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.
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   

2.
The proteolytic action of Arvin on human fibrinogen   总被引:10,自引:2,他引:10       下载免费PDF全文
1. Human fibrinogen was subjected to proteolysis by enzyme preparations (clinical Arvin and IRC-50 Arvin) from the venom of Agkistrodon rhodostoma. 2. IRC-50 Arvin releases three peptides from fibrinogen, and these were identified as fibrinopeptides AP, AY and A. 3. The less purified ;clinical' Arvin releases, in addition to fibrinopeptides AP, AY and A, small amounts of two heptapeptides derived from fibrinopeptides AP and A, probably because it contains another enzyme as well as Arvin. 4. No fibrinopeptide B is released by either Arvin preparation. 5. Thus, although Arvin is known to differ from ;reptilase' from Bothrops jararaca in that it does not activate the enzyme that cross-links fibrin (fibrin-stabilizing factor), it is identical with reptilase with respect to the peptides that it liberates from fibrinogen.  相似文献   

3.
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   

4.
The time dependence of the release of fibrinopeptides from fibrinogen was studied as a function of the concentration of fibrinogen, thrombin, and Gly-Pro-Arg-Pro, an inhibitor of fibrin polymerization. The release of fibrinopeptides during fibrin assembly was shown to be a highly ordered process. Rate constants for individual steps in the formation of fibrin were evaluated at pH 7.4, 37 degrees C, gamma/2 = 0.15. The initial event, thrombin-catalyzed proteolysis at Arg-A alpha 16 to release fibrinopeptide A (kcat/Km = 1.09 X 10(7) M-1s-1) was followed by association of the resulting fibrin I monomers. Association of fibrin I was found to be a reversible process with rate constants of 1 X 10(6) M-1s-1 and 0.064 s-1 for association and dissociation, respectively. Assuming random polymerization of fibrin I monomer, the equilibrium constant for fibrin I association (1.56 X 10(7) M-1) indicates that greater than 80% of the fibrin I protofibrils should contain more than 10 monomeric units at 37 degrees C, pH 7.4, when the fibrin I concentration is 1.0 mg/ml. Association of fibrin I monomers was shown to result in a 6.5-fold increase in the susceptibility of Arg-B beta 14 to thrombin-mediated proteolysis. The 6.5-fold increase in the observed specificity constant from 6.5 X 10(5) M-1s-1 to 4.2 X 10(6) M-1s-1 upon association of fibrin I monomers and the rate constant for fibrin association indicates that most of the fibrinopeptide B is released after association of fibrin I monomers. The interaction between a pair of polymerization sites in fibrin I dimer was found to be weaker than the interaction of fibrin I with Gly-Pro-Arg-Pro and weaker than the interaction of fibrin I with fibrinogen.  相似文献   

5.
Concanavalin A dimer interacts with fibrinogen and soluble fibrin at pH 5.2 Analysis of the binding data shows that there are in both cases four binding sites per molecule and that the dissociation constant does not change by removal of fibrinopeptides A and B. Ultracentrifugal studies shows that no aggregates of fibrinogen or fibrin are formed through concanavalin A binding and that up to four molecules of concanavalin A dimer can be bind to one molecule of fibrinogen or fibrin. These results imply that the four carbohydrate chains in the molecule are accessible to concanavalin A dimer. There is a diminution in the coagulation of fibrinogen by thrombin at low relative lectin concentrations and an increase at high concentrations. However, the lectin always favours the aggregation of fibrin monomers and does not have any inhibitory effect on the release of fibrinopeptides. We conclude that the electric charge in the neighbourhood of the carbohydrate in both chains, Bβ and γ plays an important role in the attraction between monomeric fibrin and fibrinogen-monomeric fibrin. The different effect of concanavalin A on the coagulation, depending on the relative concentration of the lectin, would be the result of the screening of this electric charge favouring either the interaction of fibrinogen-monomeric fibrin or the polymerization of monomeric fibrin.  相似文献   

6.
The active thrombin is formed in the blood stream when the blood coagulation system is activated. It attacks fibrinogen, splits off two fibrinopeptides A and fibrinogen is transformed into des-AA fibrin which is able to polymerize spontaneously forming protofibrils. At high thrombin concentration the enzyme splits off two fibrinopeptides B and des-AA fibrin units are transformed into des-AABB fibrin. These two forms of fibrin are widely used in the biological experiments. However des-AA fibrin is obtained usually from fibrinogen using the snake poisons (such as reptilase). Des-AA fibrin was obtained also by physiological enzyme thrombin, but that des-AA fibrin samples had the contamination of des-AABB fibrin. At the present paper we have described the method of the des-AA fibrin preparation by thrombin without any contamination of des-AABB fibrin.  相似文献   

