpha1 (6.0 x 10(5) m(-1) s(-1)) were independent of FDP M(w). The intrinsic catalytic efficiency of t-PA was 12-fold higher than that measured under identical conditions with intact fibrin as the cofactor. At sub-saturating levels of cofactor and substrate, rates were strongly dependent on FDP M(w) with DSPAalpha1 but not t-PA. Loss of activity with decreasing FDP M(w) correlated with loss of finger-dependent binding of the activators to the FDPs. TAFIa treatment of the FDPs resulted in 90- and 215-fold decreases in the catalytic efficiencies of t-PA (0.20 x 10(5) m(-)(1) s(-1)) and DSPAalpha1 (0.028 x 10(5) m(-1) s(-1)), yielding cofactors that were still 30- and 50-fold better than fibrinogen with t-PA and DSPAalpha1, respectively. Our results show that for both activators the products released during fibrinolysis are very effective cofactors for plasminogen activation, and both t-PA and DSPAalpha1 cofactor activity are strongly down-regulated by TAFIa.  相似文献   

5.
We showed, using the method of lysis of fibrin plates and five substrate proteins in a thin layer of agar gel, that inorganic orthophosphate (0.001-0.06 M) enhances by 50-250% the activatory functions of streptokinase, urokinase, and tissue plasminogen activator and, in general, by 1.2-12.0 times enhances protein lysis by trypsin, alpha-chymotrypsin, subtilisin, papain, bacterial metalloprotease, and even pepsin at a concentration < 4 mM. At higher concentrations, phosphate sharply inhibited pepsin activity and inhibited by 40-50% gelatin lysis by papain and gelatin (at a peak concentration) and casein lysis by metalloprotease. Inorganic pyrophosphate ions at concentrations of 10(-8)-10(-1) M enhanced the cleavage of a number of proteins by serine proteases and, at concentrations of 10(-5) -10(-3) M, the activities of pepsin, plasminogen tissue activator, and streptokinase by 100 and 40%, respectively. The pyrophosphate concentrations of > 10(-3) and >10(-4) M inhibited pepsin- and metalloprotease-induced lysis of virtually all proteins. ATP increased casein lysis by serine proteases, metalloprotease, and pepsin by 20-60% at concentration of 10(-3) M and by 30-260% at 10(-2) M concentration. At concentrations of 10-2 M, it inhibited the cleavage of some proteins by trypsin, chymotrypsin, papain, and metalloprotease by 20-100%, and, at concentrations of 10(-3) M, lysis of albumin with pepsin and other proteins (except for fibrinogen) by metalloprotease. A GTP concentration of 10(-7)-10(-2) M increased protein degradation by serine proteases, papain, and gelatin lysis by pepsin by 20-90%, whereas albumin lysis was inhibited by 40-70%. The presence of 10(-6)-10(-5) M GTP led to a slightly increased degradation of hemoglobin and casein by bacterial metalloprotease, while 10(-3) M GTP induced a drop in the activity of the metalloprotease by 20-50%. ADP could enhance gelatin lysis by trypsin, casein lysis by pepsin and papain, and inhibited metalloprotease activity by 20-100% (at 10(-3) M). Peculiarities of the effects of AMP and GD(M)P on gelatin lysis were found.  相似文献   

6.
To clarify the role of chaperones in the development of amyloid diseases, the interaction of the chaperonin GroEL with misfolded proteins and recombinant prions has been studied. The efficiency of the chaperonin-assisted folding of denatured glyceraldehyde-3-phosphate dehydrogenase (GAPDH) was shown to be decreased in the presence of prions. Prions are capable of binding to GroEL immobilized on Sepharose, but this does not prevent the interaction between GroEL and other denatured proteins. The size of individual proteins (GroEL, GAPDH, and the recombinant prion) and aggregates formed after their mixing have been determined by the dynamic light scattering analysis. It was shown that at 25°C, the non-functioning chaperonin (equimolar mixture of GroEL and GroES in the absence of Mg-ATP) bound prion yielding large aggregates (greater than 400 nm). The addition of Mg-ATP decreased significantly the size of the aggregates to 70–80 nm. After blocking of one of the chaperonin active sites by oxidized denatured GAPDH, the aggregate size increased to 1200 nm, and the addition of Mg-ATP did not prevent the aggregation. These data indicate the significant role of chaperonins in the formation of amyloid structures and demonstrate the acceleration of aggregation in the presence of functionally inactive chaperonins. The suggested model can be used for the analysis of the efficiency of antiaggregants in the system containing chaperonins.  相似文献   

