首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Mast cell tryptase is a tetrameric serine protease that is stored in complex with negatively charged heparin proteoglycans in the secretory granule. Tryptase has potent proinflammatory properties and has been implicated in diverse pathological conditions such as asthma and fibrosis. Previous studies have shown that tryptase binds tightly to heparin, and that heparin is required in the assembly of the tryptase tetramer as well as for stabilization of the active tetramer. Because the interaction of tryptase with heparin is optimal at acidic pH, we investigated in this study whether His residues are of importance for the heparin binding, tetramerization, and activation of the tryptase mouse mast cell protease 6. Molecular modeling of mouse mast cell protease 6 identified four His residues, H35, H106, H108, and H238, that are conserved among pH-dependent tryptases and are exposed on the molecular surface, and these four His residues were mutated to Ala. In addition, combinations of different mutations were prepared. Generally, the single His-Ala mutations did not cause any major defects in heparin binding, activation, or tetramerization, although some effect of the H106A mutation was observed. However, when several mutations were combined, large defects in all of these parameters were observed. Of the mutants, the triple mutant H106A/H108A/H238A was the most affected with an almost complete inability to bind to heparin and to form active tryptase tetramers. Taken together, this study shows that surface-exposed histidines mediate the interaction of mast cell tryptase with heparin and are of critical importance in the formation of active tryptase tetramers.  相似文献   

2.
Cynomolgus monkey tryptase was purified to homogeneity from lung tissue. Reducing SDS-PAGE analysis of the monkey enzyme produced a doublet at 30–32 kDa, which reacted with antibodies against human lung tryptase on a Western blot. N-terminal sequence analysis of the monkey enzyme yielded a sequence that was identical to human tryptase out to 15 residues. Gel filtration chromatography either in the presence or absence of heparin indicated that the monkey enzyme had a molecular mass of approximately 250 kDa and 140 kDa, respectively, consistent with the formation of a tetramer that can bind heparin. Other similarities between human and monkey tryptase included the ability to degrade vasoactive intestinal peptide and a resistance to inhibition by biological serine protease inhibitors. However, the two enzymes displayed markedly different pH stability profiles. Monkey tryptase was unstable at pH values over 7.0, even in the presence of heparin, displaying a half-life of 10.9 min at pH 8.0. In addition, the stabilizing effect of heparin was pH dependent, being most prevalent at lower pH values. Therefore, the biological activity of monkey tryptase may be controlled by both pH and the availability of heparin.  相似文献   

3.
The novel tetrameric structure of human beta-tryptase faces each active site into the central pore, thereby restricting access of most biologic protease inhibitors. The mechanism by which the anti-tryptase mAb B12 inhibits human beta-tryptase peptidase and proteolytic activities at neutral pH, but augments proteolytic activity at acidic pH, was examined. At neutral pH, B12-beta-tryptase complexes are inactive. At acidic pH, B12 (intact and Fab) minimally affects peptidase activity when added to beta-tryptase tetramers, but does induce susceptibility to inhibition by soybean trypsin inhibitor and antithrombin III. Surprisingly, B12 Fab-beta-tryptase complexes formed at both neutral and acidic pH exhibit the apparent molecular mass of a complex with 1 beta-tryptase monomer and 1 Fab by gel filtration. B12 does not compete with heparin for binding to tryptase at either neutral or acidic pH. Thus, B12 directly disrupts beta-tryptase tetramers to monomers that are inactive at neutral pH, whereas at acidic pH, are active and more accessible to protein inhibitors and substrates.  相似文献   

