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 46 毫秒
1.
Out of 29 disulfide bonds in human fibrinogen, 7 were cleaved during limited reduction under nondenaturing conditions in calcium-free buffer: 2 A alpha 442Cys-A alpha 472Cys and 2 gamma 326Cys-gamma 339Cys intrachain disulfide bonds in the carboxy-terminal ends of the A alpha- and gamma-chains and the symmetrical disulfide bonds at gamma 8Cys, gamma 9Cys, and A alpha 28Cys. We studied the loss of thrombin clottability that followed limited reduction and the increase in the susceptibility of the fibrinogen A alpha 19-A alpha 20 bond to hydrolysis by thrombin. Using differential scanning calorimetry, we show that the extent of unfolding and denaturation of specific domains following limited reduction is small. Heat absorption peaks corresponding to the melting of the major regions of compact structure give high calorimetric enthalpies, as in untreated nonreduced fibrinogen, indicating that substantial regions of native structure are still present in partially reduced fibrinogen. Thrombin releases fibrinopeptide A at an identical rate as in nonreduced fibrinogen while fibrinopeptide B release is slower. Sedimentation velocity studies show that thrombin treatment leads to complex formation; however, gelation does not occur. Amino-terminal analysis indicates that the second thrombin cleavage in the A alpha-chain at A alpha 19-A alpha 20 takes place only after fibrinopeptide A release. Thus, the loss of clottability appears to result from perturbation of carboxy-terminal polymerization sites, probably a consequence of gamma 326Cys-gamma 339Cys intrachain disulfide bond cleavage. The thrombin-treated partially reduced fibrinogen remains soluble in buffered saline and fully expresses at least one epitope, B beta 15-21, unique to fibrin. Furthermore, this nonclottable form accelerates the tissue plasminogen activator dependent conversion of plasminogen to plasmin.  相似文献   

2.
During cleavage of fibrinogen by thrombin, fibrinopeptide A (FpA) release precedes fibrinopeptide B (FpB) release. To examine the basis for this ordered release, we synthesized A'beta fibrinogen, replacing FpB with a fibrinopeptide A-like peptide, FpA' (G14V). Analyses of fibrinopeptide release from A'beta fibrinogen showed that FpA release and FpA' release were similar; the release of either peptide followed simple first-order kinetics. Specificity constants for FpA and FpA' were similar, demonstrating that these peptides are equally competitive substrates for thrombin. In the presence of Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization, the rate of FpB release from normal fibrinogen was reduced 3-fold, consistent with previous data; in contrast, the rate of FpA' release from A'beta fibrinogen was unaffected. Thus, with A'beta fibrinogen, fibrinopeptide release from the beta chain is similar to fibrinopeptide release from the alpha chain. We conclude that the ordered release of fibrinopeptides is dictated by the specificity of thrombin for its substrates. We analyzed polymerization, following changes in turbidity, and found that polymerization of A'beta fibrinogen was similar to that of normal fibrinogen. We analyzed clot structure by scanning electron microscopy and found that clots from A'beta fibrinogen were similar to clots from normal fibrinogen. We conclude that premature release of the fibrinopeptide from the N terminus of the beta chain does not affect polymerization of fibrinogen.  相似文献   

3.
Mullin JL  Gorkun OV  Lord ST 《Biochemistry》2000,39(32):9843-9849
We analyzed the polymerization of BbetaA68T fibrinogen, the recombinant counterpart of fibrinogen Naples, a variant known to have decreased thrombin binding. When polymerized with equal thrombin concentrations, BbetaA68T fibrinogen had a longer lag time and lower rate of lateral aggregation, V(max), than normal recombinant fibrinogen, but a similar final turbidity. At thrombin concentrations that equalized the rates of fibrinopeptide A release, BbetaA68T fibrinogen polymerized with a lag time and V(max) similar to normal, but reached a significantly lower final turbidity. Similar results were produced when BbetaA68T was polymerized with Ancrod, which cleaves fibrinopeptide A at the same rate from either fibrinogen, and when BbetaA68T desA monomers were polymerized. The polymerization of desAB fibrin monomers, which circumvents fibrinopeptide release, was the same for both fibrinogens. We confirmed that turbidity was indicative of fiber thickness by scanning electron microscopy of fibrin clots. Here, we present the first experimental evidence of fibrin polymerization with a normal period of protofibril formation and rate of lateral aggregation, but with a significantly decreased extent of lateral aggregation. We conclude that the decreased lateral aggregation seen in BbetaA68T fibrinogen is due to an altered step in the enzymatic phase of its polymerization process. We propose that during normal polymerization a subtle conformational change in the E domain occurs, between the release of FpA and FpB, and that this change modulates the mechanism of lateral aggregation. Without this change, the lateral aggregation of BbetaA68T fibrinogen is impaired such that variant clots have thinner fibers than normal clots.  相似文献   

