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 787 毫秒
1.
The structure and growth of a blood clot depend on the localization of tissue factor (TF), which can trigger clotting during the hemostatic process or promote thrombosis when exposed to blood under pathological conditions. We sought to understand how the growth, structure, and mechanical properties of clots under flow are shaped by the simultaneously varying TF surface density and its exposure area. We used an eight-channel microfluidic device equipped with a 20- or 100-μm-long collagen surface patterned with lipidated TF of surface densities ~0.1 and ~2 molecules/μm2. Human whole blood was perfused at venous shear, and clot growth was continually measured. Using our recently developed computational model of clot formation, we performed simulations to gain insights into the clot’s structure and its resistance to blood flow. An increase in TF exposure area resulted not only in accelerated bulk platelet, thrombin, and fibrin accumulation, but also in increased height of the platelet mass and increased clot resistance to flow. Moreover, increasing the TF surface density or exposure area enhanced platelet deposition by approximately twofold, and thrombin and fibrin generation by greater than threefold, thereby increasing both clot size and its viscous resistance. Finally, TF effects on blood flow occlusion were more pronounced for the longer thrombogenic surface than for the shorter one. Our results suggest that TF surface density and its exposure area can independently enhance both the clot’s occlusivity and its resistance to blood flow. These findings provide, to our knowledge, new insights into how TF affects thrombus growth in time and space under flow.  相似文献   

2.
《Biophysical journal》2023,122(4):697-712
During clotting under flow, thrombin rapidly generates fibrin, whereas fibrin potently sequesters thrombin. This co-regulation was studied using microfluidic whole blood clotting on collagen/tissue factor, followed by buffer wash, and a start/stop cycling flow assay using the thrombin fluorogenic substrate, Boc-Val-Pro-Arg-AMC. After 3 min of clotting (100 s?1) and 5 min of buffer wash, non-elutable thrombin activity was easily detected during cycles of flow cessation. Non-elutable thrombin was similarly detected in plasma clots or arterial whole blood clots (1000 s?1). This thrombin activity was ablated by Phe-Pro-Arg-chloromethylketone (PPACK), apixaban, or Gly-Pro-Arg-Pro to inhibit fibrin. Reaction-diffusion simulations predicted 108 nM thrombin within the clot. Heparin addition to the start/stop assay had little effect on fibrin-bound thrombin, whereas addition of heparin-antithrombin (AT) required over 6 min to inhibit the thrombin, indicating a substantial diffusion limitation. In contrast, heparin-AT rapidly inhibited thrombin within microfluidic plasma clots, indicating marked differences in fibrin structure and functionality between plasma clots and whole blood clots. Addition of GPVI-Fab to blood before venous or arterial clotting (200 or 1000 s?1) markedly reduced fibrin-bound thrombin, whereas GPVI-Fab addition after 90 s of clotting had no effect. Perfusion of AF647-fibrinogen over washed fluorescein isothiocyanate (FITC)-fibrin clots resulted in an intense red layer around, but not within, the original FITC-fibrin. Similarly, introduction of plasma/AF647-fibrinogen generated substantial red fibrin masses that did not penetrate the original green clots, demonstrating that fibrin cannot be re-clotted with fibrinogen. Overall, thrombin within fibrin is non-elutable, easily accessed by peptides, slowly accessed by average-sized proteins (heparin/AT), and not accessible to fresh fibrinogen.  相似文献   

3.
Gel formation by fibrin oligomers without addition of monomers   总被引:1,自引:0,他引:1  
P A Janmey  J D Ferry 《Biopolymers》1986,25(7):1337-1344
Soluble fibrin oligomers were formed by reacting fibrinogen with thrombin under fine clotting conditions where the action of thrombin is the rate-determining step for polymerization, and by inhibiting the reaction shortly before gelation. Oligomeric fibrin was separated from unreacted fibrinogen and small oligomers by gel permeation chromatography. Electron microscopy revealed that the largest soluble fibrin oligomers resemble the protofibrils present in fine clots, but are somewhat shorter and entirely lack the twisted, trifunctional junctions that contribute to the elastic properties of fine clots. When thrombin was added to the soluble fibrin oligomers, polymerization resumed and clots were formed at a more rapid rate than from fibrinogen at the same concentration and resulted in a less-opaque clot under coarse clotting conditions. The results confirm a prediction of a theory for the polymerization of fibrin and provide additional evidence that the final state of a coarse fibrin clot depends on the mobility of protofibrils during its formation.  相似文献   