7.
The kinetics of the activation of plasminogen by tissue-type plasminogen activator were studied in the presence and the absence of CNBr-digested fibrinogen as a soluble cofactor. Michaelis-Menten kinetics applied and the kinetic parameters obtained were very similar to those previously reported for the activation in the presence of solid phase fibrin (Hoylaerts, M., Rijken, D. C., Lijnen, H. R., and Collen, D. (1982) J. Biol. Chem. 257, 2912-2919). The affinity of the enzyme for plasminogen dramatically increases in the presence of the soluble cofactor while the catalytic rate constant does not change significantly (KM drops from 83 to 0.18 microM and kcat increases from 0.07 to 0.28 s-1 for tissue-type plasminogen activator of melanoma origin). Fragments containing the lysine-binding sites of plasminogen compete with plasminogen for interaction with CNBr-digested fibrinogen. The dissociation constant of this interaction was found to be 4.5 microM for the high affinity lysine-binding site. No difference was found in the kinetic parameters for the activation of plasminogen by either tissue-type plasminogen activator of melanoma origin or by glycosylated forms of tissue-type plasminogen activator obtained by recombinant DNA technology. The present findings obtained in a homogenous liquid milieu support the previously proposed mechanism of the activation of plasminogen by tissue-type plasminogen activator in the presence of fibrin. This mechanism involves binding of both tissue-type plasminogen activator and plasminogen to fibrin.  相似文献   

8.
An elastase-dependent pathway of plasminogen activation   总被引:1,自引:0,他引:1  
R Machovich  W G Owen 《Biochemistry》1989,28(10):4517-4522
In reaction mixtures containing Glu-plasminogen, alpha 2-antiplasmin, and tissue plasminogen activator or urokinase, either pancreatic or leukocyte elastase enhances the rate of plasminogen activation by 2 or more orders of magnitude. This effect is the consequence of several reactions. (a) In concentrations on the order of 100 nM, elastase degrades plasminogen within 10 min to yield des-kringle1-4-plasminogen (mini-plasminogen), which is 10-fold more efficient than Glu-plasminogen as a substrate for plasminogen activators. Des-kringle1-4-plasminogen is insensitive to cofactor activities of fibrin(ogen) fragments or an endothelial cell cofactor. (b) Des-kringle1-4-plasmin is one-tenth as sensitive as plasmin to inhibition by alpha 2-antiplasmin: k" = 10(6) M-1 s-1 versus 10(7) M-1 s-1. (c) alpha 2-Antiplasmin is disabled efficiently by elastase, with a k" of 20,000 M-1 s-1. The elastase-dependent reactions are not influenced by 6-aminohexanoate. In diluted (10-fold) blood plasma, the capacity of endogenous inhibitors to block plasmin expression is suppressed by 30 microM elastase. It is proposed that elastases provide an alternative pathway for Glu-plasminogen activation and a mechanism for controlling initiation of fibrinolysis by urokinase-type plasminogen activators.  相似文献   

9.
Fibrin (Fn) enhances plasminogen (Pg) activation by tissue-type plasminogen activator (tPA) by serving as a template onto which Pg and tPA assemble. To explore the contribution of the Pg/Fn interaction to Fn cofactor activity, Pg variants were generated and their affinities for Fn were determined using surface plasmon resonance (SPR). Glu-Pg, Lys-Pg (des(1-77)), and Mini-Pg (lacking kringles 1-4) bound Fn with K(d) values of 3.1, 0.21, and 24.5 μm, respectively, whereas Micro-Pg (lacking all kringles) did not bind. The kinetics of activation of the Pg variants by tPA were then examined in the absence or presence of Fn. Whereas Fn had no effect on Micro-Pg activation, the catalytic efficiencies of Glu-Pg, Lys-Pg, and Mini-Pg activation in the presence of Fn were 300- to 600-fold higher than in its absence. The retention of Fn cofactor activity with Mini-Pg, which has low affinity for Fn, suggests that Mini-Pg binds the tPA-Fn complex more tightly than tPA alone. To explore this possibility, SPR was used to examine the interaction of Mini-Pg with Fn in the absence or presence of tPA. There was 50% more Mini-Pg binding to Fn in the presence of tPA than in its absence, suggesting that formation of the tPA-Fn complex exposes a cryptic site that binds Mini-Pg. Thus, our data (a) indicate that high affinity binding of Pg to Fn is not essential for Fn cofactor activity, and (b) suggest that kringle 5 localizes and stabilizes Pg within the tPA-Fn complex and contributes to its efficient activation.  相似文献   