4.
Mast cell tryptase is stored as an active tetramer in complex with heparin in mast cell secretory granules. Previously, we demonstrated the dependence on heparin for the activation/tetramer formation of a recombinant tryptase. Here we have investigated the structural requirements for this activation process. The ability of heparin-related saccharides to activate a recombinant murine tryptase, mouse mast cell protease-6 (mMCP-6), was strongly dependent on anionic charge density and size. The dose-response curve for heparin-induced mMCP-6 activation displayed a bell-shaped appearance, indicating that heparin acts by binding to more than one tryptase monomer simultaneously. The minimal heparin oligosaccharide required for binding to mMCP-6 was 8-10 saccharide units. Gel filtration analyses showed that such short oligosaccharides were unable to generate tryptase tetramers, but instead gave rise to active mMCP-6 monomers. The active monomers were inhibited by bovine pancreatic trypsin inhibitor, whereas the tetramers were resistant. Furthermore, monomeric (but not tetrameric) mMCP-6 degraded fibronectin. Our results suggest a model for tryptase tetramer formation that involves bridging of tryptase monomers by heparin or other highly sulfated polysaccharides of sufficient chain length. Moreover, our results raise the possibility that some of the reported activities of tryptase may be related to active tryptase monomers that may be formed according to the mechanism described here.  相似文献   

5.
Tryptases betaI and betaII were heterologously expressed and purified in yeast to functionally characterize the substrate specificity of each enzyme. Three positional scanning combinatorial tetrapeptide substrate libraries were used to determine the primary and extended substrate specificity of the proteases. Both enzymes have a strict primary preference for cleavage after the basic amino acids, lysine and arginine, with only a slight preference for lysine over arginine. betaI and betaII tryptase share similar extended substrate specificity, with preference for proline at P4, preference for arginine or lysine at P3, and P2 showing a slight preference for asparagine. Measurement of kinetic constants with multiple substrates designed for beta-tryptases reveal that selectivity is highly dependent on ground state substrate binding. Coupled with the functional determinants, structural determinants of tryptase substrate specificity were identified. Molecular docking of the preferred substrate sequence to the three-dimensional tetrameric tryptase structure reveals a novel extended substrate binding mode that involves interactions from two adjacent protomers, including P4 Thr-96', P3 Asp-60B' and Glu-217, and P1 Asp-189. Based on the determined substrate information, a mechanism-based tetrapeptide-chloromethylketone inhibitor was designed and shown to be a potent tryptase inhibitor. Finally, the cleavage sites of several physiologically relevant substrates of beta-tryptases show consistency with the specificity data presented here.  相似文献   

6.
Tryptase was shown to be stabilized as an enzymatically active tetramer by association with heparin and dissociated to inactive monomers in the absence of heparin at 37 degrees C in physiologic buffer and in plasma. There was a 50% loss of tryptase activity at 37 degrees C by 6-8 min in both physiologic buffer and plasma. When heparin glycosaminoglycan was present, tryptase retained nearly full activity for 2 h in buffer and in plasma. Tryptase activity also decayed under standard assay conditions in the presence of synthetic ester and peptide substrates unless bound to heparin. That tryptase is bound to heparin at the pH and physiologic NaCl concentrations employed was shown by chromatography of tryptase on heparin-agarose, gel filtration, and velocity sedimentation. Elution of tryptase from heparin-agarose occurred at 0.8 M NaCl. Maximal stabilization of tryptase by heparin occurred at a weight ratio to tryptase that was equal to or greater than unity. Kcat/Km ratios for tryptase-heparin at 0.15 M NaCl and 37 degrees C were 0.9 X 10(6) s-1 M-1 for tosyl-L-Gly-Pro-Lys-p-nitroanilide and 1.7 X 10(6) s-1 M-1 for p-tosyl-L-arginine methyl ester and are among the highest reported for tryptic enzymes. The mechanism of heparin-dependent stabilization of tryptase was not due to indirect ion binding properties of heparin and was analyzed by Superose 12 high performance liquid chromatography. Active enzyme eluted with an apparent Mr of 132,000 +/- 10,000 (n = 3, +/- S.D.), whereas tryptase inactivated by incubation without heparin eluted with an apparent Mr of 34,000. The tetrameric structure of diisopropyl fluorophosphate-inhibited tryptase was also preserved after incubation with heparin at 37 degrees C but was reduced to monomeric subunits after incubation without heparin. That no appreciable degradation of tryptase occurs under conditions that cause dissociation of subunits was directly shown by electrophoresis in sodium dodecyl sulfate-polyacrylamide gels. Two different subunits of 34,000 and 33,000 Mr (after reduction) present in the intact enzyme (calculated to be 134,000 Mr) were also detected unchanged after inactivation of tryptase by dissociation of its subunits. Thus, the selective localization and association of heparin and tryptase in the human mast cell secretory granule most likely plays a major role in the regulation of tryptase after secretion.  相似文献   