4.
The storage and loss shear moduli (G', G″) of human fibrin clots have been measured in small oscillating deformations over a frequency range of 0.01 to 160 Hz with the modified Birnboim transducer apparatus. Most clots were prepared by the action of thrombin on purified fibrinogen, under various conditions of pH and ionic strength to produce networks ranging from coarse to fine structure; some were liaated by fibrinoligase. The fine, unligated clot showed very little mechanical loss or frequency dependence of G' over the experimental frequency range, though loss mechanisms evidently appear at higher frequencies; G' was proportional to the 1.5 power of fibrin concentration. The coarse, unligated clot showed a slight increase of G' with frequency, reflecting some relaxation mechanisms with time constants whose reciprocals lie in the experimental frequency range. Ligation did not greatly affect the magnitude of G'. However, clots prepared by dilution of solutions of fibrin monomer in 1 M sodium bromide had smaller moduli by a factor of ten than corresponding clots prepared by the action of thrombin of fibrinogen. Oscillatory measurements in the Birnboim apparatus with closed-end (annular pumping) geometry revealed a low-frequency anomaly which was shown to be due to permeation of fluid through the clot structure, and from these measurements the Darcy constants for coarse clots were calculated. From the Darcy constants, the average thicknesses of the fibrous elements of the structures were estimated to be from 300 to 700 A.  相似文献   

5.
Microfluidic devices allow for the controlled perfusion of human or mouse blood over defined prothrombotic surfaces at venous and arterial shear rates. To mimic in vivo injuries such a plaque rupture, the need exists to link lipidated tissue factor (TF) to surface-bound collagen fibers. Recombinant TF was relipidated in liposomes of phosphatidylserine/phosphatidylcholine/biotin-linked phosphatidylethanolamine (20:79:1 PS/PC/bPE molar ratio). Collagen was patterned in a 250-μm-wide stripe and labeled with biotinylated anticollagen antibody which was then bound with streptavidin, allowing the subsequent capture of the TF liposomes. To verify and detect the TF liposome-collagen assembly, individual molecular complexes of TF-factor VIIa on collagen were visualized using the proximity ligation assay (PLA) to produce discretely localized fluorescent events that were strictly dependent on the presence of factor VIIa and primary antibodies against TF or factor VIIa. Perfusion for 450 s (wall shear rate, 200 s(-1)) of corn trypsin inhibitor (CTI, a factor XIIa inhibitor) treated whole blood over the stripe of TF-collagen enhanced platelet adhesion by 30 ± 8% (p < 0.001) and produced measurable fibrin (>50-fold increase) as compared to surfaces lacking TF. PS/PC/bPE liposomes lacking TF resulted in no enhancement of platelet deposition. Essentially no fibrin was formed during perfusion over collagen surfaces or collagen surfaces with liposomes lacking TF despite the robust platelet deposition, indicating a lack of kinetically significant platelet-borne tissue factor in healthy donor blood. This study demonstrates a reliable approach to link functionally active TF to collagen for microfluidic thrombosis studies.  相似文献   

6.
Studies suggest that patients with deep vein thrombosis and diabetes often have hypercoagulable blood plasma, leading to a higher risk of thromboembolism formation through the rupture of blood clots, which may lead to stroke and death. Despite many advances in the field of blood clot formation and thrombosis, the influence of mechanical properties of fibrin in the formation of thromboembolisms in platelet-poor plasma is poorly understood. In this paper, we combine the concepts of reactive molecular dynamics and coarse-grained molecular modeling to predict the complex network formation of fibrin clots and the branching of fibrin monomers. The 340-kDa fibrinogen molecule was converted into a coarse-grained molecule with nine beads, and using our customized reactive potentials, we simulated the formation and polymerization process of a fibrin clot. The results show that higher concentrations of thrombin result in higher branch-point formation in the fibrin clot structure. Our results also highlight many interesting properties, such as the formation of thicker or thinner fibers depending on the thrombin concentration. To the best of our knowledge, this is the first successful molecular polymerization study of fibrin clots to focus on thrombin concentration.  相似文献   

