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.
Factor VIII (FVIII) is the protein defective in the bleeding disorder hemophilia A. Approximately 5% of hemophilia A patients have normal amounts of a dysfunctional FVIII protein and are termed cross-reacting material (CRM)-positive. The majority of genetic alterations that result in CRM-positive hemophilia A are missense mutations within the A2-domain. To determine the mechanistic basis of the genetic defects within the A2-domain for FVIII function we constructed six mutations within the FVIII cDNA that were previously found in five CRM-positive hemophilia A patients (R527W, S558F, I566T, V634A, and V634M) and one CRM-reduced hemophilia A patient (DeltaF652/3). The specific activity for each mutant secreted into the conditioned medium from transiently transfected COS-1 cells correlated with published data for the patients plasma-derived FVIII, confirming the basis of the genetic defect. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis analysis of immunoprecipitated FVIII protein radiolabeled in COS-1 cells showed that all CRM-positive mutant proteins were synthesized and secreted into the medium at rates similar to wild-type FVIII. The majority of the DeltaF652/3 mutant was defective in secretion and was degraded within the cell. All mutant FVIII proteins were susceptible to thrombin cleavage, and the A2-domain fragment from the I566T mutant had a reduced mobility because of use of an introduced potential N-linked glycosylation site that was confirmed by N-glycanase digestion. To evaluate interaction of FVIII with factor IXa, we performed an inhibition assay using a synthetic peptide corresponding to FVIII residues 558 to 565, previously shown to be a factor IXa interaction site. The concentration of peptide required for 50% inhibition of FVIII activity (IC50) was reduced for the I566T (800 mumol/L) and the S558F (960 mumol/L) mutants compared with wild-type FVIII (> 2,000 mumol/L). N-glycanase digestion increased I566T mutant FVIII activity and increased its IC50 for the peptide (1,400 mumol/L). In comparison to S558F, a more conservative mutant (S558A) had a sixfold increased specific activity that also correlated with an increased IC50 for the peptide. These results provided support that the defects in the I566T and S558F FVIII molecules are caused by steric hindrance for interaction with factor IXa.  相似文献   

2.
One of the functions of von Willebrand factor (vWF) is to serve as a carrier of clotting factor VIII (FVIII). Deficiency of this function results in the von Willebrand disease (vWD) variant type 2N, which resembles hemophilia A. We describe a new sandwich enzyme-linked immunosorbent assay (ELISA) to study the ability of vWF to bind exogenous recombinant FVIII (rFVIII), in which anti-vWF-coated plates are incubated with plasma vWF, followed by exogenous FVIII and a peroxidase-coupled anti-FVIII antibody. Dose-response curves obtained using normal plasma vWF and purified normal vWF revealed a hyperbolic relationship between the optical density and the vWF concentration. The assay allows the quantification of FVIII binding with values expressed in U/dL; 100 U/dL was the amount present in normal plasma. The sensitivity and specificity of the method are demonstrated by its ability to measure binding levels as low as 1 to 2 U/dL and the fact that no FVIII binding was observed using plasma known to contain less than 1 U/dL vWF. To verify the accuracy of the assay, three patients with type 2N vWD with characterized vWF gene mutations were studied using an existing chromogenic assay and our ELISA. A patient who was homozygous for the R53W mutation and had no FVIII binding capacity according to the chromogenic method showed undetectable FVIII binding by ELISA. The remaining two patients, one who was homozygous for the R91Q mutation and one with compound heterozygosity for the R91Q and R53W mutations, showed markedly decreased FVIII binding by the chromogenic method and yielded ELISA values ranging from 4 to 8 U/dL. Therefore, although the two methods produce qualitatively similar results, the ELISA method offers the advantage of allowing quantification of the FVIII binding function. FVIII binding was also analyzed in 20 patients with type 1 vWD; we found a decrease of FVIII binding that was proportionate to the decrease in vWF levels, showing a normal FVIII binding activity/vWF molecule ratio. We define the binding activity measured by this assay as vWF:FVIII binding activity and propose its use in the functional analysis of vWF.  相似文献   

