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 19 毫秒
1.
To date, no dominant mutation has been identified in a significant proportion of patients with type 1 von Willebrand disease (VWD). In this study, we examined 70 families as part of the Canadian Type 1 VWD Study. The entire VWF gene was sequenced for 1 index case, revealing 2 sequence variations: intron 30 (5312-19A>C) and exon 28 at Tyr1584Cys (4751A>G). The Tyr1584Cys variation was identified in 14.3% (10 of 70) of the families and was in phase with the 5312-19A>C variation in 7 (10.0%) families. Both variants were observed in 2 of 10 UK families with type 1 VWD, but neither variant was found in 200 and 100 healthy, unrelated persons, respectively. Mean von Willebrand factor antigen (VWF:Ag), VWF ristocetin cofactor (VWF:RCo), and factor VIII coagulant activity (FVIII:C) for the index cases in these families are 0.4 U/mL, 0.36 U/mL, and 0.54 U/mL, respectively, and VWF multimer patterns show no qualitative abnormalities. Aberrant VWF splicing was not observed in these patients, and both alleles of the VWF gene are expressed as RNA. Molecular dynamic simulation was performed on a homology model of the VWF-A2 domain containing the Tyr1584Cys mutation. This showed that no significant structural changes occur as a result of the substitution but that a new solvent-exposed reactive thiol group is apparent. Expression studies revealed that the Tyr1584Cys mutation results in increased intracellular retention of the VWF protein. We demonstrate that all the families with the Tyr1584Cys mutation share a common, evolved VWF haplotype, suggesting that this mutation is ancient. This is the first report of a mutation that segregates in a significant proportion of patients with type 1 VWD.  相似文献   

2.
3.
In a normal population, VWF plasma levels (VWF:Ag) and VWF activity (VWF:RCo) increase by approximately 0.17 and 0.15 IU mL?1 per decade, but the influence of age is unknown in patients with type 1 von Willebrand disease (VWD). In a retrospective cohort study, the medical records of 31 type 1 VWD patients over the age of 30, who had been followed for ≥5 years, were reviewed for baseline clinical data and previously performed VWF:Ag, VWF:RCo and factor VIII levels (FVIII:C). VWF multimer analysis was normal in 28/31 cases performed. Mean age at diagnosis was 33 (range 16–60 years), and duration of follow‐up ranged from 5 to 26 years (mean 11 years). Patients had 2–10 time points of VWD testing (mean of 5.2). The mean VWF:Ag, VWF:RCo and FVIII:C at time of diagnosis were 0.44 IU mL?1 0.34 IU mL?1 and 0.75 IU mL?1. At last follow‐up, the mean VWF:Ag, VWF:RCo and FVIII:C were significantly increased to 0.71 IU L?1, 0.56 IU mL?1 and 0.90 IU mL?1 (≤ 0.001, <0.001, and 0.0081 respectively). Here 18/31 patients had VWF:Ag, VWF:RCo and FVIII: C levels that increased into the normal range. The rate of change in VWF:Ag, VWF:RCo and FVIII was 0.30 IU mL?1 (0.21–0.39, CI 95%, < 0.0001), 0.20 IU mL?1 per decade (0.13–0.27, CI 95%, P = 0.0001) and 0.20 IU mL?1 (0.11–0.29, CI 95%, P = 0.0011). Patients with type 1 VWD experience age‐related increases to VWF:Ag and VWF:RCo which can result in normalization of VWF levels. Further studies are required to determine if the bleeding phenotype resolves with the increases in VWF:Ag and VWF:RCo levels.  相似文献   