7.
INTEREST has focused recently on the clinical use of proteolytic enzymes similar in properties to thrombin which can directly cleave fibrinogen. Potentially the most important are arvin, derived from the venom of Agkistrodon rhodostoma and reptilase, isolated from the venom of Bothrops atrox. These only release fibrinopeptide A from fibrinogen1–3, whereas thrombin cleaves fibrinopeptides A and B from fibrinogen to form fibrin. Thrombin also activates fibrin stabilizing factor (FSF) which introduces amide bonds between the subunits of soluble fibrin4. FSF rapidly forms covalent links between pairs of γ(C)-chains giving γ(C)-dimers and in a slower reaction α(A)-chains are linked to produce high molecular weight polymers5. Although reptilase, like thrombin, activates FSF6, arvin apparently does not, which would explain why the fibrin formed by arvin seems to be more friable than that produced by thrombin or reptilase7.  相似文献   

8.
The impact of delayed fibrinopeptide-A release on polymerization and structure of fibrin gels was studied utilizing a heterozygously transmitted variant fibrinogen. An arginine to histidine substitution at position 16 of the alpha chain of the abnormal fibrinogen delayed release of an abnormal fibrinopeptide-A (A) by thrombin and completely blocked release of A by reptilase. When clotted with thrombin, patient fibrin formed more slowly than normal fibrin, but clottability was normal and gel fiber mass/length ratios were decreased less than 10%. Gels formed with reptilase clotted slowly, demonstrated reduced clottability, but had normal fiber mass/length ratios. Reptilase clotted the normal but not the variant component of the patient fibrinogen. Thrombin-induced cleavage of fibrinopeptide-B prior to A occurred in these experiments, but polymerization of this species beyond trimers has been reported to be minimal under the conditions used. With time, A is removed by thrombin resulting in the slow production of normal fibrin monomer from the abnormal component. These monomers subsequently polymerize. The minimal change in gel fiber size caused by slow A release implies that fibrin fiber size is primarily a function of ionic environment and not of the sequence of peptide release.  相似文献   

9.
Mixed vesicles of dimyristoylphosphatidylcholine (DMPC) and a polymerizable lipid containing one diene group per chain are studied by freeze fracture electron microscopy and by the photobleaching (fluorescence recovery after photobleaching) technique. Large thin-walled vesicles of some micron in diameter become more stable after photochemical polymerization. Before polymerization bilayers of the diene lipid exhibit a liquid crystal-to-gel transition at Tg = 31 degrees C. Upon polymerization the transition remains but shifts to a slightly higher temperature (Tg* = 34 degrees C). The transitions in both cases are accompanied by a freezing in of the lateral mobilities. The mixed vesicle exhibits lateral phase separation after polymerization. Before polymerization the two lipids appear miscible at all compositions in the fluid state and at DMPC concentrations at or below 50 mol % in the solid state. After polymerization a two-dimensional solution of the polymer in DMPC is obtained at T greater than Tg*, while lateral phase segregation into DMPC-rich domains and patches of the polymer is observed at T less than Tg*. The domain structure appears identical irrespective of whether polymerization is performed at T greater than Tg or at T less than Tg. A typical value of the diameter of the polymerized lipid domains (approximately 400 A) indicates a rather small aggregation number (N less than 100 monomers). The lateral diffusion coefficient in butadiene-lipid bilayers only decreases from D1 = 3.10(-7) cm2/s to D1 = 8.10(-8) cm2/s (that is by a factor of 4) upon polymerization. This is consistent with the freeze fracture finding of a small aggregation number.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
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   

11.
E Mihalyi  J W Donovan 《Biochemistry》1985,24(14):3443-3448
When clotting is effected by thrombin in the presence of calcium, the endotherm for the D nodules of fibrinogen broadens significantly and then becomes narrow again, while increasing in size. Clotting effected by the snake venom enzyme Ancrod, which releases only the A fibrinopeptides from the E nodule, shows only the broadening of the D endotherm. Accordingly, significant interactions of the D nodules of fibrinogen become possible only when the B fibrinopeptides of the E nodule are released on clotting. When calcium present during clotting is removed from the fibrin clot with ethylenediaminetetraacetic acid, the endotherm for the D nodules of fibrin shows nearly complete reversal if clotting was effected with Ancrod but appears to be divided into two endotherms if clotting was effected with thrombin. At neutral pH, new endotherms were observed for fibrinogen in the temperature range 105-140 degrees C.  相似文献   