3.
Factor VIII (FVIII) gene rearrangements between the intron 22 F8A sequence in the FVIII gene and either of the two homologous F8A sequences 500 kilobases telomeric to the FVIII gene have recently been found to be responsible for the severe hemophilia A phenotype. We studied 27 patients with severe hemophilia A and 19 with moderate and mild hemophilia, and found FVIII gene rearrangement in 12 patients with severe hemophilia A and none in the patients with moderate or mild disease. Nine of the rearrangements were with the distal telomeric F8A sequence, two were with the proximal sequence, and one had variant distal rearrangement with loss of the FVIII intron 22 F8A band. Two patients with FVIII gene rearrangement had high responding inhibitors, contrary to one previous study suggesting that the presence of a FVIII gene rearrangement is correlated with the absence of inhibitor development. Carrier detection was performed in 17 female relatives, at risk of being carriers, from eight kindreds; 13 were carriers, being heterozygous for the normal and rearranged alleles. The rearrangement assay is particularly useful for carrier determination in families with sporadic cases of hemophilia not helped by linkage analysis with restriction fragment-length polymorphism or intragenic dinucleotide repeat analysis. In all five families with rearrangements and sporadic hemophilia, the mothers of all index patients were found to be carriers.  相似文献   

4.
Factor V (FV) is a large (2,196 amino acids) nonenzymatic cofactor in the coagulation cascade with a domain organization (A1-A2-B-A3-C1-C2) similar to the one of factor VIII (FVIII). FV is activated to factor Va (FVa) by thrombin, which cleaves away the B domain leaving a heterodimeric structure composed of a heavy chain (A1-A2) and a light chain (A3-C1-C2). Activated protein C (APC), together with its cofactor protein S (PS), inhibits the coagulation cascade via limited proteolysis of FVa and FVIIIa (APC cleaves FVa at residues R306, R506, and R679). The A domains of FV and FVIII share important sequence identity with the plasma copper-binding protein ceruloplasmin (CP). The X-ray structure of CP and theoretical models for FVIII have been recently reported. This information allowed us to build a theoretical model (994 residues) for the A domains of human FV/FVa (residues 1-656 and 1546-1883). Structural analysis of the FV model indicates that: (a) the three A domains are arranged in a triangular fashion as in the case of CP and the organization of these domains should remain essentially the same before and after activation; (b) a Type II copper ion is located at the A1-A3 interface; (c) residues R306 and R506 (cleavage sites for APC) are both solvent exposed; (d) residues 1667-1765 within the A3 domain, expected to interact with the membrane, are essentially buried; (e) APC does not bind to FVa residues 1865-1874. Several other features of factor V/Va, like the R506Q and A221V mutations; factor Xa (FXa) and human neutrophil elastase (HNE) cleavages; protein S, prothrombin and FXa binding, are also investigated.  相似文献   

5.
Factor Xa and thrombin (factor IIa) activate factor VIII (fVIII) by different proteolytic pathways. Thrombin cleaves fVIII at Arg372 between the A1 and A2 domains, at Arg740 between the A2 and B domains, and at Arg1689 between the B and A3 domains to form an A1/A2/A3-C1-C2 heterotrimer. We now report a stable porcine fVIIIaXa preparation obtained by Mono S HPLC at pH 6. NH2-terminal sequence analysis of purified subunits of fVIIIaXa revealed that factor Xa cleaves fVIII at Arg219 within the A1 domain and at Arg490 within the A2 domain, as well as at Arg372, Arg740, and Arg1689. Analytical ultracentrifugation of the fVIIIaXa preparation yielded results consistent with a single, 148 kDa species, similar to previous results with fVIIIaIIa [Lollar, P., & Parker, C. G. (1989) Biochemistry 28, 666-674]. Thus, the major species in the fVIIIaXa preparation contains five subunits, including fragments of the A1 and A2 domains that remain noncovalently bound. Fluorescence anisotropy measurements indicated there was no difference in the affinity of fVIIIaXa and fVIIIaIIa for a fluorescent dye-labeled, active-site-blocked derivative of porcine factor IXa. Additionally, the fVIIIaXa preparation bound dye-labeled factor IXa with 1:1 stoichiometry, indicating that all fVIIIaXa molecules in the preparation can bind factor IXa. However, fVIIIaXa had 4-fold less procoagulant activity than fVIIIaIIa. Kinetic analysis of fVIIIa cofactor activity using purified factor IXa and factor X suggested this difference is due to greater activity of fVIIIaIIa relative to fVIIIaXa within the intrinsic fXase complex, rather than a difference in their stabilities.  相似文献   