4.
Kroner  PA; Foster  PA; Fahs  SA; Montgomery  RR 《Blood》1996,87(3):1013-1021
In this report we describe the further investigation of the von Willebrand factor (vWF)/FVIII interaction in a type 1 von Willebrand disease patient characterized by discrepant VIII:C levels as determined by one-stage and two-stage VIII:C assays. A solid-phase binding assay shows that this patient's plasma vWF is moderately defective in capturing recombinant FVIII. Sequence analysis of the FVIII-binding domain encoded by the vWF mRNA of the affected individual identified mutations in both vWF alleles. In allele A, the mutations C2344T and T2451A result in the substitution of Trp for Arg19 (R19W) and of G1n for His54 (H54Q) in mature vWF, respectively. This allele also contains a reported polymorphism (A2365G, Thr26Ala). Allele B, which is underexpressed at the RNA level, contains a one-nucleotide deletion in the FVIII-binding domain (delta G2515) that results in the premature termination of translation. Analysis of the binding of FVIII by full- length vWF transiently expressed in COS-7 cells confirms that the combined R19W and H54Q substitutions are the cause of the defective vWF/FVIII interaction in this patient. The FVIII-binding defect of vWF containing either mutation alone is approximately half that of the double mutant, which suggests that the effect of these mutations is additive. The mutant proteins are recognized equally well by vWF monoclonal antibodies MBC105.4, 32B12, and 31H3, which block the binding of FVIII by vWF, indicating that amino acids Arg19, Thr26, and His54 are not critical residues in the epitopes of these antibodies.  相似文献   

5.
Significant linkage of types 2A and 2B von Willebrand disease (VWD) to the von Willebrand factor (VWF) gene have been reported, as well as mutations in the VWF gene. However, data for the partial quantitative variant are less consistent. An inconsistency of association between the type 1 VWD phenotype and genotype has been reported recently. We undertook linkage analysis of 12 families with definite or possible type 1 VWD patients. One family with classic type 1 VWD had a high lod score (Z = 5.28, theta = 0.00). A total lod score of 10.68 was obtained for the four families with fully penetrant disease. In two families linkage was rejected, while three families did not show conclusive evidence of linkage. This study corroborates ABO blood group influence, especially in patients with mild deficiencies and/or incomplete penetrance. Indirect genetic analysis may be an option for diagnosing asymptomatic or presymptomatic type 1 VWD carriers, particularly in families showing higher penetrance. The study indicates defects of the VWF locus are to be expected in more than half of the families studied. However, as defects at different loci may be the cause of this phenotype, the results of the segregation analyses should be interpreted with caution, especially in studies involving small families, or mild expressions of the disorder or incomplete penetrance.  相似文献   

6.
Type Vicenza variant of von Willebrand disease (VWD) is characterized by a low plasma von Willebrand factor (VWF) level and supranormal VWF multimers. Two candidate mutations, G2470A and G3864A at exons 17 and 27, respectively, of the VWF gene were recently reported to be present in this disorder. Four additional families, originating from northeast Italy, with both mutations of type Vicenza VWD are now described. Like the original type Vicenza subjects, they showed a mild bleeding tendency and a significant decrease in plasma VWF antigen level and ristocetin cofactor activity but normal platelet VWF content. Unlike the original patients, ristocetin-induced platelet aggregation was found to be normal. Larger than normal VWF multimers were also demonstrated in the plasma. Desmopressin (DDAVP) administration increased factor VIII (FVIII) and VWF plasma levels, with the appearance of even larger multimers. However, these forms, and all VWF oligomers, disappeared rapidly from the circulation. The half-life of VWF antigen release and of elimination was significantly shorter than that in healthy counterparts, so that at 4 hours after DDAVP administration, VWF antigen levels were close to baseline. Similar behavior was demonstrated by VWF ristocetin cofactor activity and FVIII. According to these findings, it is presumed that the low plasma VWF levels of type Vicenza VWD are mainly attributed to reduced survival of the VWF molecule, which, on the other hand, is normally synthesized. In addition, because normal VWF-platelet GPIb interaction was observed before or after DDAVP administration, it is proposed that type Vicenza VWD not be considered a 2M subtype.  相似文献   