7.
Heparin antagonists are potent inhibitors of mast cell tryptase   总被引:7,自引:0,他引:7  
Tryptase may be a key mediator in mast cell-mediated inflammatory reactions. When mast cells are activated, they release large amounts of these tetrameric trypsin-like serine proteases. Tryptase is present in a macromolecular complex with heparin proteoglycan where the interaction with heparin is known to be essential for maintaining enzymatic activity. Recent investigations have shown that tryptase has potent proinflammatory activity, and inhibitors of tryptase have been shown to modulate allergic reactions in vivo. Many of the tryptase inhibitors investigated previously are directed against the active site. In the present study we have investigated an alternative approach for tryptase regulation. We show that the heparin antagonists Polybrene and protamine are potent inhibitors of both human lung tryptase and of recombinant mouse tryptase (mouse mast cell protease 6). Protamine inhibited tryptase in a competitive manner whereas Polybrene showed noncompetitive inhibition kinetics. Treatment of tetrameric, active tryptase with Polybrene caused dissociation into monomers, accompanied by complete loss of enzymatic activity. The present report thus suggests that heparin antagonists potentially may be used in treatment of mast cell-mediated diseases such as asthma.  相似文献   

8.
Tryptase, a serine protease, is the major protein component in mast cells. In an animal model of asthma, tryptase has been established as an important mediator of inflammation and late airway responses induced by antigen challenge. Human tryptase is notable for its tetrameric structure, requirement of heparin for stability, and resistance to endogenous inhibitors. Human protryptase was expressed as a recombinant protein in Pichia pastoris. The recombinant protein consisted of two forms of protryptase, one containing the entire propeptide and the other containing only the Val-Gly dipeptide at its amino terminus. Isolation of active recombinant tryptase required a two column purification protocol and included a heparin- and dipeptidyl peptidase I-dependent activation step. Purified recombinant tryptase migrated as a tetramer on a gel filtration column and displayed kinetic parameters identical to those of a native tryptase obtained from HMC-1 cells, a human mast cell line. Recombinant and HMC-1 tryptase exhibited comparable sensitivities to an array of protein and low-molecular-weight inhibitors, including one that is highly specific for tryptase (APC-1167). Similarly, the recombinant enzyme cleaved both alpha- and beta-chains of fibrinogen to generate fibrinogen fragments indistinguishable from those generated by HMC-1-derived tryptase. Thus, recombinant tryptase expressed in P. pastoris displays physical and enzymatic properties essentially identical to the native enzyme. This system provides a cost-effective and easy to manipulate expression system that will enable the functional characterization of this unique enzyme.  相似文献   

9.
Tryptase, a serine protease with trypsin-like substrate cleavage properties, is one of the key effector molecules during allergic inflammation. It is stored in large quantities in the mast cell secretory granules in complex with heparin proteoglycan, and these complexes are released during mast cell degranulation. In the present paper, we have studied the mechanism for tryptase activation. Recombinant mouse tryptase, mouse mast cell protease 6 (mMCP-6), was produced in a mammalian expression system. The mMCP-6 fusion protein contained an N-terminal 6 x His tag followed by an enterokinase (EK) site replacing the native activation peptide (6xHis-EK-mMCP-6). In the absence of heparin, barely detectable enzyme activity was obtained after enterokinase cleavage of 6xHis-EK-mMCP-6 over a pH range of 5.5-7.5. However, when heparin was present, 6xHis-EK-mMCP-6 yielded active enzyme when enterokinase cleavage was performed at pH 5.5-6.0 but not at neutral pH. Affinity chromatography analysis showed that mMCP-6 bound strongly to heparin-Sepharose at pH 6.0 but not at neutral pH. After enterokinase cleavage of the sample at pH 6.0, mMCP-6 occurred in inactive monomeric form as shown by FPLC analysis on a Superdex 200 column. When heparin was added at pH 6.0, enzymatically active higher molecular weight complexes were formed, e.g., a dominant approximately 200 kDa complex that may correspond to tryptase tetramers. No formation of active tetramers was observed at neutral pH. When injected intraperitoneally, mMCP-6 together with heparin caused neutrophil influx, but no signs of inflammation were seen in the absence of heparin. The present paper thus indicates a crucial role for heparin in the formation of active mast cell tryptase.  相似文献   