10.
Plasminogen activation is a key event in the fibrinolytic system that results in the dissolution of blood clots, and also promotes cell migration and tissue remodelling. The recent structure determinations of microplasmin in complex with the bacterial plasminogen activators staphylokinase and streptokinase have provided novel insights into the molecular mechanisms of plasminogen activation and cofactor function. These bacterial proteins are cofactor molecules that contribute to exosite formation and enhance the substrate presentation to the enzyme. At the same time, they modulate the specificity of plasmin towards substrates and inhibitors, making a 'specificity switch' possible.  相似文献   

11.
M E Baker 《FEBS letters》1985,182(1):47-52
Using a computer program designed to detect evolutionary relationships between proteins, I find that residues 72-110 of the mature sequence of human tissue-type plasminogen activator (t-PA) and 39 residues at the carboxy terminus of human albumin have a comparison score that is 8.8 standard deviation units higher than that obtained with a comparison of randomized sequences of these proteins. The probability (p) of getting this score by chance is approximately 10(-18), indicating that part of t-PA and albumin are derived from a common ancestor. I also find that alpha-fetoprotein, a relative of albumin is related to t-PA. Part of this region on t-PA has been previously shown to be related to epidermal growth factor. t-PA, albumin, alpha-fetoprotein, and epidermal growth factor have diverse biological activities. The finding that these proteins are related suggests some new approaches for studying their functions.  相似文献   

12.
Investigation of the clearance of proteins from the air spaces is important for an understanding of the resolution of pulmonary edema and also because of current interest in delivery of therapeutic peptides via the distal air spaces. Few experimental studies have examined the size dependence for alveolar clearance of large macromolecules; there have been no human studies. In anesthetized rabbits, we measured clearance of cyanocobalamin and different-sized human proteins instilled into the air spaces. After 8 h, the amounts of instilled tracer recovered in the lungs were [57Co]cyanocobalamin, 19.4 +/- 3.0% (Stokes radius 0.65 nm); 125I-labeled insulin, 64.6 +/- 3.9% (1.2 nm); 131I-labeled albumin, 87.0 +/- 4.0% (3.5 nm); and 125I-labeled immunoglobulin G, 91.8 +/- 3.3% (5.5 nm) (P < 0.05). Sieving of different-sized proteins occurred across the alveolar epithelial barrier because tracer concentrations in air space lavage fluid after 8 h were decreased more for the smaller tracers than the larger ones. Size selectivity for alveolar protein clearance in humans with resolving alveolar edema was investigated by measuring the changes in albumin and total protein concentration. The fraction of total protein concentration made up of albumin was greater in the edema fluid than in the plasma initially. The albumin fraction decreased with time in 9 of 10 patients with resolving edema, from 0.62 +/- 0.2 to 0.58 +/- 0.10 (P < 0.05) after 10 +/- 5 h. Thus both rabbit studies and human studies provide evidence for size-dependent clearance of protein from the air spaces of the lung.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Helicobacter pylori is the causative agent of chronic gastritis, peptic ulcer, and gastric malignancies. A number of virulence factors have been described including several adhesins, a cytotoxin, neutrophil-activating protein, and expression of binding of extracellular matrix proteins, like collagen type IV, laminin, and vitronectin. H. pylori strains commonly express binding of soluble plasminogen. Coccoid forms also express binding. Plasminogen binding was optimal at pH 7.0. The binding is mediated by two cell surface proteins of 42 and 57 kDa. Scatchard plot analysis showed a straight line with a K(d) of 7 x 10(-7) M. Lysine and E-aminocaproic acid inhibited binding. The binding domain on the plasminogen molecule is the fifth kringle, miniplasminogen. Plasminogen is converted to plasmin by tissue plasminogen activator. During H. pylori infection the activity of tissue plasminogen activator is decreased and that of urokinase increased. This is reversed after eradication therapy. The plasminogen binding and conversion to plasmin is the only proteolytic activity of H. pylori, and may enhance tissue penetration and be involved in carcinogenesis.  相似文献   