4.
We used fluorescence correlation spectroscopy (FCS) to study the activation of fibrinogen by thrombin and the subsequent aggregation of fibrin monomers into fibrin polymers at a very low and at physiological fibrinogen concentrations. In the labeling procedure used the fibrinogen was randomly labeled and the label was bound to the fibrinopeptide A and/or to the part of fibrinogen which after activation takes part in fibrin formation. We measured a diffusion coefficient for fibrinogen of 2.48 x 10(-7) +/- 0.10 x 10(-7) cm2/s. After activation with thrombin both fibrinopeptide A and fibrin polymerization products could be demonstrated. From our findings we suggest a model for the formation of a three-dimensional network as two parallel processes, elongation and branching and that fibrin oligomers are not only intermediates in the polymerization process but also are substrates for branching.  相似文献   

5.
Localization of a fibrin polymerization site   总被引:6,自引:0,他引:6  
The formation of a fibrin clot is initiated after the proteolytic cleavage of fibrinogen by thrombin. The enzyme removes fibrinopeptides A and B and generates fibrin monomer which spontaneously polymerizes. Polymerization appears to occur though the interaction of complementary binding sites on the NH2-terminal and COOH-terminal (Fragment D) regions of the molecule. A peptide has been isolated from the gamma chain remnant of fibrinogen Fragment D1 which has the ability to bind to the NH2-terminal region of fibrinogen as well as to inhibit fibrin monomer polymerization. The peptide reduces the maximum rate and extent of the polymerization of thrombin or batroxobin fibrin monomer and increases the lag time. The D1 peptide does not interact with disulfide knot, fibrinogen, or Fragment D1, but it binds to thrombin-treated disulfide knot with a Kd of 1.45 X 10(-6) M at approximately two binding sites per molecule of disulfide knot. Fibrin monomer formed either by thrombin or batroxobin binds approximately two molecules of D1 peptide per molecule of fibrin monomer, indicating that the complementary site is revealed by the loss of fibrinopeptide A. The NH2-terminal sequence (Thr-Arg-Trp) and COOH-terminal sequence (Ala-Gly-Asp-Val) of the D1 peptide were determined. Therefore the gamma 373-410 region of fibrinogen contains a polymerization site which is complementary to the thrombin-activated site on the NH2-terminal region of fibrinogen.  相似文献   

6.
1. Improved methods for the purification of lamprey thrombin and fibrinogen are presented. 2. Lamprey thrombin releases two fibrinopeptides from lamprey fibrinogen during the transformation into fibrin. Bovine thrombin releases only one of these, a peptide referred to as fibrinopeptide B. The differences in the by-products of fibrin formation are reflected in the different N-terminal amino acid compositions of the two types of fibrin. 3. The fibrinopeptide that is not removed from the lamprey fibrinogen by bovine thrombin can subsequently be released by treatment of that fibrin with lamprey thrombin. 4. Under the conditions used, lamprey thrombin releases both fibrinopeptides at about the same rate. 5. The differences in interaction among these pairs of related proteins are extreme manifestations of the phenomenon loosely referred to as `species specificity'.  相似文献   

7.
Wound healing is a complex process initiated by the formation of fibrin fibers and endothelialization. Normally, this process is triggered in a wound by thrombin cleavage of fibrinopeptides on fibrinogen molecules, which allows them to self spontaneously-assemble into large fibers that provide the support structure of the clot and promote healing. We have found that the fibrous structures can also form without thrombin on most polymer or metal surfaces, including those commonly used for stents. We show that the relatively hydrophobic E and D regions of the fibrinogen molecule are adsorbed on these surfaces, exposing the αC domains, which in turn results in the formation of large fiber structures that promote endothelial cell adhesion. We show that the entire process can be suppressed when stents or other substrates are coated with polymers that are functionalized to bind the αC domains, leading to the development of potentially nonthrombogenic implant materials.  相似文献   

8.
Heparin, in Langmuirian fashion, binds stoichiometrically with high affinity, Kd approximately 100 nM, to both fibrinogen and fibrin adsorbed as monomolecular films to lecithin-coated, microscopic, polystyrene-divinylbenzene beads. Complex formation inhibits aggregation of fibrin-coated beads, and it also results in dissociation of preformed aggregates of fibrin-coated beads. These phenomena are not caused by desorption of fibrin(ogen), indirect inhibition of thrombin activity, or mere electrostatic repulsion of charged particles. Instead, these data are consistent with the proposal that the complexed heparin interferes directly with dimer formation between fibrin molecules adsorbed to colliding beads. We describe these phenomena and their application to the development of sensitive analytical methods for quantitating heparin. Based on these observations, we also propose a role for endogenous heparin in the physiologic regulation of fibrin-mediated adhesion of surfaces.  相似文献   

