首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
A polysaccharide YCP was prepared from a marine filamentous fungus Keissleriella sp. YS4108, which exhibited as a molecular weight (Mw) of 2.4x10(3) kDa and its three sulfated derivatives (YCP-SL, YCP-SM and YCP-SH) were synthesized, the degree of substitution (DS) of which were determined to be 0.13, 0.99 and 1.3, with the average molecular weight 0.64x10(3), 0.57x10(3) and 0.45x10(3) kDa, respectively. Anticoagulant activity and antiplatelet aggregation activity of these sulfated derivates were evaluated by activated partial thromboplastin time (APTT), prothrombin time (PT), thrombin time (TT) and platelet aggregation assay. The results showed that YCP sulfates significantly prolonged APTT, TT and PT. The derivates showed no effects on thrombin in the presence or in the absence of antithrombin III (AT III) or heparin cofactor II (HC II), while the derivates effectively inhibited factor Xa in the presence of AT III. At the same time, YCP-SH also possessed potent antiplatelet aggregation activity in vitro compared with aspirin. YCP sulfates specifically interfered with different stages of the coagulation cascade, and the anticoagulant activity improved with the increasing DS and decreased Mw.  相似文献   

2.
The molecular interactions between components of the heparin-catalyzed antithrombin III/thrombin reaction were investigated by light scattering. When heparin was added to antithrombin III, the molecular weight increased to a maximum and then decreased to that of a 1:1 (antithrombin III X heparin) complex. The initial molecular weights at low heparin to antithrombin III ratios were consistent with the formation of a 2:1 (antithrombin III X heparin) complex in which only one antithrombin III molecule had undergone the conformational change measured by protein fluorescence enhancement. The peak molecular weight never reached that of a complete 2:1 complex. This behavior was observed for bovine and human antithrombin III in the presence of both unfractionated heparin and high molecular weight-high affinity heparin. Pentosane polysulfate also caused some multiple associations. Bovine antithrombin III and thrombin formed a 1:1 complex that underwent further aggregation within minutes, while the human proteins did not aggregate on this time scale after forming the 1:1 complex. In the presence of stoichiometric amounts of heparin, the bovine proteins formed an initial complex of Mr = 230,000 (corresponding to a dimer of heparin-antithrombin III-thrombin) which underwent further aggregation. The human proteins, however, formed a 1:1 (antithrombin III X thrombin) initial complex in the presence of heparin, followed by aggregation. These interactions of thrombin and antithrombin with heparin suggest complex interactions that could relate to heparin function.  相似文献   

3.
The inhibition of thrombin by antithrombin III (AT III) and heparin has been studied in pure systems to determine the kinetics of inhibition during human prothrombin activation. The present study shows that prothrombinase-catalyzed prothrombin activation resulted in the generation of thrombin and meizothrombin(des F1). In the absence of heparin the second-order rate constants of the inactivation of both thrombin and meizothrombin(des F1) formed in the reaction mixture appeared to be identical, k = 3.7 X 10(5) M-1 min-1. The rate constant of inhibition of purified thrombin was 6.5 X 10(5) M-1 min-1. In the presence of heparin the decay of the amidolytic activity was biexponential and could be modeled by a four-parameter equation to determine the pseudo first-order rate constants of inhibition as well as the composition of the reaction with respect to the levels of thrombin and meizothrombin(des F1). The ratio of thrombin over meizothrombin(des F1) varied with the initial prothrombin concentration. Heparin catalyzed the AT III inhibition of thrombin but not meizothrombin(des F1) formed during the prothrombin activation. Thrombin, generated by (Xa-Va-phospholipid-Ca2+) was inhibited by AT III/heparin more slowly than purified thrombin, and the saturation kinetics of the inhibition with respect to AT III differed from those found with purified thrombin.  相似文献   

4.
The effect of heparin on the kinetics of inactivation of thrombin by antithrombin III (AT) has been investigated in order to distinguish between two possible mechanisms. Either (1) heparin activates AT to make it a (kinetically) more effective inhibitor, or (2) heparin makes thrombin more susceptible to inhibition by AT. The results were consistent only with mechanism 1. The experimental approach was to premix heparin with either thrombin or AT and then to measure the rate of association of the two proteins in the rapid-mixing stop-flow spectrophotometer. Reactions were followed spectrophotometrically by observing displacement of the dye proflavine from the active site of thrombin as AT binds. Only premixing AT with heparin accelerated the reaction compared to control (no heparin); the observed second-order rate constant was enhanced by a factor of 200–400. Premixing of thrombin with heparin was without effect on the rate of association with AT. If heparin was premixed with both proteins before reaction, the rate was as slow as the control, indicating that heparin bound to thrombin is actually inhibitory to the association of enzyme with activated AT.  相似文献   