14.
Nag N  Krishnamoorthy G  Rao BJ 《The FEBS journal》2005,272(24):6228-6243
Changes in the oligomeric status of MutS protein was probed in solution by dynamic light scattering (DLS), and corroborated by sedimentation analyses. In the absence of any nucleotide cofactor, free MutS protein [hydrodynamic radius (Rh) of 10-12 nm] shows a small increment in size (Rh 14 nm) following the addition of homoduplex DNA (121 bp), whereas the same increases to about 18-20 nm with heteroduplex DNA containing a mismatch. MutS forms large aggregates (Rh > 500 nm) with ATP, but not in the presence of a poorly hydrolysable analogue of ATP (ATPgammaS). Addition of either homo- or heteroduplex DNA attenuates the same, due to protein recruitment to DNA. However, the same protein/DNA complexes, at high concentration of ATP (10 mm), manifest an interesting property where the presence of a single mismatch provokes a much larger oligomerization of MutS on DNA (Rh > 500 nm in the presence of MutL) as compared to the normal homoduplex (Rh approximately 100-200 nm) and such mismatch induced MutS aggregation is entirely sustained by the ongoing hydrolysis of ATP in the reaction. We speculate that the surprising property of a single mismatch, in nucleating a massive aggregation of MutS encompassing the bound DNA might play an important role in mismatch repair system.  相似文献   

15.
Akao M  Ueshima S  Okada K  Fukao H  Seki T  Ariga T  Matsuo O 《Life sciences》2003,72(15):1695-1704
The liver produces a variety of proteins including plasminogen. Plasminogen is pro-enzyme that is converted into plasmin by plasminogen activator. Plasmin has a broad substrate spectrum and participates in several biological processes, such as fibrinolysis, tissue remodeling, cell migration, angiogenesis and embryogenesis. In the present study, the regulation of plasminogen expression in mouse hepatocytes was investigated in the primary culture system. Expression level of plasminogen mRNA in the culture at the low cell density condition (0.2 x 10(5) cells / cm(2)) was compared with that at the high cell density condition (1.0 x 10 (5) cells / cm(2)). In the low cell density culture, the expression level of plasminogen mRNA decreased by a time-dependent manner. However, mRNAs for albumin and alpha(2)-antiplasmin were not influenced by the low cell density culture. On the other hand, in the high cell density culture, plasminongen mRNA expressed constantly as well as albumin and alpha(2)-antiplasmin mRNAs. Thus, the decrease in plasminogen mRNA expression could specifically occur when the density of hepatocytes was low. The down-regulation of plasminogen mRNA in the low cell density culture is not observed in the presence of cycloheximide, suggesting that the de novo protein synthesis is required for the regulatory mechanism. These findings indicate that the expression of plasminogen mRNA from hepatocyte is dependent on the cell density and the stimulation by cell-cell contact may be an important factor for the constitutive expression of plasminogen gene in hepatocytes.  相似文献   

16.
Membrane-type serine protease 1 (MT-SP1) was recently cloned, and we now report its biochemical characterization. MT-SP1 is predicted to be a type II transmembrane protein with an extracellular protease domain. This localization was experimentally verified using immunofluorescent microscopy and a cell-surface biotinylation technique. The substrate specificity of MT-SP1 was determined using a positional scanning-synthetic combinatorial library and substrate phage techniques. The preferred cleavage sequences were found to be (P4-(Arg/Lys)P3-(X)P2-(Ser)P1-(Arg)P1'-(Ala)) and (P4-(X)P3-(Arg/Lys)P2-(Ser)P1(Arg) P1'(Ala)), where X is a non-basic amino acid. Protease-activated receptor 2 (PAR2) and single-chain urokinase-type plasminogen activator are proteins that are localized to the extracellular surface and contain the preferred MT-SP1 cleavage sequence. The ability of MT-SP1 to activate PARs was assessed by exposing PAR-expressing Xenopus oocytes to the soluble MT-SP1 protease domain. The latter triggered calcium signaling in PAR2-expressing oocytes at 10 nm but failed to trigger calcium signaling in oocytes expressing PAR1, PAR3, or PAR4 at 100 nm. Single-chain urokinase-type plasminogen activator was activated using catalytic amounts of MT-SP1 (1 nm), but plasminogen was not cleaved under similar conditions. The membrane localization of MT-SP1 and its affinity for these key extracellular substrates suggests a role of the proteolytic activity in regulatory events.  相似文献   