12.
Multiple factors affect the thrombin-catalyzed conversion of fibrinogen to fibrin, including: fibrinopeptide (FPA and FPB) release leading to exposure of two types of polymerization domains ("A" and "B," respectively) in the central portion of the molecule, and exposure of a noncatalytic "secondary" thrombin-binding site in fibrin. Fibrinogen containing the FPA sequence but lacking the B beta 1-42 sequence ("des-(B beta 1-42)-fibrinogen"), was compared to native fibrinogen (containing both FPA and FPB) to investigate the role played by B beta 1-42 in the polymerization of alpha-fibrin (i.e. fibrin lacking FPA), to compare reptilase and thrombin cleavage of FPA from fibrinogen, and to explore the location and function of the secondary thrombin-binding site. Electron microscopy of evolving polymer structures (mu, 0.14; pH 7.4) plus turbidity measurements, showed that early thin fibril formation as well as subsequent lateral fibril associations were impaired in des-(B beta 1-42)-alpha-fibrin, thus indicating that the B beta 1-42 sequence contributes to the A polymerization site. Reptilase-activated des-(B beta 1-42)-alpha-fibrin polymerized even more slowly than thrombin-activated des-(B beta 1-42)-alpha-fibrin, differences that disappeared when repolymerization of preformed fibrin monomers was carried out. Since existing data indicate that thrombin releases FPA in a concerted manner, resulting in relatively rapid evolution of fully functional divalent alpha-fibrin monomers, it can be inferred that delayed fibrin assembly of reptilase fibrin is due to slower formation of divalent alpha-fibrin monomers. Thrombin-activated des-(B beta 1-42)-alpha-fibrin polymerized more rapidly at low ionic strength (mu, 0.04) than did native alpha,beta-fibrin, a reversal of their behavior at physiological ionic strength (mu, 0.14). Concomitant measurement of FPA release revealed modest slowing of release at low ionic strength from des-(B beta 1-42)-fibrinogen (t1/2, 36.5 versus 21.5 min) and marked slowing from native fibrinogen (t1/2, 138 versus 22.2 min). This behavior correlated with increased thrombin binding to native alpha,beta-fibrin at low ionic strength, coupled with weak thrombin binding to des-(B beta 1-42)-alpha-fibrin, and indicates that secondary thrombin binding plays an important role in regulating thrombin diffusion and catalytic activity. Des-(B beta 1-42)-fibrinogen lacks or has a markedly defective secondary thrombin-binding site, from which we conclude that the B beta 15-42 sequence in fibrin plays a major role in forming or providing this site.  相似文献   

13.
Assembly of fibrin. A light scattering study.   总被引:9,自引:0,他引:9  
Using stopped flow light scattering, we show that assembly of fibrin following activation with non-rate-limiting amounts of thrombin or reptilase occurs in two steps, of which the first is end-to-end polymerization of fibrin monomers to protofibrils and the second is lateral association of protofibrils to fibers, in agreement with Ferry's original proposal. Polymerization is found to proceed as a bimolecular association of bifunctional monomers; the overall rate varies as the inverse first power of the concentration; end-to-end association of two monomers, of a monomer and an oligomer, and of two oligomers occurs with the same rate constant. The value of the rate constant is 8.2 C 10(5) M-1 s-1 in 0.5 M NaCl, is three times larger in 0.1 M NaCl (0.05 M Tris, pH 7.4), and is the same following activation by reptilase and by thrombin. The onset of growth of fibers from protofibrils takes 12 times longer in 0.5 than in 0.1 M salt, i.e. thick fibers ("coarse" gels) form from short protofibrils, and thin fibers ("fine" gels) form from longer protofibrils. Jumps of salt concentration at times when protofibrils, but not fibers, have formed result in immediate growth of thick fibers at low salt from long protofibrils formed at high salt. The rate of fiber growth in these experiments varies as the inverse first power of the concentration. 3the instant of gelation (formation of a network of fibers) falls in the later half of the time during which the scattering rises due to fiber growth; the rise of gel rigidity after gelation is found to continue beyond the end of this period. Jumps from low to high salt result in retention of whatever fibers have formed at low salt and a very small additional increase of the scattering due to further fiber growth at high salt. From a variety of evidence, we conclude that the properties of fibrin are determined by kinetics and not equilibria of assembly steps. Results obtained here agree with the following scheme of fibrin assembly: monomers polymerize to protofibrils; long protofibrils associate laterally to fibers; occasionally a long protofibril associates with two different fibers to form an interfiber connection; fiber growth does not reverse to yield stabler, more compact, structures and terminates in formation of a network of fibers. The typical delay of fiber growth is the time during which protofibrils form from monomers. Measurements at rate-limiting concentrations of thrombin have allowed estimation of turnover rates of fibrinopeptides that agree with kinetic parameters obtained with direct assay of fibrinopeptide. Release of fibrinopeptide B causes more rapid fiber formation. Addition of thrombin after activation by reptilase, at a time when protofibrils, but not fibers, have formed, is followed rapidly by fiber formation; this proves that thrombin readily removes fibrinopeptide B from protofibrils. On the basis of these new results and earlier work (in particular, Blomb?ck, B., Hessel, B., Hogg, D., and Therkildsen, L...  相似文献   