6.
The classification of factor VIII deficiency, generally used based on plasma levels of factor VIII, consists of severe (<1% normal factor VIII activity), moderate (1% to 4% factor VIII activity), or mild (5% to 25% factor VIII activity). A recent communication described four individuals bearing identical factor VIII mutations. This resulted in a severe bleeding disorder in two patients who carried a normal factor V gene, whereas the two patients who did not display severe hemophilia were heterozygous for the factor V(LEIDEN) mutation, which leads to the substitution of Arg506 --> Gln mutation in the factor V molecule. Based on the factor VIII level measured using factor VIII-deficient plasma, these two patients were classified as mild/moderate hemophiliacs. We studied the condition of moderate to severe hemophilia A combined with the factor V(LEIDEN) mutation in vitro in a reconstituted model of the tissue factor pathway to thrombin. In the model, thrombin generation was initiated by relipidated tissue factor and factor VIIa in the presence of the coagulation factors X, IX, II, V, and VIII and the inhibitors tissue factor pathway inhibitor, antithrombin-III, and protein C. At 5 pmol/L initiating factor VIIa x tissue factor, a 10-fold higher peak level of thrombin formation (350 nmol/L), was observed in the system in the presence of plasma levels of factor VIII compared with reactions without factor VIII. Significant increase in thrombin formation was observed at factor VIII concentrations less than 42 pmol/L (approximately 6% of the normal factor VIII plasma concentration). In reactions without factor VIII, in which thrombin generation was downregulated by the addition of protein C and thrombomodulin, an increase of thrombin formation was observed with the factor V(LEIDEN) mutation. The level of increase in thrombin generation in the hemophilia A situation was found to be dependent on the factor V(LEIDEN) concentration. When the factor V(LEIDEN) concentration was varied from 50% to 150% of the normal plasma concentration, the increase in thrombin generation ranged from threefold to sevenfold. The data suggested that the analysis of the factor V genotype should be accompanied by a quantitative analysis of the plasma factor V(LEIDEN) level to understand the effect of factor V(LEIDEN) in hemophilia A patients. The presented data support the hypothesis that the factor V(LEIDEN) mutation can increase thrombin formation in severe hemophilia A.  相似文献   

7.
Agonist-bound heptahelical receptors activate heterotrimeric G proteins by catalyzing exchange of GDP for GTP on their alpha subunits. In search of an approximation of the receptor-alpha subunit complex, we have considered the properties of A326S Gialpha1, a mutation discovered originally in Gsalpha (Iiri, T., Herzmark, P., Nakamoto, J. M., Van Dop, C., and Bourne, H. R. (1994) Nature 371, 164-168) that mimics the effect of receptor on nucleotide exchange. The mutation accelerates dissociation of GDP from the alphai1beta1gamma2 heterotrimer by 250-fold. Nevertheless, affinity of mutant Gialpha1 for GTPgammaS is high in the presence of Mg2+, and the mutation has no effect on the intrinsic GTPase activity of the alpha subunit. The mutation also uncouples two activities of betagamma: stabilization of the GDP-bound alpha subunit (which is retained) and retardation of GDP dissociation from the heterotrimer (which is lost). For wild-type and mutant Gialpha1, beta gamma prevents irreversible inactivation of the alpha subunit at 30 degreesC. However, the mutation accelerates irreversible inactivation of alpha at 37 degreesC despite the presence of beta gamma. Structurally, the mutation weakens affinity for GTPgammaS by steric crowding: a 2-fold increase in the number of close contacts between the protein and the purine ring of the nucleotide. By contrast, we observe no differences in structure at the GDP binding site between wild-type heterotrimers and those containing A326S Gialpha1. However, the GDP binding site is only partially occupied in crystals of G protein heterotrimers containing A326S Gialpha1. In contrast to original speculations about the structural correlates of receptor-catalyzed nucleotide exchange, rapid dissociation of GDP can be observed in the absence of substantial structural alteration of a Galpha subunit in the GDP-bound state.  相似文献   

8.
Factor V (FV) activation is the result of cleavages at Arg709, Arg1018 and Arg1545 by thrombin or FXa. The relative importance of these cleavages in tissue factor (TF) induced thrombin generation in plasma and in a purified system was elucidated with recombinant FV in which the three sites had been eliminated one by one or in combinations. The mutants were analyzed with a clotting assay using FV-deficient plasma and in a TF induced thrombin generation system using plasma or purified components. Surprisingly, in the standard FV clotting assay, all mutants gave similar clotting activities and the thrombin generation curves obtained with wild-type and thrombin-resistant FV were similar. Differences in clotting activities and thrombin generation patterns between wild-type and thrombin-resistant FV were only observed when lower TF concentrations were used. The thrombin generation curve obtained in plasma containing wt FV was characterized by a short lag phase and a subsequent phase of rapid thrombin generation (propagation phase). The Arg709 to Gln mutation yielded a slightly prolonged lag phase and the rate of thrombin generation during the propagation phase was approximately 5-fold lower than that observed with wt FV. The Arg1018 to Ile mutation only slightly affected the thrombin generation curve, whereas the Arg1545 to Gln mutation yielded a prolonged lag phase and decreased maximum thrombin activity. Thrombin-resistant FV (mutated at all three sites) yielded a prolonged lag phase and poor thrombin generation during the propagation phase. The purified system further demonstrated the importance of the three cleavage sites for rapid and sustained thrombin generation. The results demonstrate that cleavages at positions 709, 1018 and 1545 are not required for assembly of a FXa-FV complex expressing low but significant prothrombinase activity but that all three sites in different ways are important for the creation of a FVa which maximally supports the FXa-mediated activation of prothrombin.  相似文献   