17.
Previously, we have shown that the small heat shock protein with apparent molecular mass 27 kDa (Hsp27) does not affect the thermal unfolding of F-actin, but effectively prevents aggregation of thermally denatured F-actin [Pivovarova AV, Mikhailova VV, Chernik IS, Chebotareva NA, Levitsky DI & Gusev NB (2005) Biochem Biophys Res Commun331, 1548-1553], and supposed that Hsp27 prevents heat-induced aggregation of F-actin by forming soluble complexes with denatured actin. In the present work, we applied dynamic light scattering, analytical ultracentrifugation and size exclusion chromatography to examine the properties of complexes formed by denatured actin with a recombinant human Hsp27 mutant (Hsp27-3D) mimicking the naturally occurring phosphorylation of this protein at Ser15, Ser78, and Ser82. Our results show that formation of these complexes occurs upon heating and accompanies the F-actin thermal denaturation. All the methods show that the size of actin-Hsp27-3D complexes decreases with increasing Hsp27-3D concentration in the incubation mixture and that saturation occurs at approximately equimolar concentrations of Hsp27-3D and actin. Under these conditions, the complexes exhibit a hydrodynamic radius of approximately 16 nm, a sedimentation coefficient of 17-20 S, and a molecular mass of about 2 MDa. It is supposed that Hsp27-3D binds to denatured actin monomers or short oligomers dissociated from actin filaments upon heating and protects them from aggregation by forming relatively small and highly soluble complexes. This mechanism might explain how small heat shock proteins prevent aggregation of denatured actin and by this means protect the cytoskeleton and the whole cell from damage caused by accumulation of large insoluble aggregates under heat shock conditions.  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.
Binding of plasminogen to cultured human endothelial cells   总被引:26,自引:0,他引:26  
Endothelial cells are known to release the two major forms of plasminogen activator, tissue plasminogen activator (TPA) and urokinase. We have previously demonstrated that plasminogen (PLG) immobilized on various surfaces forms a substrate for efficient conversion to plasmin by TPA (Silverstein, R. L., Nachman, R. L., Leung, L. L. K., and Harpel, P. C. (1985) J. Biol. Chem. 260, 10346-10352). We now report the binding of human PLG to cultured human umbilical vein endothelial cell (HUVEC) monolayers, utilizing a newly devised cell monolayer enzyme-linked immunosorbent assay system. PLG binding to HUVEC was concentration dependent and saturable at physiologic PLG concentration (2 microM). Binding of PLG was 70-80% inhibited by 10 mM epsilon-aminocaproic acid, suggesting that it is largely mediated by the lysine-binding sites of PLG. PLG bound at an intermediate level to human fibroblasts, poorly to human smooth muscle cells, and not at all to bovine smooth muscle or bovine endothelial cells; unrelated proteins such as human albumin and IgG failed to bind HUVEC. PLG binding to HUVEC was rapid, reaching a steady state within 20 min, and quickly reversible. 125I-PLG bound to HUVEC with an estimated Kd of 310 +/- 235 nM (S.E.); each cell contained 1,400,000 +/- 1,000,000 (S.E.) binding sites. Functional studies demonstrated that HUVEC-bound PLG is activatable by TPA according to Michaelis-Menten kinetics (Km, 5.9 nM). Importantly, surface-bound PLG was activated with a 12.7-fold greater catalytic efficiency than fluid phase PLG. These results indicate that PLG binds to HUVEC in a specific and functional manner. Binding of PLG to endothelial cells may play a pivotal role in modulating thrombotic events at the vessel surface.  相似文献   

20.
We present evidence from small-angle X-ray scattering synchrotron experiments that porcine stomach mucin (MUC6) contains a double-globular comb structure. Analysis of the amino acid sequence of the peptide comb backbone indicates that the globular structure is determined by both the charge and hydrophobicity of the amino acids and the placement of the short hydrophilic carbohydrate side chains (approximately 2.5 nm). The double-globular structure is, thus, due to a block copolymer type hydrophobic polyampholyte charge instability in contrast to the random copolymer instabilities observed previously with synthetic polyelectrolytes (particularly polystyrene sulfonates). Careful filtering was required to exclude multimonomer aggregates from the X-ray measurements. A double Guinier analysis ( R g approximately 26 nm) and a double power law fit are consistent with two globules per chain in low salt conditions. The average radius of the globules is approximately 10 nm in salt- free condition (double Guinier fit) and the average distance of intrachain separation of the globules is 48 nm. The addition of salt causes a significant decrease in the radius of gyration (14 nm 100 mM NaCl) of the chains and is attributed to the contraction of the glycosylated peptide spacer between the two globules (the globular size continues to be approximately 10 nm and the globule separation is then 18 nm). Without salt, the scaling of the semidilute mesh size (xi) as a function of the mucin concentration (c) is xi approximately c (-0.45)compared with xi approximately c (-0.28) in high salt conditions, highlighting the globular nature of the chains. In contrast, hydrophilic flexible polyelectrolytes have a stronger concentration dependence of xi when excess salt is added.  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