9.
The structure of the ternary complex of human alpha-thrombin with a covalently bound analogue of fibrinopeptide A and a C-terminal hirudin peptide has been determined by X-ray diffraction methods at 0.25 nm resolution. Fibrinopeptide A folds in a compact manner, bringing together hydrophobic residues that slot into the apolar binding site of human alpha-thrombin. Fibrinogen residue Phe8 occupies the aryl-binding site of thrombin, adjacent to fibrinogen residues Leu9 and Val15 in the S2 subsite. The species diversity of fibrinopeptide A is analysed with respect to its conformation and its interaction with thrombin. The non-covalently attached peptide fragment hirudin(54-65) exhibits an identical conformation to that observed in the hirudin-thrombin complex. The occupancy of the secondary fibrinogen-recognition exosite by this peptide imposes restrictions on the manner of fibrinogen binding. The surface topology of the thrombin molecule indicates positions P1'-P3', differ from those of the canonical serine-proteinase inhibitors, suggesting a mechanical model for the switching of thrombin activity from fibrinogen cleavage to protein-C activation on thrombomodulin complex formation. The multiple interactions between thrombin and fibrinogen provide an explanation for the narrow specificity of thrombin. Structural grounds can be put forward for certain congenital clotting disorders.  相似文献   

10.
Unmodified and polyethylene glycol (PEG) modified neutral and negatively charged liposomes were prepared by freeze-thaw and extrusion followed by chromatographic purification. The effects of PEG molecular weight (PEG 550, 2000, 5000), PEG loading (0-15 mol%), and liposome surface charge on fibrinogen adsorption were quantified using radiolabeling techniques. All adsorption isotherms increased monotonically over the concentration range 0-3 mg/ml and adsorption levels were low. Negatively charged liposomes adsorbed significantly more fibrinogen than neutral liposomes. PEG modification had no effect on fibrinogen adsorption to neutral liposomes. An inverse relationship was found between PEG loading of negatively charged liposomes and fibrinogen adsorption. PEGs of all three molecular weights at a loading of 5 mol% reduced fibrinogen adsorption to negatively charged liposomes. Protein adsorption from diluted plasma (10% normal strength) to four different liposome types (neutral, PEG-neutral, negatively charged, and PEG-negatively charged) was investigated using gel electrophoresis and immunoblotting. The profiles of adsorbed proteins were similar on all four liposome types, but distinctly different from the profile of plasma itself, indicating a partitioning effect of the lipid surfaces. alpha2-macroglobulin and fibronectin were significantly enriched on the liposomes whereas albumin, transferrin, and fibrinogen were depleted compared to plasma. Apolipoprotein AI was a major component of the adsorbed protein layers. The blot of complement protein C3 adsorbed on the liposomes suggested that the complement system was activated.  相似文献   

11.
Adsorption of the polylysine and of the copolypeptides: L-lysine/L-serine and L-lysine/L-phenylalanine on phospholipid monolayers has been investigated. The charge density of the monolayers was varied by using the negatively charged phosphatidyl serine and the neutral phosphatidyl choline at different ratios. The surface concentrations of the adsorbed polypeptides was determined by measuring the surface radiation of their radioactive label.The adsorbing capacity of the monolayer surfaces increases with their negative charge, however with respect to polypeptides the surface activity sequence is pL < pLS < pLφ. From the dependence of adsorption on the ionic strength it was concluded that it is controlled by three types of interaction: (1) electrostatic attraction to the negatively charged surface; (2) electrostatic repulsion between adsorbed polybases; (3) hydrophobic interactions involving specific structural arrangements. This is true even of the apparently neutral PC monolayer where the fixed phosphate groups form an electrical double layer with the more mobile choline groups which can be interpenetrated by the charged groups of the basic polypeptides.  相似文献   