7.
Menache 《Haemophilia》1998,4(S3):44-47
Nine patients with von Willebrand disease type 3, six with type 2B, one with type 2A, and one patient with type 1/2N were infused with one dose of ≈50 or 100 IU ristocetin cofactor activity (RCoF) per kg body weight of von Willebrand factor (vWF) (Human), a product with a very low content of factor VIII (FVIII). Blood samples were collected over 96 h. The data for RCoF and vWF antigen (vWF:Ag) were fitted to a 1-compartment model decay. The data for FVIII:C were fitted to a model with a linear time 'synthesis' term and a 1-compartment decay. Results in von Willebrand disease type 3 patients (nine patients; 10 infusions) indicated a volume of distribution of 39.9 and 39.8 mL kg−1 for RCoF and vWF:Ag, respectively. The FVIII:C rate of synthesis was 6.4 U dL−1 h−1 (range: 4.4–8.8). The decay rates for FVIII:C, RCoF, and vWF:Ag were 0.041 (h−1) [ t 1/2: 16.9 h]; 0.061 (h−1) [ t 1/2: 11.3 h] and 0.006 (h−1) [ t 1/2: 12.4 h], respectively. In patients with von Willebrand disease type 2 ( n  = 8) the RCoF mean volume of distribution was 46 mL kg−1. The factor VIIIC mean rate of synthesis was 5.5 U dL−1h−1 and the decay rate 0.043 (h−1) [ t 1/2: 16.1 h]. The rate of decay for RCoF and vWF:Ag were 0.050 (h−1) [ t 1/2: 13.9 h] and 0.044 (h−1) [ t 1/2: 15.7 h], respectively.  相似文献   

8.
A family with von Willebrand disease has been identified in which different members of the same sibship exhibit different abnormalities of von Willebrand factor (vWF). The two most severely affected sibs (bleeding time over 20 min) had abnormalities of vWF similar to those seen in type IIC. The smallest detectable multimer was increased and the triplet structure of individual multimers was replaced with a single band. The largest multimers could not be detected and there were relatively more small multimers than intermediate sized forms. vWF antigen (vWF:Ag) was decreased to 12.5-17% by electroimmunoassay (EIA) and to 3.2-5.5% by immunoradiometric assay (IRMA). In the less severely affected sibling (bleeding time 12.5 min) there was a similar relative increase in the smallest detectable multimer. However, the larger multimers were present and the relative concentration of large to small multimers was similar to normal. The triplet structure was altered in that the relative proportion of satellite bands to the central predominant band was decreased. vWF:Ag concentrations were moderately decreased (40-80% by EIA and 25-35% by IRMA). The father and grandfather showed a vWF multimeric pattern similar to the less severely affected sibling but there was no decrease in vWF:Ag concentration and their bleeding times were normal. These observations suggest that the interplay of several genetic factors is responsible for the expression of von Willebrand disease in this family.  相似文献   

9.
PURPOSE OF REVIEW: Type 1 von Willebrand disease (VWD) is the most common form of VWD, but has remained [corrected] the least well understood. Recent work is changing this situation. This review summarizes recent analysis of the genetic basis of the disease. RECENT FINDINGS: Linkage analysis demonstrates that dominantly inherited, fully penetrant VWD is present in approximately 50% of type 1 families. Between 55 and 70% of index cases analysed have a candidate von Willebrand factor gene (VWF) mutation, but no mutations are present in the promotor, or protein coding sequences or splice sites of remaining cases [corrected] Missense mutations throughout VWF predominate. Blood group O is much more common in type 1 von Willebrand disease than in the general population and is particularly prevalent in cases with incompletely penetrant mutations or no VWF mutation. SUMMARY: Type 1 von Willebrand disease can be divided into three groups where (1) fully penetrant VWF mutations appear sufficient to explain the low plasma von Willebrand factor and bleeding, (2) VWF mutation may act as a risk factor for bleeding in combination with blood group O and/other unknown genetic factors, and (3) classic VWF mutations are absent but VWF may still play a role in some cases and blood group O is common.  相似文献   