7.
Thrombin exosite 1 binds the predominant gamma(A)/gamma(A)-fibrin form with low affinity. A subpopulation of fibrin molecules, gamma(A)/gamma'-fibrin, has an extended COOH terminus gamma'-chain that binds exosite 2 of thrombin. Bivalent binding to gamma(A)/gamma'-fibrin increases the affinity of thrombin 10-fold, as determined by surface plasmon resonance. Because of its higher affinity, thrombin dissociates 7-fold more slowly from gamma(A)/gamma'-fibrin clots than from gamma(A)/gamma(A)-fibrin clots. After 24 h of washing, however, both gamma(A)/gamma'- and gamma(A)/gamma(A)-fibrin clots generate fibrinopeptide A when incubated with fibrinogen, indicating the retention of active thrombin. Previous studies demonstrated that heparin heightens the affinity of thrombin for fibrin by simultaneously binding to fibrin and exosite 2 on thrombin to generate a ternary heparin-thrombin-fibrin complex that protects thrombin from inhibition by antithrombin and heparin cofactor II. In contrast, dermatan sulfate does not promote ternary complex formation because it does not bind to fibrin. Heparin-catalyzed rates of thrombin inhibition by antithrombin were 5-fold slower in gamma(A)/gamma'-fibrin clots than they were in gamma(A)/gamma(A)-fibrin clots. This difference reflects bivalent binding of thrombin to gamma(A)/gamma'-fibrin because (a) it is abolished by addition of a gamma'-chain-directed antibody that blocks exosite 2-mediated binding of thrombin to the gamma'-chain and (b) the dermatan sulfate-catalyzed rate of thrombin inhibition by heparin cofactor II also is lower with gamma(A)/gamma'-fibrin than with gamma(A)/gamma(A)-fibrin clots. Thus, bivalent binding of thrombin to gamma(A)/gamma'-fibrin protects thrombin from inhibition, raising the possibility that gamma(A)/gamma'-fibrin serves as a reservoir of active thrombin that renders thrombi thrombogenic.  相似文献   

8.
During thrombolytic therapy and after recanalization is achieved, reduction in the volume of mural thrombi is desirable. Mural thrombi are known to contribute to rethrombosis and reocclusion. The lysis rate of mural thrombi has been demonstrated to increase with fluid flow in different experimental models, but the mechanisms responsible are unknown. An experimental and a theoretical study were developed to determine the contribution of outer convective transport to the lysis of mural fibrin clots. Normal human plasma containing recombinant tissue-type plasminogen activator (tPA; 0.5 microg/mL) was (re)perfused over mural fibrin clots with fluorescently labeled fibrin at low arterial, arterial, or higher wall shear stresses (4, 18, or 41 dyn/cm(2), respectively) and lysis was monitored in real time. Flow accelerated lysis, but significantly only at the highest shear stress: The average lysis front velocity was 3 x 10(-5) cm/s at 41 dyn/cm(2) vs. almost half of that at the lower shear stresses. Confocal microscopy showed fibrin fibers dissolving only in a narrow region close to the surface when permeation velocity was predicted to be low. A heterogeneous transport-reaction finite element model was used to describe mural fibrinolysis. After scaling the effects of outer and inner convection, inner diffusion, and chemical reactions, a simplified inner diffusion/reaction model was used. Correlation to fibrin lysis data in purified systems dictated higher rates of plasmin(ogen) and tPA adsorption onto fibrin and a decreased catalytic rate of plasmin-mediated fibrin degradation, compared with published parameters. At each shear stress, the model predicted a temporal pattern of lysis of mural fibrin (similar to that observed experimentally), and protease accumulation in a narrow fibrin region and significant lysis inhibition by plasma alpha(2)-antiplasmin (according to the literature). Increasing outer convection accelerated mural fibrinolysis, but the model did not predict the big increase in lysis rate at the highest shear stress. At higher than arterial flows, additional mechanisms not accounted for in the current model, such as fibrin collapse at the fibrin front, may regulate the lysis of mural clots and determine the outcome of thrombolytic therapy.  相似文献   