10.
A trypsin-like enzyme (tryptase) has been purified to homogeneity from the granules of a human cytolytic lymphocyte (CTL) line, Q31, by a three-step procedure. By including 0.3% (v/v) Triton X-100 and 1 mg/ml heparin in purification buffers, near total yields of tryptase activity were obtained during the purification. The enzyme, referred to as Q31 tryptase, migrated in polyacrylamide gels with sodium dodecyl sulfate at a position corresponding to 28 kDa with and to 45 kDa without 2-mercaptoethanol. It had an amino-terminal sequence identical to a previously reported human CTL tryptase at 20 of 22 positions identified. It hydrolyzed N alpha-carbobenzyloxy-L-lysyl-thiobenzyl ester (BLT), and this BLT esterase activity was most efficient at slightly alkaline pH and was relatively more active near neutral pH than mouse CTL tryptase. Human alpha 1-protease inhibitor, human antithrombin III, phenylmethanesulfonyl fluoride, and p-aminobenzamidine inhibited the Q31 tryptase. The inhibition by human antithrombin III was rapid enough to be of physiological significance. A survey of oligopeptide p-nitroanilides found that the best substrate for human Q31 tryptase is H-D-(epsilon-carbobenzyloxy)Lys-L-Pro-L-Arg-p-nitroanilide. The Q31 tryptase appears to have broad specificity for amino acid residues at P2 and P3, i.e. at 2 and 3 residues amino-terminal to the scissile bond.  相似文献   

11.
Fukuoka Y  Schwartz LB 《Biochemistry》2004,43(33):10757-10764
beta-Tryptase is a trypsin-like serine protease stored in mast cell secretory granules primarily as an enzymatically active tetramer. The current study aims to determine whether monomeric beta-tryptase also can exhibit enzyme activity, as suggested previously. At neutral pH beta-tryptase tetramers in the absence of heparin or dextran sulfate spontaneously convert to inactive monomers. Addition of a polyanion to these monomers at neutral pH fails to convert them back to a tetramer or to an enzymatically active state. In contrast, at acidic pH addition of a polyanion resurrects enzyme activity. Whether this activity is associated with tetramers or monomers depends on the concentration of beta-tryptase. Under the experimental conditions employed at pH 6 in the presence of heparin, the monomer concentration at which 50% conversion to tetramers occurs is 193 ng/mL. Activity against tripeptide substrates by monomers is detected at pH 6 but not at pH 7.4, whereas tetramer activity is greater at pH 7.4 than pH 6.0. Active monomers are inhibited by soybean trypsin inhibitor, bovine pancreatic trypsin inhibitor, antithrombin III, and alpha2-macroglobulin, whereas active tetramers are resistant to these inhibitors. Active monomers form complexes with these inhibitors and cleave both antithrombin III and alpha2-macroglobulin. These inhibitors also prevent reconstitution of monomers to tetramers, indicating that inactive monomers become active monomers before becoming active tetramers. The ability of tryptase monomers to become active at acidic pH raises the possibilities of expanded substrate specificities as well as inhibitor susceptibilities where the low-pH environments associated with inflammation or poor vascularity are encountered in vivo.  相似文献   