5.
The time course of the hemostasis parameters (activated partial thromboplastin time, prothrombin time, thrombin time, fibrinogen, plasminogen, D-dimer, soluble fibrin-monomer complexes (SFMC), as well as antithrombin III (AT III) and protein C (antiplasmin) activities) were estimated in eight male volunteers aged from 20 to 40 years during 21-h bed rest after dehydratation induced by diuretics, which was followed by intravenous infusion of colloid and crystalloid solutions. It was shown that noncompensated dehydratation during bed rest did not cause significant alterations in the procoagulant component of hemostasis. The intravenous infusion of the blood substitute sterofundin (crystalloid) and venofundin (colloid) to the bed-resting subjects resulted in an increase in coagulation within the physiological norm. A decrease in the AT III activity and plasminogen level resulting in an increase in SFMC were observed. The fibrinogen concentration remained stable, which indicated the absence of acute reaction of the body to the experimental effects.  相似文献   

6.
Heparin and heparin fragments in the molecular mass range 1,700-20,000 Da were examined for their ability to accelerate the antithrombin III (AT III)-dependent inhibition of human factor Xa and the prothrombin converting complex (prothrombinase) during human prothrombin activation. The prothrombinase reaction was modeled by a 3-parameter 2-exponential equation to determine the initial rate of prothrombin activation and the pseudo-first order rate constants of inhibition of prothrombinase and in situ generated thrombin activity. The catalytic specific activities of the heparins increased with increasing molecular size for both the inhibition of prothrombinase and factor Xa. A 10-fold increase over the entire Mr range was found. In contrast to results obtained by others (Ellis, V., Scully, M. F., and Kakkar, V. V. (1986) Biochem. J. 233, 161-165; Barrowcliffe, T. W., Havercroft, S. J., Kemball-Cook, G., and Lindahl, U. (1987) Biochem. J. 243, 31-37), all the heparins showed a 5-fold higher rate of inhibition of factor Xa when compared with the inhibition of prothrombinase, indicating that the factor Va-mediated protection of factor Xa from inhibition by AT III/heparin is independent of the molecular size of the heparin. Our original approach has also revealed a hitherto unrecognized phenomenon, namely, in addition to the accelerating effect of the heparins on the rate of formation of the inactive AT III-factor Xa complex, heparins with Mr greater than 4,500 reduce the initial rate of thrombin generation in the presence of AT III in a concentration-dependent way. We hypothesize that the formation of the dissociable ternary AT III-heparin-factor Xa complex results in a (partial) loss of factor Xa activity towards its natural substrate prothrombin.  相似文献   

7.
Heparin was fractionated by affinity chromatography on immobilized antithrombin III followed by gel filtration on Sephadex G-100. Eighteen fractions were obtained ranging in molecular weight from 9,700 to 34,300 as determined by sedimentation equilibrium. The binding stoichiometries of antithrombin III and thrombin interactions with the heparin of these fractions were measured, using changes in intrinsic and extrinsic fluorescence. Catalytic activity also was measured for each of the heparin fractions. As the molecular weight of heparin varied from about 10,000 to 30,000, the average number of antithrombin and thrombin sites/heparin molecule varied from 1.0 to 2.1 and 2.4 to 6.8. In addition, the molar specific activity increased 5.7-fold, an increase which correlated directly with the product of the number of antithrombin III and thrombin molecules bound. Thus as the number of bound molecules increased with increased molecular weight, the rate of reaction/bound antithrombin III increased in proportion to the number of bound thrombin molecules and vice versa. This can be explained by assuming that heparin functions as a template for both proteins, that all bound thrombin and antithrombin III molecules are accessible to each other, and that the rate at which a bound molecule reacts is proportional to the number of molecules of its interacting counterpart bound. These observations and conclusions are similar to those of Hoylaerts et al. (Hoylaerts, M., Owen, W. G., and Collen, D. (1984) J. Biol. Chem. 259, 5670-5677), who demonstrated that the rate at which single molecules of antithrombin III, covalently attached to heparin, react increases as the thrombin binding capacity (chain length) of heparin increases.  相似文献   