9.
Proteolytic degradation of fibrin, the major structural component in blood clots, is critical both during normal wound healing and in the treatment of ischemic stroke and myocardial infarction. Fibrin-containing clots experience substantial strain due to platelet contraction, fluid shear, and mechanical stress at the wound site. However, little is understood about how mechanical forces may influence fibrin dissolution. We used video microscopy to image strained fibrin clots as they were degraded by plasmin, a major fibrinolytic enzyme. Applied strain causes up to 10-fold reduction in the rate of fibrin degradation. Analysis of our data supports a quantitative model in which the decrease in fibrin proteolysis rates with strain stems from slower transport of plasmin into the clot. We performed fluorescence recovery after photobleaching (FRAP) measurements to further probe the effect of strain on diffusive transport. We find that diffusivity perpendicular to the strain axis decreases with increasing strain, while diffusivity along the strain axis remains unchanged. Our results suggest that the properties of the fibrin network have evolved to protect mechanically loaded fibrin from degradation, consistent with its function in wound healing. The pronounced effect of strain upon diffusivity and proteolytic susceptibility within fibrin networks offers a potentially useful means of guiding cell growth and morphology in fibrin-based biomaterials.  相似文献   

10.
We tested the hypothesis that in vitro peroxynitrite (ONOO, a product of activated inflammatory cells) may affect fibrinolysis in human blood through the reduction of platelet-related fibrinolysis resistance. It was found that ONOO (25–300 µM) accelerated lysis of platelet-fibrin clots (in PRP) dose-dependently, whereas fibrinolysis of platelet-free clots was slightly inhibited by ≥1000 µM stressor. Concentrations of ONOO affecting the lysis of platelet-rich clots, inhibited clot retraction (CR) in a dose-dependent manner. Thromboelastometry (ROTEM) measurements performed in PRP showed that treatment with ONOO (threshold conc. 100 µM) prolongs clotting time, and reduces alpha angle, and clot formation velocity parameters indicating for reduced thrombin formation rate. In PRP, ONOO (threshold conc. 100 µM) reduced the collagen-evoked exposure of phosphatidylserine (PS) on platelets’ plasma membrane, the shedding of platelet-derived microparticles (PMP), and inhibited platelet-dependent thrombin generation (measured in artificial system), dose-dependently. As judged by confocal microscopy, similar ONOO concentrations altered the architecture of clots formed in collagen-treated PRP. Clots formed in the presence of ONOO were less dense and were composed of thicker fibers, which make them more susceptible to lysis. In platelet-depleted plasma, ONOO (up to milimolar concentration) did not alter clot structure. Blockage of PS exposed on platelets resulted in an alteration of clot architecture toward more prone to lysis. ONOO, at lysis-affecting concentrations, inhibited the collagen-evoked secretion of fibrinolytic inhibitors from platelets. We conclude that physiologically relevant ONOO concentrations may accelerate the lysis of platelet-fibrin clots predominantly via downregulation of platelet-related mechanisms including: platelet secretion, clot retraction, platelet procoagulant response, and the alteration in clot architecture associated with it.  相似文献   

11.
Vascular injury triggers two intertwined processes, platelet deposition and coagulation, and can lead to the formation of an intravascular clot (thrombus) that may grow to occlude the vessel. Formation of the thrombus involves complex biochemical, biophysical, and biomechanical interactions that are also dynamic and spatially-distributed, and occur on multiple spatial and temporal scales. We previously developed a spatial-temporal mathematical model of these interactions and looked at the interplay between physical factors (flow, transport to the clot, platelet distribution within the blood) and biochemical ones in determining the growth of the clot. Here, we extend this model to include reduction of the advection and diffusion of the coagulation proteins in regions of the clot with high platelet number density. The effect of this reduction, in conjunction with limitations on fluid and platelet transport through dense regions of the clot can be profound. We found that hindered transport leads to the formation of smaller and denser clots compared to the case with no protein hindrance. The limitation on protein transport confines the important activating complexes to small regions in the interior of the thrombus and greatly reduces the supply of substrates to these complexes. Ultimately, this decreases the rate and amount of thrombin production and leads to greatly slowed growth and smaller thrombus size. Our results suggest a possible physical mechanism for limiting thrombus growth.  相似文献   

12.