12.
The catalytic activity of human tryptase, a mast cell neutral endoprotease, is expressed when the enzyme is in its tetrameric form, but is lost under physiologic conditions concomitant with a quaternary structural alteration involving conversion to a monomeric form. The associated changes in the CD spectra noted in the current study indicate accompanying alterations in the secondary structure of the protein. In particular, the progressive disappearance of the negative minimum centered at 228 nm suggests an effect on beta-sheet structure, which may be important for monomer-monomer interaction and/or stabilization of catalytic activity. Dextran sulfate, like heparin, stabilizes the catalytic activity and quaternary structure of tryptase and also maintains the native secondary structure of the enzyme at and beyond a temperature of 40 degrees C. Dextran sulfate-stabilized tryptase therefore was used as an immunogen to which were produced three murine mAb (B2, C11, and G4) recognizing the catalytically active form of the enzyme. Inactive tryptase bound to plastic microtiter wells was not recognized by any of the newly made antibodies, whereas inactive tryptase in solution was recognized by G4, which when biotinylated, could be used as a detector antibody in a sandwich ELISA for tryptase. Each of the newly made mAb recognized the catalytically active form of tryptase. Thus, alterations in epitopes, perhaps reflecting tertiary structural alterations as well as changes in secondary and quaternary conformations, occur with tryptase inactivation. A pragmatic result of these newly generated antibodies is the affinity purification to homogeneity of active tryptase by sequential chromatography with B2 coupled to CH-Sepharose and heparin-agarose. Tryptase purified by this technique had a specific activity with p-tosyl-L-arginine methyl ester of 117 +/- 9 U/mg and had 3.9 +/- 0.3 active sites per molecule of active enzyme (134,000 m.w.) as titrated with p-nitrophenyl-p'-guanidinobenzoate. The spectral and immunologic data in the current study are consistent with concerted conformational alterations in the secondary and tertiary as well as quaternary structures of tryptase associated with loss of catalytic activity. Failure to reverse any of these alterations with dextran sulfate suggests that the pathway of tetramer assembly in vivo is more complicated than simple subunit association.  相似文献   

13.
The interaction of leech-derived tryptase inhibitor (LDTI) with bovine liver capsule tryptase (BLCT) and bovine trypsin has been studied using both thermodynamic and kinetic approaches. Several differences were detected: (i) the equilibrium affinity of LDTI for BLCT (Ka = 8.9 x 10(5) M(-1)) is about 600-fold lower than that for bovine trypsin (Ka = 5.1 x 10(8) M(-1)); (ii) LDTI behaves as a purely non-competitive inhibitor of BLCT, while it is a purely competitive inhibitor of bovine trypsin. These functional data are compared with those previously reported for the LDTI binding to human tryptase, where tight inhibition occurs at two of the four active sites of the tetramer (Ka = 7.1 x 10(8) M(-1)). Amino acid sequence alignment of BLCT, human betaII-tryptase and bovine trypsin allows us to infer some possible structural basis for the observed functional differences.  相似文献   

14.
Both mast cells and complement participate in innate and acquired immunity. The current study examines whether beta-tryptase, the major protease of human mast cells, can directly generate bioactive complement anaphylatoxins. Important variables included pH, monomeric vs tetrameric forms of beta-tryptase, and the beta-tryptase-activating polyanion. The B12 mAb was used to stabilize beta-tryptase in its monomeric form. C3a and C4a were best generated from C3 and C4, respectively, by monomeric beta-tryptase in the presence of low molecular weight dextran sulfate or heparin at acidic pH. High molecular weight polyanions increased degradation of these anaphylatoxins. C5a was optimally generated from C5 at acidic pH by beta-tryptase monomers in the presence of high molecular weight dextran sulfate and heparin polyanions, but also was produced by beta-tryptase tetramers under these conditions. Mass spectrometry verified that the molecular mass of each anaphylatoxin was correct. Both beta-tryptase-generated C5a and C3a (but not C4a) were potent activators of human skin mast cells. These complement anaphylatoxins also could be generated by beta-tryptase in releasates of activated skin mast cells. Of further biologic interest, beta-tryptase also generated C3a from C3 in human plasma at acidic pH. These results suggest beta-tryptase might generate complement anaphylatoxins in vivo at sites of inflammation, such as the airway of active asthma patients where the pH is acidic and where elevated levels of beta-tryptase and complement anaphylatoxins are detected.  相似文献   