12.
Cha T  Guo A  Zhu XY 《Biophysical journal》2006,90(4):1270-1274
Electrostatic interaction is known to play important roles in the adsorption of charged lipids on oppositely charged surfaces. Here we show that, even for charge neutral (zwitterionic) lipids, electrostatic interaction is critical in controlling the adsorption and fusion of lipid vesicles to form supported phospholipid bilayers (SPBs) on surfaces. We use terminally functionalized alkanethiol self-assembled monolayers (SAMs) to systematically control the surface charge density. Charge neutral egg phophatidylcholine (eggPC) vesicles readily fuse into SPBs on either a positively charged 11-aminino-1-undecanethiol SAM or a negatively charged 10-carboxy-1-decanethiol SAM when the density of surface charge groups is > or = 80%. These processes depend critically on the buffer environment: fusion of adsorbed vesicles to form SPBs on each charged molecular surface does not occur when the molecular ion of the buffer used is of the opposite charge type. We attribute this to the high entropic repulsion (electric double layer repulsion) due to the large size of molecular counterions. On the other hand, such a critical dependence on buffer type is not observed when charged lipids are used. This study suggests the general importance of controlling electrostatic interaction in the formation of stable SPBs.  相似文献   

13.
The effect of desialylation of fibrinogen on its conversion to fibrin has been investigated with particular reference to the kinetics of clot formation and structure. Also examined was the role of sialic acid in fibrinogen (factor I) poor in factor XIII (fibrinstabilizing factor) and factor I containing F XIII. The removal of more than 90% of the sialic acid of fibrinogen does not alter the thrombin clotting time, the clot solubility in monochloroacetic acid, the extent of cross-linking in the fibrin polymer, or the firmness and elasticity of the evolved clot. The data indicate that the sialic acid residues of fibrinogen do not contribute significantly to its conversion to fibrin by thrombin.  相似文献   

14.
1. Four clotting factors, Cf-1(C), Cf-2(C), Cf-1(T) and Cf-2(T) were isolated from Agkistrodon acutus (collected on mainland China and Taiwan) venom by Komori et al. (1987). It was reported that all factors possessed coagulant activity in the conversion of fibrinogen to fibrin, although they showed different chemical properties and antigenicities. 2. Their role in the clot formation system was clarified and compared with that of thrombin. Clotting factors from A. acutus venom released only fibrinopeptide A from the A alpha chain of fibrinogen, while thrombin released fibrinopeptide A and B from the A alpha and B beta chains. 3. Cf-1(C) and Cf-2(T), like thrombin, rapidly activated factor XIII in the presence of calcium ions, whereas Cf-2(C) and Cf-1(T) had little effect on factor XIII. These effects are shown by Cf-1(C) and Cf-2(T) forming a clot that remained insoluble in 8 M urea or 0.44 M monochloroacetic acid, whereas Cf-2(C) and Cf-1(T) formed a soluble clot in these agents.  相似文献   

15.
Magnetically induced birefringence was used to monitor fibrin polymerization after the release of the small negatively charged A fibrinopeptides from human fibrinogen by the action of the snake-venom-derived enzymes reptilase and ancrod. A range of conditions was investigated. Fibrin polymerization in solutions of purified fibrinogen shows a distinct break near the gelation point. On addition of Ca2+ or albumin the lag period is shortened, fibre thickness is increased and the break in assembly almost vanishes, probably because both of these additives promote lateral aggregation. There are minor differences in the kinetics, depending on the venom enzyme used. The kinetics of fibrin assembly in model systems containing either Ca2+ or albumin and in human plasma with a largely dormant coagulation cascade are very similar. Therefore in the latter condition there is no significant alteration in the assembly process due to interaction between fibrin or the venom enzymes and any of the plasma proteins. When the cascade is activated, the polymerization progress curves have a character that resembles a combination of the reactions observed when the venom enzymes and endogenously generated thrombin separately induce coagulation, except for a region near gelation where, paradoxically, polymerization appears to be slower on activation. The low-angle neutron-diffraction patterns from oriented gels made with thrombin or reptilase are identical. Therefore at low resolution the packing of the monomers within fibres is the same when fibrinopeptide A only or both fibrinopeptides A and B are removed.  相似文献   

16.
Endothelial cells and platelets are reported to have receptors for alpha-thrombin. To visualize the binding of alpha-thrombin to these cells, we developed a method to label thrombin with colloidal gold. Formed by electrostatic adsorption of thrombin to the negatively charged gold, the resulting probe is stable for weeks and consists of approximately 30 thrombin molecules adsorbed to each 16.5 nm gold particle. The probe retained about 10% of the enzymatic activity (fibrinogen clotting) of the unlabeled native thrombin and 20% of the ability of the native thrombin to aggregate platelets in platelet-rich plasma (PRP). In PRP, approximately 90% of the observed probes were bound to fibrin strands, with the remaining probes (650 per cell) attached to activated platelets. In contrast, washed, paraformaldehyde-fixed human platelets exhibited a marked increase in probe density (4900 per cell). Time-dependent ultrastructural studies (2-240 min) of binding of the thrombin-gold probe to confluent cultures of porcine aortic endothelial cells revealed that the initial binding (7300 probes per cell) occurred randomly at the cell surface. A limited number (25%) of the probes clustered at coated-pit regions and were internalized (60-240 min). The probe induced a limited amount of cellular retraction similar to that achieved with unlabeled thrombin. These results suggest that the thrombin-gold probe is suitable for investigations of the localization of thrombin receptors on cell surfaces and the interaction of thrombin with these receptors during thrombotic events.  相似文献   