Background

Formation of compact and poorly lysable clots has been reported in thromboembolic disorders. Little is known about clot properties in bleeding disorders.

Objectives

We hypothesized that more permeable and lysis-sensitive fibrin clots can be detected in women with heavy menstrual bleeding (HMB).

Methods

We studied 52 women with HMB of unknown cause and 52 age-matched control women. Plasma clot permeability (Ks), turbidity and efficiency of fibrinolysis, together with coagulation factors, fibrinolysis proteins, and platelet aggregation were measured.

Results

Women with HMB formed looser plasma fibrin clots (+16% [95%CI 7–18%] Ks) that displayed lower maximum absorbancy (-7% [95%CI -9 – -1%] ΔAbsmax), and shorter clot lysis time (-17% [95%CI -23 – -11%] CLT). The HMB patients and controls did not differ with regard to coagulation factors, fibrinogen, von Willebrand antigen, thrombin generation markers and the proportion of subjects with defective platelet aggregation. The patients had lower platelet count (-12% [95%CI -19 – -2%]), tissue plasminogen activator antigen (-39% [95%CI -41 – -29%] tPA:Ag), and plasminogen activator inhibitor-1 antigen (-28% [95%CI -38 – -18%] PAI-1:Ag) compared with the controls. Multiple regression analysis upon adjustment for age, body mass index, glucose, and fibrinogen showed that decreased tPA:Ag and shortened CLT were the independent predictors of HMB.

Conclusions

Increased clot permeability and susceptibility to fibrinolysis are associated with HMB, suggesting that altered plasma fibrin clot properties might contribute to bleeding disorders of unknown origin.  相似文献   

13.
Incorporation of thrombospondin into fibrin clots   总被引:9,自引:0,他引:9  
Thrombospondin is a major platelet glycoprotein which is released from platelets during blood coagulation. We examined the interaction of thrombospondin with polymerizing fibrin. Thrombospondin, purified from human platelets and labeled with 125I, became incorporated into clots formed from both plasma and purified fibrinogen. Plasma clots contained somewhat less thrombospondin than clots formed from equivalent concentrations of fibrinogen. In plasma clots and fibrin clots formed in the presence of factor XIII, thrombospondin was cross-linked in the clot; thrombospondin in the supernatant remained largely monomeric. Cross-linking of thrombospondin by factor XIII, however, only slightly increased the amount of thrombospondin which was incorporated into the clot. In contrast, incorporation of 125I-fibronectin into clots was dependent upon cross-linking. Most of the incorporation of 125I-thrombospondin occurred during fibrin polymerization as judged by parallel studies of the incorporation of 125I-fibrinogen. The amount of thrombospondin incorporated into a clot was directly related to thrombospondin concentration and was only weakly dependent on fibrinogen concentration. Incorporation was not saturated at thrombospondin:fibrin (mol/mol) ratios as high as 2/1. Thrombospondin, however, modified the final structure of fibrin clots in a concentration-dependent manner as monitored by opacity. When tryptic digests of 125I-thrombospondin were studied, the 270-kilodalton core became incorporated into fibrin whereas the 30-kilodalton heparin binding fragment was excluded. These results indicate that thrombospondin specifically co-polymerizes with fibrin during blood coagulation and may be an important modulator of clot structure.  相似文献   