9.
Factor VIII (FVIII) and factor V (FV) are homologous coagulation cofactors sharing a similar domain organization (A1-A2-B-A3-C1-C2) and are both extensively glycosylated within their B-domains. In mammalian cell expression systems, compared with FV, the FVIII primary translation product is inefficiently transported out of the endoplasmic reticulum. Here we show that FVIII is degraded within the cell by a lactacystin-inhibitable pathway, implicating the cytosolic 20 S proteasome machinery. Protein chaperones calnexin (CNX) and calreticulin (CRT) preferentially interact with glycoproteins containing monoglucosylated N-linked oligosaccharides and are proposed to traffic proteins through degradative and/or secretory pathways. Utilizing co-immunoprecipitation assays, intracellular FVIII was detected in association with CNX maximally within 30 min to 1 h following synthesis, whereas FV could not be detected in association with CNX. In contrast, both FVIII and FV displayed interaction with CRT during transit through the secretory pathway. B-domain deleted FVIII significantly reduced the CNX and CRT interaction, indicating the B-domain may represent a primary CNX and CRT interaction site. In the presence of inhibitors of glucose trimming, the interactions of FVIII with CNX, and of FVIII and FV with CRT, were significantly reduced whereas the secretion of FVIII, and not FV, was inhibited. In addition, transfection in a glucosidase I-deficient Chinese hamster ovary cell line (Lec23) demonstrated that both degradation and secretion of FVIII were inhibited, with little effect on the secretion of FV. These results support that CNX and CRT binding, mediated at least in part by the B-domain of FVIII, is required for efficient FVIII degradation and secretion. In contrast, FV does not require CNX interaction for efficient secretion. The results suggest a unique requirement for carbohydrate processing and molecular chaperone interactions that may limit the productive secretion of FVIII.  相似文献   

10.
We have developed a radioimmunoassay for human thrombin using rabbit anti-human thrombin IgG. The assay can measure 2 ng thrombin/ml plasma, 500-fold more sensitive than clotting assays. Human prothrombin is less reactive in the assay than thrombin by at least four orders of magnitude, and there is no demonstrable cross-reactivity with human factor Xa, the clotting factor structurally most similar to thrombin. The assay does not detect thrombin bound to anthithrombin III. Using the assay, we have demonstrated that plasma from 20 normal subjects does not contain detectable thrombin. We measured thrombin generation in clotting blood in polypropylene tubes and observed that thrombin appears (approximately equal to 3 ng/ml) within 45 S-5 min after venipuncture. This material is thrombin, not intermediates of prothrombin activation, since it disappears after addition of heparin, which promotes thrombin antithrombin III complex formation. After a plateau of 2-10 min, there is further thrombin generation, which results in clotting after 15-27 min at a level of 40-50 ng thrombin/ml. The thrombin generated 9-25 min before clotting may activate factors V and VIII and stimulate platelet aggregation and release. In contrast, the cascade hypothesis assigns a role for thrombin only late in blood clotting. Radioimmunoassay of thrombin and other clotting factors will be useful for clinical and physiological studies of blood clotting especially since the assay seems specific for thrombin and is independent of other activities that affect bioassays.  相似文献   

11.
The structural abnormalities and functional characteristics of dysfunctional prothrombin variants in two new kindreds have been determined. Prothrombin Corpus Christi (family 1) was purified and found to have markedly reduced fibrinogen clotting activity, yet normal amidolytic and near-normal platelet aggregating activity. A transition (C to T) at nucleotide position 8885, present in the heterozygous form in affected family members, resulted in the substitution of Cys for Arg 382. This substitution results in the loss of a positive charge within the fibrinogen-binding exosite of thrombin, thus accounting for the observed functional defect. A heterozygous C to T transition was also present at position 19994 in other family members with a hypoprothrombinemic phenotype. This mutation results in the replacement of Gln 541 (CAA) by a premature stop codon (TAA). Prothrombin Dhahran (family 2) was found to have markedly reduced fibrinogen clotting activity, but normal amidolytic activity. Affected family members were found to have a G to A transition at nucleotide position 7312 resulting in the substitution of His for Arg 271. This substitution results in the abolition of a factor Xa cleavage site, yielding meizothrombin rather than thrombin, on activation of prothrombin Dhahran by factor Xa. All but one of the above mutations occur at CpG dinucleotides, thus further supporting the observation of a high incidence of CpG transitions in hereditary dysprothrombinemia. The significant bleeding tendencies of individuals homozygous for prothrombin Dhahran (prothrombin clotting activity 5% to 7%) contrast sharply with the absence of significant chronic bleeding in the proband expressing prothrombin Corpus Christi (prothrombin clotting activity 2%). Our findings underscore the capacity of thrombin to contribute to clinical hemostasis by mechanisms other than its fibrinogen clotting activity.  相似文献   