17.
Factor XIII zymogen activation is a complex series of events that involve fibrinogen acting in several different roles. This report focuses on the role of fibrinogen as a cofactor in factor XIII activation by thrombin. We demonstrate that fibrinogen has two distinct activities that lead to an increased rate of factor XIII activation. First, the thrombin proteolytic activity is increased by fibrin. The cleavage rates of both a small chromogenic substrate and the factor XIII activation peptide are increased in the presence of either the major fibrin isoform, gammaA/gammaA fibrin, or a minor variant form, gammaA/gamma' fibrin. This enhancement of thrombin activity by fibrin is independent of fibrin polymerization and requires only cleavage of the fibrinopeptides. Subsequently, gammaA/gamma' fibrinogen accelerates plasma factor XIII activation by a non-proteolytic mechanism. This increased rate of activation results in a slightly more rapid cross-linking of fibrin gammaA and gamma' chains and a significantly more rapid cross-linking of fibrin alpha chain multimers. Together, these results show that although both forms of fibrin increase the rate of activation peptide cleavage by thrombin, gammaA/gamma' fibrinogen also increases the rate of factor XIII activation in a non-proteolytic manner. A revised model of factor XIII activation is presented below.  相似文献   

18.
Inhibition of the enzymatic activity of thrombin by concanavalin A   总被引:1,自引:0,他引:1  
Concanavalin A, a carbohydrate lectin derived from the jack bean, prolongs the thrombin clotting time of human plasma or purified fibrinogen. Prolongation is due to delay in peptide release from fibrinogen. The rate of fibrin monomer polymerization is not affected. Hydrolysis of protamine sulfate by thrombin is inhibited by concanavalin A. All inhibitory effects are prevented by α-methyl-D-mannoside. Concanavalin A does not delay clotting of fibrinogen by reptilase (releases fibrinopeptide A only) or by Ancistrodon contortrix contortrix (releases fibrinopeptide B initially followed by a small amount of A). It is concluded that concanavalin A binds to a carbohydrate on the thrombin molecule thus inhibiting its enzymatic activity.  相似文献   

19.
The effect of plasmin-derived fibrin(ogen) degradation products on alpha-thrombin cleavage of plasma Factor XIII was studied to identify the fibrin polymer structure that promotes Factor XIIIa formation. Fibrin polymers derived from fibrinogen and Fragment X enhanced the rate of thrombin cleavage of plasma Factor XIII in plasma or buffered solutions. The concentrations of fibrinogen and Fragment X that promoted half-maximal rates of Factor XIIIa formation were 5 and 40 micrograms/ml, respectively. Fragments Y, D, E, D-dimer, and photooxidized fibrinogen did not enhance thrombin cleavage of Factor XIII. Although purified Fragment D1 inhibited fibrin gelation, the soluble protofibrils promoted thrombin activation of Factor XIII. Noncrosslinked fibrin fibers failed to enhance thrombin cleavage of Factor XIII. In conclusion, soluble fibrin oligomers function to promote thrombin cleavage of plasma Factor XIII during blood clotting.  相似文献   

20.
Nonsubstrate interaction of thrombin with fibrinogen promotes sequential cleavage of fibrinopeptides A and B (fpA and fpB, respectively) from the latter, resulting in its conversion into fibrin. The recently established crystal structure of human thrombin in complex with the central part of human fibrin clarified the mechanism of this interaction. Here, we reveal new details of the structure and present the results of molecular modeling of the fpA- and fpB-containing portions of the Aalpha and Bbeta chains, not identified in the complex, in both fibrinogen and protofibrils. The analysis of the results reveals that in fibrinogen the fpA-containing portions are in a more favorable position to bind in the active site cleft of bound thrombin. Surface plasmon resonance experiments establish that the fpB-containing portions interact with the fibrin-derived dimeric D-D fragment, suggesting that in protofibrils they bind to the newly formed DD regions bringing fpB into the vicinity of bound thrombin. These findings provide a coherent rationale for the preferential removal of fpA from fibrinogen at the first stage of fibrin assembly and the accelerated cleavage of fpB from protofibrils and/or fibrils at the second stage.  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