14.
Association of fibrin with the platelet cytoskeleton   总被引:2,自引:0,他引:2  
We have previously postulated that surface membrane proteins become specifically associated with the internal platelet cytoskeleton upon platelet activation (Tuszynski, G.P., Walsh, P.N., Piperno, J., and Koshy, A. (1982) J. Biol. Chem. 257, 4557-4563). Four lines of evidence are in support of this general hypothesis since we now show that platelet surface receptors for fibrin become specifically associated with the platelet Triton-insoluble cytoskeleton. 1) Fibrin was detected immunologically in the washed Triton-insoluble cytoskeletons of thrombin-activated platelets under conditions where fibrin polymerization and resultant precipitation was blocked with Gly-Pro-Arg-Pro, a synthetic peptide that inhibits polymerization of fibrin monomer. 2) Radiolabeled fibrin bound to thrombin-activated platelets and became associated with the cytoskeleton. 3) The amount of radiolabeled fibrin bound to thrombin-activated thrombasthenic platelets and their cytoskeletons amounted to about 20% of the fibrin bound to thrombin-activated control platelets and their cytoskeletons. 4) The association of fibrin with cytoskeletons and with the platelet surface was nearly quantitatively blocked by an antibody prepared against cytoskeletons (anti-C), an antibody against isolated membranes of Pronase-treated platelets (anti-M1), and a monoclonal antibody to the platelet surface glycoprotein complex, GPIIb-GPIII (anti-GPIII). These antibodies blocked ADP and thrombin-induced platelet aggregation as well as thrombin-induced clot retraction. Analysis of the immunoprecipitates obtained with anti-C, anti-M1, and anti-GPIII from detergent extracts of 125I-surface labeled platelets revealed that these antibodies recognized GPIIb-GPIII. These data suggest that thrombin activation of platelets results in the specific association of fibrin with the platelet cytoskeleton, that this association may be mediated by the GPIIb-GPIII complex, and that these mechanisms may play an important role in platelet aggregation and clot retraction induced by thrombin.  相似文献   

15.
Fibrin clot structure is highly dependent on factor XIII activity. Activated FXIII catalyzes the formation of the peptide bonds between the gamma and alpha chains in noncovalently bound fibrin polymers and incorporates various adhesive and antifibrinolytic proteins into the final fibrin clot. In the absence of activated FXIII, clots are unstable and susceptible to fibrinolysis. Several studies have examined the effects of FXIII polymorphisms on final fibrin clot structure and clinical thrombotic risk. The Val34Leu FXIII polymorphism is associated with increased activation by thrombin. In the presence of saturating thrombin concentrations, however, FXIIIa specific enzyme activity is not affected by genetic polymorphisms. Fibrin clots formed in the presence of the FXIII 34Leu polymorphisms do tend to be thinner and less porous, however. The effects of prothrombin concentrations on clot structure have suggested that thinner clots are more resistant to fibrinolysis and associated with increased thrombotic risk. Most clinical studies of 34Leu FXIII carriers, however, have demonstrated a lower incidence of both venous and arterial thrombosis in carriers of the mutant allele compared to Val/Val carriers. One recent study has suggested that the interactions between FXIII phenotype and plasma fibrinogen concentrations significantly influence clinical thrombotic risk.  相似文献   

16.
Low concentrations of actin filaments (F-actin) inhibit the rate and extent of turbidity developed during polymerization of purified fibrinogen by thrombin. Actin incorporates into the fibrin clot in a concentration-dependent manner that does not reach saturation, indicating nonspecific trapping of actin filaments in the fibrin network. Actin does not retard activation of fibrinogen by thrombin, but rather the alignment of fibrin protofibrils into bundles which constitute the coarse clot. In contrast, equivalent F-actin concentrations have little or no effect on the turbidity of plasma clots. The difference is attributed to the presence of a plasma protein, gelsolin, that severs actin filaments. Purified gelsolin greatly reduces the effect of F-actin on the turbidity of a pure fibrin clot and decreases the fraction of actin incorporated by the clot. A calculation of the extent to which the gelsolin concentrations used in these experiments reduce the fraction of actin filaments which are long enough to impede each other's rotational diffusion indicates that it is the overlapping actin filaments which retard the association of fibrin protofibrils. The findings suggest that one role for the F-actin depolymerizing and particularly actin severing activities in blood is to prevent actin filaments released by tissue injury from interfering with the formation of coarse fibrin clots.  相似文献   