12.
Recurrent DNA inversions, which disrupt the factor VIII (FVIII) gene, generally occur between a region of intron 22 (int22h) and one of two homologous copies of this region, located 300 to 400 kb telomeric to the FVIII gene. This report describes a patient with severe hemophilia A and a high level inhibitor with atypical hybridization patterns. A Bcl I Southern blot assay was altered to 17.5, 16, and 14 kb. His mother and two out of four aunts tested had normal and abnormal restriction patterns which led to a total of five different fragments, suggesting that they were carriers. The Xba I plus Kpn I restriction fragment-length polymorphism in intron 22 by Southern blotting using the same probe (probe a) yielded the 6.2 kb polymorphic band, with a clearly separated 6.6 kb band from the non-factor VIII region; an alternative int22h hybridization probe (probe x) detected no additional fragment. These results suggest that probe a as well as probe x could recognize an intron-22-sized fragment. This report shows a variation in the number of int22h copies although we could not find the inversion junction.  相似文献   

13.
The binding of factor VIII to von Willebrand factor (vWF) is essential for the protection of factor VIII against proteolytic degradation in plasma. We have characterized the binding kinetics of human factor VIII with vWF using a centrifugation binding assay. Purified or plasma vWF was immobilized with a monoclonal antibody (MoAb RU1) covalently linked to Sepharose (Pharmacia LKB Biotechnology, Uppsala, Sweden). Factor VIII was incubated with vWF-RU1-Sepharose and unbound factor VIII was separated from bound factor VIII by centrifugation. The amount of bound factor VIII was determined from the decrease of factor VIII activity in the supernatant. Factor VIII binding to vWF-RU1-Sepharose conformed to the Langmuir model for independent binding sites with a Kd of 0.46 +/- 0.12 nmol/L, and a stoichiometry of 1.3 factor VIII molecules per vWF monomer at saturation, suggesting that each vWF subunit contains a binding site for factor VIII. Competition experiments were performed with a recombinant vWF (deltaA2-rvWF), lacking residues 730 to 910 which contain the epitope for MoAB RU1. DeltaA2-rvWF effectively displaced previously bound factor VIII, confirming that factor VIII binding to vWF-RU1-Sepharose was reversible. To determine the association rate constant (k(on)) and the dissociation rate constant (k(off)), factor VIII was incubated with vWF-RU1-Sepharose for various time intervals. The observed association kinetics conformed to a simple bimolecular association reaction with k(on) = 5.9 +/- 1.9 x 10(6) M(-1) s(-1) and k(off) = 1.6 +/- 1.2 x 10(-3) s(-1) (mean +/- SD). Similar values were obtained from the dissociation kinetics measured after dilution of preformed factor VIII-vWF-RU1-Sepharose complexes. Identical rate constants were obtained for factor VIII binding to vWF from normal pooled plasma and to vWF from plasma of patients with hemophilia A. The kinetic parameters in this report allow estimation of the time needed for complex formation in vivo in healthy individuals and in patients with hemophilia A, in which monoclonally purified or recombinant factor VIII associates with endogenous vWF. Using the plasma concentration of vWF (50 nmol/L in monomers) and the obtained values for K(on) and K(off), the time needed to bind 50% of factor VIII is approximately 2 seconds.  相似文献   

14.
Patients with severe coagulation factor VIII deficiency require frequent infusions of human factor VIII (hFVIII) concentrates to treat life-threatening hemorrhages. Because these patients are immunologically hFVIII-naive, a significant treatment complication is the development of inhibitors or circulating alloantibodies against hFVIII, which bind the replaced glycoprotein, increase its plasma clearance, and inhibit its activity, preventing subsequent treatments from having a therapeutic effect. A genetic approach toward the induction of immunologic unresponsiveness to hFVIII has the conceptual advantage of a long-term, stable elimination of undesired immune responses against hFVIII. Here, we report that in a factor VIII (FVIII)-deficient mouse model for severe hemophilia A, genetic modification of donor bone marrow cells with a retroviral vector encoding hFVIII, and transplant to hemophiliac mouse recipients, results in the induction of immune tolerance to FVIII in 50% of treated animals after immunization with hFVIII, despite the fact that hFVIII protein or activity is undetectable. In tolerized animals, the titers of anti-hFVIII binding antibodies and of hFVIII inhibitor antibodies were significantly reduced, and there was evidence for hFVIII unresponsiveness in CD4(+) T cells. Importantly, the plasma clearance of hFVIII was significantly decreased in tolerized animals and was not significantly different from that seen in a FVIII-naive hemophiliac mouse. This model system will prove useful for the evaluation of genetic therapies for hFVIII immunomodulation and bring genetic therapies for hFVIII tolerance closer to clinical application for patients with hemophilia A.  相似文献   