8.
Because thrombin has been implicated in sepsis, it has been proposed that antithrombin III (AT III) is beneficial due to its anticoagulatory and antiadhesive effects. Using intravital microscopy, we visualized leukocyte-endothelium interactions in postcapillary venules of the feline mesentery exposed to lipopolysaccharide (LPS). At a concentration of AT III that blocks leukocyte adhesion in postischemic mesentery, we found no role for thrombin in LPS-induced rolling, adhesion and emigration, or microvascular dysfunction. Furthermore, AT III did not attenuate leukocyte-endothelial interactions after tumor necrosis factor-alpha superfusion of the mesentery. In contrast, fucoidan, a selectin inhibitor, prevented almost all LPS-induced rolling and reduced adhesion, emigration, and microvascular dysfunction. In a model of endotoxemia, leukocyte recruitment into mesentery or lungs was unaffected by AT III. Finally, in a human cell system that mimics the flow conditions in vivo, human neutrophils rolled, adhered, and emigrated similar to the feline postcapillary microvessels, and AT III had no effect on leukocyte recruitment induced by LPS. If AT III has beneficial effects in endotoxemia, it is not due to a direct effect upon leukocyte rolling, adhesion, or emigration in postcapillary venules in vivo.  相似文献   

9.
Properties of thrombin- and elastase-modified human antithrombin III   总被引:3,自引:0,他引:3  
P Gettins  B Harten 《Biochemistry》1988,27(10):3634-3639
Proteolytically modified forms of human antithrombin III have been prepared by reaction of native antithrombin with thrombin, human neutrophil elastase, or porcine pancreatic elastase. These forms have two chains disulfide linked and are of the same molecular weight as native antithrombin III. 1H NMR spectroscopy has been used to characterize these proteins and to compare them to one another and to native antithrombin III. The three modified proteins have very similar NMR spectra and histidine residues with identical pH titration parameters, and they undergo the same spectral changes upon binding heparin. They differ from native antithrombin III in all of these respects. In addition, the proteins are much more stable than native antithrombin III. The three modified proteins behave identically as a function of temperature; at 372 K, 44 K above the unfolding temperature for native antithrombin III, the proteins are still folded and possess approximately 70 unexchanged amide protons even after several hours. The unfolding of the heparin binding domain at low concentrations of deuteriated guanidine hydrochloride seen in native thrombin III is absent in the modified forms. It is concluded that the thrombin- and elastase-modified forms of antithrombin have identical structures when allowance is made for the slightly different sites of cleavage by the two types of elastase and by thrombin. This structure is very different from that of native antithrombin III.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
The effects of polyamines on blood coagulation and fibrinolysis in the presence of glycosaminoglycans (GAGs) were examined because it is known that heparin (HP) interacts with polyamines, especially with spermine. Spermine was able to reverse the prolongation of coagulation time of rabbit plasma caused by HP. The effects of various GAGs on thrombin activity in the presence of anti-thrombin III (AT) were then tested using a synthetic substrate. Inhibition of thrombin activity by GAGs was in the order HP > heparan sulfate (HS) > dermatan sulfate (DS) > chondroitin sulfate (CS) approximately hyaluronan (HA). When these GAGs were fully sulfonated, the inhibitory activity of HS, DS, CS and HA, but not HP, became stronger. The effects of GAGs on thrombin activity were reversed by polyamines, in particular spermine. The EC(50) value of spermine for reversal of HP inhibition was 30-50 microM, and the K(d) value of spermine for heparin was 41.1 microM. Analysis by surface plasmon resonance (SPR) indicated that the interaction between AT and HP was weakened by spermine through its binding to HP. The effect of HP on fibrinolysis was then examined. When Glu-plasminogen and tissue-type plasminogen activator (tPA) were used as enzyme source, HP strongly enhanced the plasmin activity and spermine reversed this effect. Analysis by SPR suggests that the structure of the active site of tPA may be changed through the ternary complex formation of tPA, HP and spermine. The results indicate that blood coagulation was enhanced and fibrinolysis was weakened by spermine in the presence of HP.  相似文献   