14.
The polymerization of fibrin induced by the enzyme thrombin was investigated in the pregelation phase by combining measurements of the release of the fibrinopeptides A with static and dynamic light-scattering at scattering angles ranging from 5°–150°. Without making any assumptions about the polymer distribution, one is led to the following conclusions: The aggregates are rodlike; the Flory-Stockmayer model is to be preferred over the percolation model, i.e., cyclic bonds are infrequent; in the early stages the experiments indicate a functionality f = 2 (number of reactive binding sites per monomer) that increases with increasing release of fibrinopeptides A, eventually approaching the value f = 4; the number of binding sites involved in a bond is about twice the number of the released fibrinopeptides A; and in the polymers the monomer units aggregate end-to-end in the very early stage and then in a staggered overlap.  相似文献   

15.
J Wilf  A P Minton 《Biochemistry》1986,25(11):3124-3133
Oligomer formation in fibrinogen solutions following addition of thrombin was studied by addition of thrombin inhibitor at various times subsequent to thrombin, followed by size-exclusion chromatography (SEC) on a high-performance SEC column capable of resolving species of molecular weights less than or equal to 10(6). Peaks corresponding to species with 1, 2, 3, and 4 or more times the molecular weight of fibrinogen were detected and quantified via nonlinear least-squares curve-fitting procedures. The evolution of each of these peaks with time is well accounted for by a kinetic model in which the predominant component of each oligomeric molecular weight species is a linear complex of fibrinogen and fibrin. The observed predominance of trimeric over dimeric oligomers even at short times suggests that the thrombin-catalyzed release of the two A fibrinopeptides from a single molecule of fibrinogen is highly cooperative.  相似文献   

16.
T W Willis  A T Tu 《Biochemistry》1988,27(13):4769-4777
Crotalus atrox venom contains a variety of proteases which render fibrinogen incoagulable and solubilize fibrin. One of these proteases was purified by using ion-exchange and gel permeation liquid chromatography. The protease, called atroxase, consists of a single nonglycosylated polypeptide chain with a molecular weight of 23,500 and an isoelectric point of pH 9.6. Amino acid analysis indicates atroxase to contain 206 residues with no sulfhydryl groups. Metal analysis found zinc and potassium at 1 mol/mol of protein, and calcium at 0.3 mol/mol of protein. Proteolytic activity is inhibited by ethylenediaminetetraacetate and alpha 2-macroglobulin. Maximal proteolytic activity occurs at pH 9.0 and 55 degrees C. Proteolytic specificity, using oxidized insulin B chain, is similar to that of several hemorrhagic toxins found within the same venom, yet atroxase shows no hemorrhagic activity and exhibits low lethality when tested on Swiss Webster mice. Atroxase, an A alpha, B beta fibrinogenase, cleaves the A alpha chain of fibrinogen first followed by the B beta chain and shows no effect on the gamma chain. The nonspecific action of the enzyme results in the extensive hydrolysis of fibrinogen which releases a variety of fibrinopeptides. Fibrin solubilization appears to occur primarily from the hydrolysis of alpha-polymer and unpolymerized alpha and beta chains. Although crude venom induces platelet aggregation, atroxase demonstrated no ability to induce or inhibit aggregation.  相似文献   