15.
The inactivation of factor Va was examined on primary cultures of human umbilical vein endothelial cells (HUVECs), either after addition of activated protein C (APC) or after addition of alpha-thrombin and protein C (PC) zymogen. Factor Va proteolysis was visualized by Western blot analysis using a monoclonal antibody (alpha HVaHC No. 17) to the factor Va heavy chain (HC), and cofactor activity was followed both in a clotting assay using factor V-deficient plasma and by quantitation of prothrombinase function. APC generation was monitored using the substrate 6-(D-VPR)amino-1-naphthalenebutylsulfonamide (D-VPR-ANSNHC4H9), which permits quantitation of APC at 10 pmol/L. Addition of APC (5 nmol/L) to an adherent HUVEC monolayer (3.5 x 10(5) cells per well) resulted in a 75% inactivation of factor Va (20 nmol/L) within 10 minutes, with complete loss of cofactor activity within 2 hours. Measurements of the rate of cleavage at Arg506 and Arg306 in the presence and absence of the HUVEC monolayer indicated that the APC-dependent cleavage of the factor Va HC at Arg506 was accelerated in the presence of HUVECs, while cleavage at Arg306 was dependent on the presence of the HUVEC surface. Factor Va inactivation proceeded with initial cleavage of the factor Va HC at Arg506, generating an M(r) 75,000 species. Further proteolysis at Arg306 generated an M(r) 30,000 product. When protein C (0.5 mumol/L), alpha-thrombin (1 nmol/L), and factor Va (20 nmol/L) were added to HUVECs an APC generation rate of 1.56 +/- 0.11 x 10(-14) mol/min per cell was observed. With APC generated in situ, cleavage at Arg506 on the HUVEC surface is followed by cleavage at Arg306, generating M(r) 75,000 and M(r) 30,000 fragments, respectively. In addition, the appearance of two novel products derived from the factor Va HC are observed when thrombin is present on the HUVEC surface: the HC is processed through limited thrombin proteolysis to generate an M(r) 97,000 fragment, which is further processed by APC to generate an M(r) 43,000 fragment. NH2-terminal sequence analysis of the M(r) 97,000 fragment revealed that the thrombin cleavage occurs in the COOH-terminus of the intact factor Va HC since both the intact HC as well as the M(r) 97,000 fragment have the same sequence. Our data demonstrate that the inactivation of factor Va on the HUVEC surface, initiated either by APC addition or PC activation, follows a mechanism whereby cleavage is observed first at Arg506 followed by a second cleavage at Arg306. The latter cleavage is dependent on the availability of the HUVEC surface. This mechanism of inactivation of factor Va is similar to that observed on synthetic phospholipid vesicles.  相似文献   