17.
Molecular basis of fibrin clot elasticity   总被引:3,自引:0,他引:3  
Blood clots must be stiff to stop hemorrhage yet elastic to buffer blood's shear forces. Upsetting this balance results in clot rupture and life-threatening thromboembolism. Fibrin, the main component of a blood clot, is formed from molecules of fibrinogen activated by thrombin. Although it is well known that fibrin possesses considerable elasticity, the molecular basis of this elasticity is unknown. Here, we use atomic force microscopy (AFM) and steered molecular dynamics (SMD) to probe the mechanical properties of single fibrinogen molecules and fibrin protofibrils, showing that the mechanical unfolding of their coiled-coil alpha helices is characterized by a distinctive intermediate force plateau in the systems' force-extension curve. We relate this plateau force to a stepwise unfolding of fibrinogen's coiled alpha helices and of its central domain. AFM data show that varying pH and calcium ion concentrations alters the mechanical resilience of fibrinogen. This study provides direct evidence for the coiled alpha helices of fibrinogen to bring about fibrin elasticity.  相似文献   

18.
Concanavalin A was employed to study the role of platelet membrane glycoproteins in platelet-fibrin interactions during clot formation. A rheological technique was used to study the interactions, measuring the clot rigidity and platelet contractile force simultaneously during the formation of network structure. Concanavalin A lowered the clot rigidity and contractile force of a platelet-rich plasma clot by a small extent. Plasma glycoproteins probably compete with platelet membranes for concanavalin A binding in platelet-rich plasma. Both native concanavalin A (tetrameric) and succinyl concanavalin A (dimeric) lowered the clot rigidity and contractile force of a washed platelet-fibrin clot dramatically, almost down to those values found for fibrin clots. Inhibition studies with alpha-methyl-D-mannoside indicated that the concanavalin A effects were specific for the concanavalin A binding capacity to platelets. The effects of native concanavalin A on platelet-fibrin clots were only partially reversible, while the succinyl concanavalin A effects were completely reversible. The observed concanavalin A effects are probably mainly due to concanavalin A binding to platelet membrane glycoproteins. The concanavalin A binding site appears to play an important role in the fibrin binding to platelets.  相似文献   

19.
Streptokinase (SK) is a potent clot dissolver but lacks fibrin clot specificity as it activates human plasminogen (HPG) into human plasmin (HPN) throughout the system leading to increased risk of bleeding. Another major drawback associated with all thrombolytics, including tissue plasminogen activator, is the generation of transient thrombin and release of clot-bound thrombin that promotes reformation of clots. In order to obtain anti-thrombotic as well as clot-specificity properties in SK, cDNAs encoding the EGF 4,5,6 domains of human thrombomodulin were fused with that of streptokinase, either at its N- or C-termini, and expressed these in Pichia pastoris followed by purification and structural-functional characterization, including plasminogen activation, thrombin inhibition, and Protein C activation characteristics. Interestingly, the N-terminal EGF fusion construct (EGF-SK) showed plasmin-mediated plasminogen activation, whereas the C-terminal (SK-EGF) fusion construct exhibited ‘spontaneous’ plasminogen activation which is quite similar to SK i.e. direct activation of systemic HPG in absence of free HPN. Since HPN is normally absent in free circulation due to rapid serpin-based inactivation (such as alpha-2-antiplasmin and alpha-2-Macroglobin), but selectively present in clots, a plasmin-dependent mode of HPG activation is expected to lead to a desirable fibrin clot-specific response by the thrombolytic. Both the N- and C-terminal fusion constructs showed strong thrombin inhibition and Protein C activation properties as well, and significantly prevented re-occlusion in a specially designed assay. The EGF-SK construct exhibited fibrin clot dissolution properties with much-lowered levels of fibrinogenolysis, suggesting unmistakable promise in clot dissolver therapy with reduced hemorrhage and re-occlusion risks.  相似文献   

20.
Measurements of small oscillatory deformations of a fibrin clot by axial motion of a rod in a closed tube reveal an anomalous mechanical loss due to permeation of fluid through the clot structure. The Darcy constant for permeation can be calculated from data at the frequency where the apparent storage and loss shear moduli are equal, without the necessity of measurements at much lower frequencies as previously employed. From the Darcy constant, the average number of fibrin monomer units (v) per cross-section of a fibrous element of the clot can be calculated; it ranges from 4 to several hundred. In the range of fibrin concentration(c) from 3 to 14 milligrams, v is approximately proportional to c-2 for clots of coarse structure and to c-0.5 for clots of fine structure.  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