11.
The antithrombin III (AT III) activity and the AT III concentration were investigated in 62 consecutive patients with acute myocardial infarction (AMI). To identify the reactant pattern of AT III in postaggressive situations, we also determined labile and acute phase proteins. Firstly, 29 patients were nourished orally and then 33 patients were fed by i.v. hyperalimentation (additional caloric intake of approximately 1,000 cal). AT III activities and concentrations as well as prealbumin and retinol-binding protein decreased concomittantly and significantly whereas haptoglobin, C-reactive protein and fibrinogen increased significantly after AMI. The changes cannot be interpreted as being alterations of the haematocrit. The alterations of AT III correlated significantly with the changes of labile proteins but not with the acute phase reactant proteins. The AT III decrease in the postinfarctional phase may promote a prethrombotic state. In In addition it can be concluded from our results that AT III reacts as (nutritive-dependent) labile protein, which is lowered in postaggressive situation and does not increase as an acute phase reactant. This is in accordance with results from recent animal experiments.  相似文献   

12.
Heparin has been shown to accelerate the inactivation of alpha-thrombin by antithrombin III (AT) by promoting the initial encounter of proteinase and inhibitor in a ternary thrombin-AT-heparin complex. The aim of the present work was to evaluate the relative contributions of an AT conformational change induced by heparin and of a thrombin-heparin interaction to the promotion by heparin of the thrombin-AT interaction in this ternary complex. This was achieved by comparing the ionic and nonionic contributions to the binary and ternary complex interactions involved in ternary complex assembly at pH 7.4, 25 degrees C, and 0.1-0.35 M NaCl. Equilibrium binding and kinetic studies of the binary complex interactions as a function of salt concentration indicated a similar large ionic component for thrombin-heparin and AT-heparin interactions, but a predominantly nonionic contribution to the thrombin-AT interaction. Stopped-flow kinetic studies of ternary complex formation under conditions where heparin was always saturated with AT demonstrated that the ternary complex was assembled primarily from free thrombin and AT-heparin binary complex at all salt concentrations. Moreover, the ternary complex interaction of thrombin with AT bound to heparin exhibited a substantial ionic component similar to that of the thrombin-heparin binary complex interaction. Comparison of the ionic and nonionic components of thrombin binary and ternary complex interactions indicated that: 1) additive contributions of ionic thrombin-heparin and nonionic thrombin-AT binary complex interactions completely accounted for the binding energy of the thrombin ternary complex interaction, and 2) the heparin-induced AT conformational change made a relatively insignificant contribution to this binding energy. The results thus suggest that heparin promotes the encounter of thrombin and AT primarily by approximating the proteinase and inhibitor on the polysaccharide surface. Evidence was further obtained for alternative modes of thrombin binding to the AT-heparin complex, either with or without the active site of the enzyme complexed with AT. This finding is consistent with the ternary complex encounter of thrombin and AT being mediated by thrombin binding to nonspecific heparin sites, followed by diffusion along the heparin surface to a unique site adjacent to the bound inhibitor.  相似文献   

13.
The anticoagulant properties of mast cell product, chondroitin sulphate E   总被引:2,自引:0,他引:2  
The anticoagulant potency in vitro of chondroitin sulphate E has been found to be similar to that of the heparinoids. In purified systems chondroitin sulphate E was shown to be principally an activator of heparin cofactor II. Maximum acceleration of heparin cofactor II:thrombin interaction was 185-fold (9.3 X 10(7) M-1 min-1), antithrombin III:thrombin interaction was 11-fold (4.16 X 10(6) M-1 min-1) and antithrombin III:factor Xa was 146-fold (3.86 X 10(6) M-1 min-1). Chondroitin sulphate E was observed to prolong the thrombin clotting time of fibrinogen in the absence of antithrombin III and heparin cofactor II. The effect appeared to be related to interference in thrombin:fibrinogen interaction rather than in fibrin monomer polymerization.  相似文献   

14.
A S-sens K5 surface acoustic wave biosensor was coupled with mass spectrometry (SAW-MS) for the analysis of a protein complex consisting of human blood clotting cascade factor alpha-thrombin and human antithrombin III, a specific blood plasma inhibitor of thrombin. Specific binding of antithrombin III to thrombin was recorded as a function of time with a S-sens K5 biosensor. Two out of five elements of the sensor chip were used as references. To the remaining three elements coated with RNA anti-thrombin aptamers, thrombin and antithrombin III were bound consecutively. The biosensor measures mass changes on the chip surface showing that 20% of about 400fmol/cm2 thrombin formed a complex with the 1.7-times larger antithrombin III. Mass spectrometry (MS) was applied to identify the bound proteins. Sensor chips with aptamer-captured (1) thrombin and (2) thrombin-antithrombin III complex (TAT-complex) were digested with proteases on the sensor element and subsequently identified by peptide mass fingerprint (PMF) with matrix assisted laser desorption/ionization time-of-flight (MALDI-ToF) mass spectrometry. A significant identification of thrombin was achieved by measuring the entire digest with MALDI-ToF MS directly from the sensor chip surface. For the significant identification of both proteins in the TAT-complex, the proteolytic peptides had to be separated by nano-capillary-HPLC prior to MALDI-ToF MS. SAW-MS is applicable to protein interaction analysis as in functional proteomics and to miniaturized diagnostics.  相似文献   