16.
The single-chain procofactor factor V is cleaved by thrombin (FVaIIa) at Arg709, Arg1018, and Arg1545 and by a variety of other proteases to generate a cofactor species with various levels of cofactor function. Having demonstrated previously that monocyte-bound forms of cathepsin G and elastase cleave and activate factor V, studies were initiated here using purified proteins to probe factor V structure/function. Electrophoretic, Western blotting, and amino-terminal sequence analyses revealed that cathepsin G cleaves factor V at several sites (Phe1031, Leu1447, Tyr1518, and potentially Tyr696), ultimately generating an amino-terminal 103 kDa heavy chain and a carboxy-terminal 80 kDa light chain (FVaCG). Elastase also cleaves factor V at several sites (Ile708, Ile819, Ile1484, and potentially Thr678), generating a cofactor species, FVaHNE, with an amino-terminal 102 kDa heavy chain and a carboxy-terminal 90 kDa light chain. Incubation of FVaIIa with either cathepsin G or elastase resulted in cleavage within the heavy chain, releasing peptides of approximately 2000 and approximately 3000 Da, respectively, generating FVaIIa/CG and FVaIIa/HNE. The functional activity of each cofactor species was assessed either by clotting assay or by employing a purified prothrombinase assay using saturating amounts of factor Xa. Significant differences in cofactor function were observed between the two assay systems. Whereas FVaIIa, FVaCG, FVaIIa/CG, FVaHNE, and FVaIIa/HNE all had similar cofactor activities in the purified prothrombinase assay, FVaCG and FVaHNE had no cofactor activity in the clotting-based assay, and FVaIIa/CG and FVaIIa/HNE had approximately 30-35% clotting activity relative to FVaIIa. These disparate results led us to examine the binding interactions of these cofactors with the various prothrombinase components. Kinetic analyses indicated that FVaIIa (Kd(app) = 0.096 nM), FVaIIa/CG (Kd(app) = 0.244 nM), and FVaIIa/HNE (Kd(app) = 0.137 nM) bound to membrane-bound factor Xa much more effectively than FVaCG (Kd(app) = 1.46 nM) and FVaHNE (Kd(app) = 0.818 nM). In contrast, studies of the activated protein C (APC)-catalyzed inactivation of each of the factor V(a) species indicated that they were all equivalent substrates for APC with no differences observed in the rate of inactivation or the cleavage mechanism, suggesting that APC interacts with the light chain at a site distinct from factor Xa. The Km values for prothrombin, as well as the kcat values for each of the FV(a) species, were all similar (approximately 0.25 microM and approximately 1900 min-1). In addition, kinetic analyses indicated that whereas FVaCG and FVaHNE exhibited a slightly reduced ability to interact with phospholipid vesicles (approximately 2-3-fold), the remaining FV(a) species assembled equally well on this surface. Collectively, these data indicate that FVaCG and FVaHNE have a diminished capacity to support factor Xa binding; however, cleavage at Arg1545 and removal of the extended B-domain in these cofactors restore near-total factor Xa binding. Thus, cleavage at Arg1545 optimizes cofactor function within prothrombinase by facilitating factor Xa binding to membrane-bound FVa.  相似文献   

17.
In preliminary studies, the generation of thrombin in vivo was found to induce a 92% loss of functional activity of factor IX (F.IX) despite the detection by Western blotting of a product resembling activated F.IX (F.IXa) and a 25% increase in F.IX antigen levels (Hoogendoorn et al, Thromb Haemost 69:1127, 1993 [abstr]). These changes were associated with evidence of increased elastase availability. To study the possibility that these two observations were related, a detailed physical and functional characterization of the hydrolysis of purified human F.IX by human neutrophil elastase (HNE) was performed in vitro. An activated partial thromboplastin time (aPTT) clotting assay demonstrated that, although HNE eliminated the potential of F.IX to be activated, it only marginally reduced the F.IXa activity. Reducing sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) indicated that HNE treatment of F.IX generated cleavage products of 30 and 20 kD that could not be distinguished from the respective heavy and light chain peptides that were identified in parallel studies when F.IX was activated by activated bovine F.XI (F.XIa), one of its physiological activators. In addition, nonreducing SDS-PAGE demonstrated that HNE-treated F.IX formed no complexes with antithrombin III (ATIII) in the presence of heparin. Furthermore, HNE-treated F.IX was unable to (1) bind the active site probe p-aminobenzamidine; (2) hydrolyze the synthetic peptide substrate CH3SO2-Leu-Gly-Arg-p-nitroanilide; and (3) activate human factor X (F.X). In contrast to dansyl-Glu-Gly-Arg-chloromethyl ketone (dEGR)-inactivated F.IXa, HNE-treated F.IX (0.01 to 10,000 pmol/L) failed to inhibit the clotting activity of F.IXa (10 pmol/L) in the aPTT. NH2-terminal sequencing indicated that HNE cleaved human F.IX at Thr140, Thr144, Ile164, Thr172, and Val181. The cleavages at Thr140/Thr144 and at Thr172/Val181 are both very close to the normal F.XIa alpha-(Arg145) and beta-(Arg180) cleavage sites, respectively. In summary, the results suggest that the activatability of F.IX is eliminated after cleavage by HNE and that the inability of HNE-treated F.IX to support F.IXa-like coagulant function is a consequence of improper active site formation. These in vitro observations support the possibility that increased HNE cleavage of F.IX in vivo may contribute to the disregulation of hemostasis that occurs in conditions such as disseminated intravascular coagulation (DIC).  相似文献   