15.
Tryptase from human mast cells has been shown (in vitro) to catalyze the destruction of fibrinogen and high-molecular-weight kininogen as well as the activation of C3a and collagenase. Although large amounts of tryptase are released in tissues by degranulating mast cells and levels as high as 1000 ng/ml have been measured in the circulation following systemic anaphylaxis, no specific physiologic inhibitor has yet been found for the protease. The current work tests several more inhibitors for their effects on tryptase and examines any effect of tryptase on these inhibitors. First, antileukoprotease and low-molecular-weight elastase inhibitor from human lung and hirudin and antithrombin III had no effect on tryptase activity in vitro. Second, the possibility that tryptase, being insensitive to the effects of inhibitors, might instead destroy them was also considered. Tryptase failed to cleave and inactivate antileukoprotease, low-molecular-weight elastase inhibitor, alpha 1 protease inhibitor, alpha 2 macroglobulin, and antithrombin III. Third, based on the knowledge that tryptase stability is regulated by its interaction with heparin, antithrombin III was used as a model heparin-binding protein to demonstrate that a protein competitor for heparin-binding sites, presumably by displacement of tryptase, destabilizes this enzyme. Conversely, tryptase, in excess, blocked the binding of antithrombin III to heparin, thereby attenuating the heparin-mediated inhibition of thrombin by antithrombin III.  相似文献   

16.
Two different pyridoxal 5'-phosphate-containing l-threonine deaminases (EC 4.3.1.19), biosynthetic and biodegradative, which catalyze the deamination of l-threonine to alpha-ketobutyrate, are present in Escherichia coli and Salmonella typhimurium. Biodegradative threonine deaminase (TdcB) catalyzes the first reaction in the anaerobic breakdown of l-threonine to propionate. TdcB, unlike the biosynthetic threonine deaminase, is insensitive to l-isoleucine and is activated by AMP. In the present study, TdcB from S. typhimurium was cloned and overexpressed in E. coli. In the presence of AMP or CMP, the recombinant enzyme was converted to the tetrameric form accompanied by significant enzyme activation. To provide insights into ligand-mediated oligomerization and enzyme activation, crystal structures of S. typhimurium TdcB and its complex with CMP were determined. In the native structure, TdcB is in a dimeric form, whereas in the TdcB.CMP complex, it exists in a tetrameric form with 222 symmetry and appears as a dimer of dimers. Tetrameric TdcB binds to four molecules of CMP, two at each of the dimer interfaces. Comparison of the dimer structure in the ligand (CMP)-free and -bound forms suggests that the changes induced by ligand binding at the dimer interface are essential for tetramerization. The differences observed in the tertiary and quaternary structures of TdcB in the absence and presence of CMP appear to account for enzyme activation and increased binding affinity for l-threonine. Comparison of TdcB with related pyridoxal 5'-phosphate-dependent enzymes points to structural and mechanistic similarities.  相似文献   

17.
18.
Myeloperoxidase (MPO) is an important component of the neutrophil response to microbial infection. In this paper we report an additional activity of MPO, the potent and selective inhibition of human mast cell tryptase. MPO inhibits human mast cell tryptase in a time-dependent manner with an IC50 of 16 nM at 1 h. In contrast, MPO does not inhibit trypsin, thrombin, plasmin, factor Xa, elastase, or cathepsin G. It is the native protein conformation of MPO and not its enzyme activity that is responsible for tryptase inhibition. Heparin, at high concentrations, can prevent the inhibition of tryptase by MPO. We have shown by size-exclusion chromatography that MPO promotes the dissociation of active tryptase tetramer to inactive monomer. These data suggest that MPO inhibits tryptase by interfering with the heparin stabilization of tryptase tetramer. We have previously shown that lactoferrin (another neutrophil-associated protein) also inhibits tryptase activity by a similar mechanism. The finding that MPO is a potent inhibitor of tryptase lends further support to the hypothesis that neutrophil proteins, such as MPO and lactoferrin, may play a regulatory role as endogenous suppressers of tryptase enzyme activity.  相似文献   