15.
M R Downing  J W Bloom  K G Mann 《Biochemistry》1978,17(13):2649-2653
Human alpha-thrombin is inhibited by the circulating protease inhibitors alpha1-antitrypsin, antithrombin III, and alpha2-macroglobulin. Kinetic analyses of the inhibitor thrombin interactions were carried out utilizing either fibrinogen or the synthetic substrate Bz-Phe-Val-Arg-p-nitroanilide as substrates to determine residual thrombin activity. These studies demonstrated that the inhibition of thrombin by alpha1-antitrypsin, antithrombin III, and alpha2-macroglobulin followed second-order kinetics. The rate constants for the inhibition of thrombin by alpha1-antitrypsin, antithrombin III, and alpha2-macroglobulin are 6.51 +/- 0.38 x 10(3), 3.36 +/- 0.34 x 10(5), and 2.93 +/- 0.02 x 10(4) M-1 min-1, respectively. Comparison of the second-order rate constants and the normal plasma levels of the three inhibitors demonstrates that, under the in vitro conditions utilized, antithrombin III is five times and alpha2-macroglobulin is one-third as effective as alpha1-antitrypsin in the inhibition of thrombin.  相似文献   

16.
The effect of bovine thrombomodulin on the specificity of bovine thrombin   总被引:8,自引:0,他引:8  
Bovine lung thrombomodulin is purified and used to investigate the basis of the change in substrate specificity of bovine thrombin when bound to thrombomodulin. Bovine thrombomodulin is a single polypeptide having an apparent molecular weight of 84,000 and associates with thrombin with high affinity and rapid equilibrium, to act as a potent cofactor for protein C activation and antagonist of reactions of thrombin with fibrinogen, heparin cofactor 2, and hirudin. Bovine thrombomodulin inhibits the clotting activity of thrombin with Kd less than 2.5 nM. Kinetic analysis of the effect of bovine thrombomodulin on fibrinopeptide A hydrolysis by thrombin indicates competitive inhibition with Kis = 0.5 nM. The active site of thrombin is little perturbed by thrombomodulin, as tosyl-Gly-Pro-Arg-p-nitroanilide hydrolysis and inhibition by antithrombin III are unaffected. Insensitivity of the reaction with antithrombin III is likewise observed with thrombin bound to thrombomodulin on intact endothelium. Antithrombin III-heparin, human heparin cofactor 2, and hirudin inhibit thrombin-thrombomodulin more slowly than thrombin. These effects may arise from a decrease in Ki of the inhibitors for thrombin-thrombomodulin or from changes in the active site not detected by tosyl-Gly-Pro-Arg-p-nitroanilide or antithrombin III. Bovine prothrombin fragment 2 inhibits thrombin clotting activity (Kd less than 7.5 microM) and acts as a competitive inhibitor of protein C activation (Kis = 2.1 microM). The data are consistent with a mechanism whereby thrombomodulin alters thrombin specificity by either binding to or allosterically altering a site on thrombin distinct from the catalytic center required for binding or steric accommodation of fibrinogen, prothrombin fragment 2, heparin cofactor 2, and hirudin.  相似文献   

17.
Heparin, other glycosaminoglycans, and synthetic sulfated polymers have antithrombotic and anticoagulant activities, which may be mediated through a range of interactions with different proteins. A simple, quantitative method has been developed for assessing the affinity of interaction between sulfated polymers and proteins in the liquid phase. This has been used to compare the binding of a range of glycosaminoglycans and other sulfated polymers to antithrombin III and thrombin, a major inhibitor of and a central protease in the coagulation system, respectively. The results are consistent with the binding of naturally occurring glycosaminoglycans to antithrombin III solely through the well-defined antithrombin III-binding pentasaccharide found in heparin, the apparent affinity of a preparation depending upon its content of this pentasaccharide. Highly sulfated synthetic polymers will, however, bind antithrombin III by a second mechanism. The affinity of heparin for thrombin decreased with decreasing molecular weight. However, results obtained with heparan sulfate preparations did not indicate any clear relationship between either molecular weight or sulfate content and thrombin binding, but suggested that there may be an oligosaccharide sequence containing N-sulfate residues which confers high affinity for thrombin. In addition, some of the synthetic sulfated polymers bound thrombin with very high affinity.  相似文献   