18.
Factor (F) VIII inhibitor development remains one of the most serious complications in the treatment of hemophilia A. Former and recent studies on inhibitor development revealed that patients with severe hemophilia A and positive inhibitor family history are at highest risk of developing an inhibitor. Comparison of recent inhibitor incidence studies on previously untreated patients indicate that the risk of inhibitor development under treatment with recombinant FVIII concentrates is comparable to the inhibitor incidence under FVIII substitution by plasma-derived concentrates. However, longer observation periods are necessary to draw final conclusions. Since inhibitor development may result in inefficacy of FVIII concentrates in the treatment of severe bleedings, the induction of immune tolerance (IT) is still of main concern. Various regimens to induce IT by application of FVIII concentrates have been conducted up to now. Success rate appears to be influenced by low to high responder status, number of exposure days before onset of treatment, and dosage of therapeutic regimen. Especially, discontinuation of IT therapy seems to be associated with failure of therapy. Taking into account available data on IT therapy, we recommend early onset of a high dosage regimen in high responder patients as soon as possible after inhibitor detection, as this is associated with higher success rate and shorter elimination time.  相似文献   

19.
Tissue factor (TF)-induced coagulation was compared in contact pathway suppressed human blood from normal, factor VIII-deficient, and factor XI-deficient donors. The progress of the reaction was analyzed in quenched samples by immunoassay and immunoblotting for fibrinopeptide A (FPA), thrombin-antithrombin (TAT), factor V activation, and osteonectin. In hemophilia A blood (factor VIII:C <1%) treated with 25 pmol/L TF, clotting was significantly delayed versus normal, whereas replacement with recombinant factor VIII (1 U/mL) restored the clot time near normal values. Fibrinopeptide A release was slower over the course of the experiment than in normal blood or hemophilic blood with factor VIII replaced, but significant release was observed by the end of the experiment. Factor V activation was significantly impaired, with both the heavy and light chains presenting more slowly than in the normal or replacement cases. Differences in platelet activation (osteonectin release) between normal and factor VIII-deficient blood were small, with the midpoint of the profiles observed within 1 minute of each other. Thrombin generation during the propagation phase (subsequent to clotting) was greatly impaired in factor VIII deficiency, being depressed to less than 1/29 (<1.9 nmol TAT/L/min) the rate in normal blood (55 nmol TAT/L/min). Replacement with recombinant factor VIII normalized the rate of TAT generation. Thus, coagulation in hemophilia A blood at 25 pmol/L TF is impaired, with significantly slower thrombin generation than normal during the propagation phase; this reduced thrombin appears to affect FPA production and factor V activation more profoundly than platelet activation. At the same level of TF in factor XI-deficient blood (XI:C <2%), only minor differences in clotting or product formation (FPA, osteonectin, and factor Va) were observed. Using reduced levels of initiator (5 pmol/L TF), the reaction was more strongly influenced by factor XI deficiency. Clot formation was delayed from 11.1 to 15.7 minutes, which shortened to 9.7 minutes with factor XI replacement. The maximum thrombin generation rate observed ( approximately 37 nmol TAT/L/min) was approximately one third that for normal (110 nmol/L TAT/min) or with factor XI replacement (119 nmol TAT/L/min). FPA release, factor V activation, and release of platelet osteonectin were slower in factor XI-deficient blood than in normal blood. The data demonstrate that factor XI deficiency results in significantly delayed clot formation only at sufficiently low TF concentrations. However, even at these low TF concentrations, significant thrombin is generated in the propagation phase after formation of the initial clot in hemophilia C blood.  相似文献   

20.
In type 2N von Willebrand disease (vWD), von Willebrand factor (vWF) is characterized by normal multimeric pattern, normal platelet-dependent function, but a markedly decreased affinity for factor VIII (FVIII). In this report, we describe the case of a vWD patient who has an abnormal vWF multimers distribution associated with a markedly decreased vWF ability to bind FVIII. Sequencing analysis of patient's vWF gene showed, at heterozygous state, a G-->A transition resulting in the substitution of Asn for Asp at position 116 of the mature vWF subunit and a C-->T transition, changing the codon for Arg 896 into a stop codon. His sister who has a subnormal vWF level, but a normal FVIII/vWF interaction, was found to be heterozygous for the Arg896ter mutation only. Recombinant vWF (rvWF) containing the candidate (Asn116) missense mutation was expressed in COS-7 cells. The expression level of Asn116rvWF was significantly decreased compared with wild-type rvWF. The multimeric pattern of Asn116rvWF was greatly impaired as shown by the decrease in high molecular weight forms. The FVIII binding ability of Asn116rvWF was dramatically decreased. These data show that the Asp116Asn substitution is the cause of both the defective FVIII/vWF interaction and the impaired multimeric pattern observed in the patient's vWF. The monoclonal antibody 31H3 against D' domain of vWF (epitope aa 66-76) that partially inhibits the FVIII binding and recognizes only nonreduced vWF, showed a decreased ability to bind Asn116rvWF when used as capture-antibody in enzyme-linked immunosorbent assay (ELISA). This result suggests that a potential conformation change in the D' domain is induced by the Asp116Asn substitution, which is localized in the D3 domain.  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