19.
The fibrinogenolytic activity of purified tryptase from human lung mast cells   总被引:13,自引:0,他引:13  
The capacity of purified tryptase from human lung mast cells to metabolize human fibrinogen, fibrin, and plasminogen was evaluated. Tryptase (5 micrograms/ml) inactivated the thrombin-induced clotting activity of fibrinogen (100 micrograms/ml) with essentially similar t 1/2 values of 4.6 min in the absence of heparin and 5.8 min in the presence of heparin (20 micrograms/ml) that were not appreciably different than with lysine-Sepharose-purified plasmin (5 micrograms/ml). Fibrinogen treated with tryptase together with heparin lost all detectable clotting activity by 4 hr at 37 degrees C, whereas fibrinogen treated with tryptase alone resulted in destruction of only 80% of fibrinogen clotting equivalents after 16 hr. Tryptase alone was observed to cleave only the alpha-chains of fibrinogen by electrophoresis of tryptase-treated, denatured, and reduced fibrinogen in polyacrylamide gradient gels. Tryptase together with heparin cleaved first the alpha-chain and then the beta-chain, the latter cleavage corresponding to complete loss of fibrinogen clotting activity by 4 hr. No fibrinogen fragments with anticoagulant activity were generated by tryptase. In contrast, plasmin left no residual clotting activity after 4 hr of incubation and generated fibrinogen fragments with anticoagulant activity. Plasmin sequentially cleaved the alpha, beta, and gamma subunits of fibrinogen. Tryptase alone (6 micrograms/ml) or together with heparin (20 micrograms/ml) failed to activate plasminogen (0.6 mg/ml) after a 60-min incubation at 37 degrees C. Addition of urokinase to tryptase-treated or untreated plasminogen resulted in essentially identical plasmin activities (0.32 and 0.34 U/ml, respectively), indicating that tryptase neither activates nor destroys plasminogen. Tryptase (700 ng) also failed to substantially solubilize cross-linked fibrin (2.6 micrograms) or the corresponding amount of fibrinogen bound to plastic microtiter plates with or without heparin. The failure to solubilize fibrinogen and, possibly, fibrin is consistent with the observation that the apparent m.w. by SDS polyacrylamide gel electrophoresis of unreduced fibrinogen is not appreciably altered by prior treatment with tryptase, even though cleavage of alpha-and beta-chains is revealed after reduction. Fibrinogenolysis by tryptase complements other mast cell mediators with anticoagulant properties such as heparin and suggests a significant prevention of coagulation by activated mast cells.  相似文献   

20.
Angiogenesis, the growth of new blood vessels, is regulated by a number of factors, including hypoxia and vascular endothelial growth factor (VEGF). Although the effects of hypoxia have been studied intensely, less attention has been given to other extracellular parameters such as pH. Thus, the present study investigates the consequences of acidic pH on VEGF binding and activity in endothelial cell cultures. We found that the binding of VEGF165 and VEGF121 to endothelial cells increased as the extracellular pH was decreased from 7.5 to 5.5. Binding of VEGF165 and VEGF121 to endothelial extracellular matrix was also increased at acidic pH. These effects were, in part, a reflection of increased heparin binding, because VEGF165 and VEGF121 showed increased retention on heparin-Sepharose at pH 5.5 compared with pH 7.5. Consistent with these findings, soluble heparin competed for VEGF binding to endothelial cells under acidic conditions. However, at neutral pH (7.5) low concentrations of heparin (0.1-1.0 microg/ml) potentiated VEGF binding. Extracellular pH also regulated VEGF activation of the extracellular signal-regulated kinases 1 and 2 (Erk1/2). VEGF165 and VEGF121 activation of Erk1/2 at pH 7.5 peaked after 5 min, whereas at pH 6.5 the peak was shifted to 10 min. At pH 5.5, neither VEGF isoform was able to activate Erk1/2, suggesting that the increased VEGF bound to the cells at low pH was sequestered in a stored state. Therefore, extracellular pH might play an important role in regulating VEGF interactions with cells and the extracellular matrix, which can modulate VEGF 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