18.
We have recently determined the structure of a native fibrinogen at 2.7-Å resolution. Not the least of the hurdles during the many years of this project was growing X-ray-grade crystals from suitably purified proteins. Small, synthetic peptides based on the parts of fibrinogen exposed by the action of thrombin contributed greatly to these experiments. In addition, trimethylamine oxide (TMAO) was found to improve the diffraction of fibrinogen crystals. The history of my interest in fibrinogen and its crystallization can be traced back in part to some early interactions with John Edsall.  相似文献   

19.
Heparin (H) was previously shown to accelerate the inactivation of alpha-thrombin (T) by antithrombin III (AT) primarily by promoting the initial binding of thrombin and AT in a ternary T.AT.H complex intermediate without significantly influencing the subsequent product formation step (Olson, S. T., and Shore, J. D. (1982) J. Biol. Chem. 257, 14891-14895). In the present study, the protein-heparin interactions which contribute to the assembly of the ternary complex intermediate and their linkage were quantitated by equilibrium binding and stopped-flow kinetic studies at pH 7.4, I 0.3, 25 degrees C, using p-aminobenzamidine (P) as a fluorescence probe. Equilibrium binding studies of the AT.H and T.H binary complex interactions monitored by the 40% enhancement in AT fluorescence or the 16-18% quenching of thrombin-bound p-aminobenzamidine fluorescence, respectively, indicated a 100-fold greater affinity of AT for heparin (KAT,H 0.23 microM) as compared to thrombin for heparin (KT,H 35-42 microM). Consistent with this large difference in affinities, rapid kinetic studies indicated that assembly of the ternary complex occurred predominantly as a bimolecular association between the AT.H binary complex and free thrombin. Thus, under pseudo-first order conditions ([AT]o, [H]o much greater than [T]o much less than [P]o), the observed thrombin inactivation rate constant (kobs) exhibited a saturable dependence on [AT]o or [H]o when [H]o much less than KT,H, reflecting a KAT,H (0.25 microM) similar to that directly determined by equilibrium binding. Moreover, competitive inhibition of the reaction by T.H binary complexes was indicated from the hyperbolic decrease in kobs produced by heparin with either high or low affinity for AT or active-site blocked thrombin at concentrations comparable to KT,H. This behavior was consistent with values of KT,H (27-51 microM) similar to those determined directly from equilibrium binding measurements. Comparison of the affinities of the binary protein-heparin interactions with the affinity of thrombin for AT.H complex in the ternary complex measured previously, indicated that the affinity of either protein for heparin was enhanced about 10-fold by the prior binding of the other protein to heparin. This linkage of the protein-heparin interactions implies that the ternary complex will be assembled at thrombin, AT, and heparin concentrations considerably lower than those predicted from previous reaction models which fail to account for this linkage.  相似文献   

20.
In this paper, we describe two simple, label-free, homogenous assays using gold nanoparticles (Au NPs)-one to detect coagulation-related proteins and the other to screen inhibition reactions. The first nanosensor functions on the basis of the fact that thrombin catalyzes fibrinogen to form long-chain fibrins, which then induce aggregation of Au NPs. We applied this sensor to study the interactions of thrombin, inhibitors, cofactors, and antidotes. We further used thrombin-conjugated Au NPs (Thr-Au NPs) to analyze the levels of fibrinogen in plasma samples via fibrinogen-induced aggregation of Thr-Au NPs. The limit of detection (LOD; S/N=3) of this sensor for fibrinogen in plasma was 10nM. The Thr-Au NP probe provided quantitative results for fibrinogen in plasma samples that correlated (R(2)=0.97) with those obtained using a clinical von Clauss clotting rate assay. In addition, the Thr-Au NP-based sensor could be used to monitor thrombin concentrations in plasma samples under physiological conditions. Compared with conventional assays, these label-free assays offer several advantages, such as rapid and simple readout by the naked eye or by UV-vis absorption spectroscopy.  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