17.
Singly dissected twitch fibers from frog muscle were studied on an optical bench apparatus after micro-injection with the pH indicator dye, phenol red. Dye-related absorbances in myoplasm, denoted by A0(lambda) and A90(lambda), were estimated as a function of wavelength lambda (450 nm less than or equal to lambda less than or equal to 640 nm) with light polarized parallel (0 degrees) and perpendicular (90 degrees) to the fiber axis respectively. At all lambda, A0(lambda) was slightly greater than A90(lambda), indicating that some of the phenol red molecules were bound to oriented structures accessible to myoplasm. The phenol red "isotropic" signal, [A0(lambda) + 2A90(lambda)]/3, a quantity equal to the average absorbance of all the dye molecules independent of their orientation, had a spectral shape that was red-shifted by approximately 10 nm in comparison with in vitro dye calibration curves measured in 140 mM KCl. The red-shifted spectrum also indicates that some phenol red molecules were bound in myoplasm. A quantitative estimate of indicator binding was obtained from measurements of the dye's apparent diffusion constant in myoplasm, denoted by Dapp. The small value of Dapp, 0.37 x 10(-6) cm2 s-1 (at 16 degrees C), can be explained if approximately 80% of the dye was bound to myoplasmic sites of low mobility. To estimate the apparent myoplasmic pH, denoted by pHapp, the isotropic absorbance of phenol red was fitted by in vitro calibration spectra. pHapp was found to be independent of dye concentration (0.2-2 mM), but varied widely (range, 6.8-7.5; mean value, 7.17) among fibers judged from functional characteristics to be normal. When fibers were subjected to acid or alkaline loads by exposure to Ringer's solution containing, respectively, dissolved CO2 or NH3, the changes in pHapp were in agreement with those expected from pH micro-electrode studies. It is concluded that in spite of the several indications for the presence of bound phenol red inside muscle cells, the pHapp signal from the indicator is useful for monitoring changes in myoplasmic pH in response to physiological and pharmacological manipulations.  相似文献   

18.
Mihalyi E 《Biophysical chemistry》2004,112(2-3):131-140
Calcium binding curves of human and bovine fibrinogen were obtained by using a calcium sensitive electrode. The two were identical and showed 2 high, 2-3 medium and more than 15 low affinity sites. Differential scanning calorimetry at neutral pH demonstrated the presence of the D and E domains of fibrinogen; however, at pH 3.5 the D-domain was split into two. The presence of the subdomains was demonstrated also by digestion by pepsin at this pH. Combination of digestion of fibrinogen and of its fragments with different enzymes and temperatures identified up to 12 subdomains in the original molecule. Clotting of fibrinogen by thrombin at pH 7.0 was investigated also by differential scanning calorimetry. In the absence of Ca2+ clotting elicited a 40% increase in the enthalpy of thermal denaturation of the D domain of fibrinogen, but the position of the peak increased only by 0.4 degrees C. However, with clotting in the presence of 10(-3) M calcium the former increased by 70-75% and the latter by 11.0 degrees C, while these parameters of the E-domain remained unchanged. Changes of bound calcium during clotting were also measured with the calcium sensitive electrode. These had to be corrected, because the drop in free calcium was partly compensated by release of some calcium that was already bound to fibrinogen. Log of the half time of calcium uptake plotted against log thrombin concentration indicated a first order process with respect to thrombin concentration, moreover, the rate determined corresponded to that of the conformation change measured by calorimetry. The calcium uptake was correlated with release of the fibrinopeptides. Release of fibrinopeptide B follows parallel to binding of calcium and that of fibrinopeptide A is about fourfold faster. Polymerization and formation of thick bundles of fibrin is connected with release of fibrinopeptide A. Clotting with Ancrod, an enzyme that releases only fibrinopeptide A, showed only minimal binding of calcium. The polymerization inhibiting tetrapeptide Gly-Pro-Arg-Pro also depressed binding of calcium. These data suggest that a calcium-binding site must be in the proximity of the site of release of fibrinopeptide B and of a polymerization site.  相似文献   

19.
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   

20.
The structural basis of the wide variability of the physical properties of fibrin clots and the process of assembly of the clot were investigated by electron microscopy of fibers formed under various ionic conditions. In addition, highly specific proteolytic enzymes from different snake venoms were used to remove selectively only the A (batroxobin) or the B (venzyme) fibrinopeptides from fibrinogen, in contrast to thrombin, which removes both pairs. Fibers produced by cleavage of only the B fibrinopeptides displayed a characteristic band pattern indistinguishable from that of fibers formed upon removal of either the A fibrinopeptides alone or of both pairs. Computer modeling studies suggest that there is a unique molecular packing that gives rise to this fibrin band pattern. These findings imply that the release of either fibrinopeptide triggers similar modes of aggregation; the intermolecular binding sites can be localized to particular molecular domains. The diameters of fibers formed with each condition of enzyme, pH, salt concentration, and temperature were measured from electron micrographs. All fibers, except for those produced at both high ionic strength and pH, had about the same average diameter of 85 +/- 13 nm. The degree of lateral aggregation of the fibers themselves varied greatly, however; fibers aggregated more readily with cleavage of both pairs of fibrinopeptides and at lower pH and salt concentrations. The formation of such thick fiber bundles increases the stability of the clot and its resistance to proteolytic dissolution.  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