10.
The present study aimed to evaluate the parent-child relationship in differentiating between unaffected healthy individuals and those with von Willebrand disease (VWD). This study was performed on 15 children between the ages of 5 and 15 years and parents with personal and familial evidence of bleeding. Diagnosis of VWD as considered 'low von Willebrand factor (VWF) level or mild type 1 VWD' in the following children: those with low VWF levels (VWF:RCo and VWF:Ag between 30 and 50 U/dl), at least one bleeding symptom and a family member with at least one bleeding symptom. Laboratory values in the parents of families 1-7 were VWF:Ag 65-90, VWF:RCo 54-87, and FVIII:C 74-110, versus VWF:Ag 33-47, VWF:RCo 30-42, and FVIII:C 36-67 in their children. The normal laboratory values in the parents of families 1-7 suggested that their children would probably have low VWF levels. Our findings are that VWF levels are increasing with age. Laboratory values in the parents of families 8-15 were VWF:Ag 30-59, VWF:RCo 32-55, and FVIII:C 44-66, versus VWF:Ag 32-48, VWF:RCo 30-54, and FVIII:C 38-55 in their children. The laboratory values in the children from families 8-15 were close to the minimum range of normal or below normal, which suggested that it was possible that the parents and children in families 8-15 could be diagnosed as having mild type 1 VWD. The present study's findings show that comparison of the VWF levels in parents and their children may be helpful in differentiating children with low VWF levels and mild type 1 VWD from children that only have low VWF levels.  相似文献   

11.
Von Willebrand disease (VWD) is an inherited bleeding disorder, caused by quantitative (type 1 and 3) or qualitative (type 2) defects in von Willebrand factor (VWF). Gene therapy is an appealing strategy for treatment of VWD because it is caused by a single gene defect and because VWF is secreted into the circulation, obviating the need for targeting specific organs or tissues. However, development of gene therapy for VWD has been hampered by the considerable length of the VWF cDNA (8.4 kb [kilobase]) and the inherent complexity of the VWF protein that requires extensive posttranslational processing. In this study, a gene-based approach for VWD was developed using lentiviral transduction of blood-outgrowth endothelial cells (BOECs) to express functional VWF. A lentiviral vector encoding complete human VWF was used to transduce BOECs isolated from type 3 VWD dogs resulting in high-transduction efficiencies (95.6% +/- 2.2%). Transduced VWD BOECs efficiently expressed functional vector-encoded VWF (4.6 +/- 0.4 U/24 hour per 10(6) cells), with normal binding to GPIbalpha and collagen and synthesis of a broad range of multimers resulting in phenotypic correction of these cells. These results indicate for the first time that gene therapy of type 3 VWD is feasible and that BOECs are attractive target cells for this purpose.  相似文献   

12.
13.
14.
While studying patient plasma containing an unusual pattern of von Willebrand factor (VWF) multimers, we discovered a previously unreported phenomenon: heavy predominance of dimeric VWF. Genomic analysis revealed a new congenital mutation (Tyr87Ser) that altered the final stages of VWF biosynthesis. This mutation in the propeptide (VWFpp) resulted in synthesis of dimeric VWF with an almost complete loss of N-terminal multimerization. The multimer pattern in patient plasma appears to result from separate alleles' synthesizing wild-type or mutant (dimeric) VWF, with homodimers composing the predominant protomeric species. We have expressed VWF protein containing the Tyr87Ser mutation and analyzed the intracellular processing and resulting VWF biological functions. The expressed dimeric VWF displayed a loss of several specific functions: collagen binding, factor VIII binding, and ristocetin-induced platelet binding. However, granular storage of dimeric VWF was normal, demonstrating that the lack of multimerization does not preclude granular storage. Although the tertiary structure of the VWFpp remains unknown, the mutant amino acid is located in a region that is highly conserved across several species and may play a major role in the multimerization of VWF. Our data suggest that one function of the highly cysteine-rich VWFpp is to align the adjacent subunits of VWF into the correct configuration, serving as an intramolecular chaperone. The integrity of the VWFpp is essential to maintain the proper spacing and alignment of the multiple cysteines in the VWFpp and N-terminus of the mature VWF.  相似文献   

15.
16.
Summary Type IIA is a variant form of von Willebrand disease (vWD) characterized by the absence of von Willebrand factor (vWF) high molecular weight multimers in plasma. Most of the candidate missense mutations potentially responsible for type IIA vWD have been found clustered within a short segment of vWF, lying between Gly742 and Glu875 of the mature subunit. The present work reports a single heterozygous T → G transversion in eight patients from a large type IIA vWD family, resulting in the substitution Phe751→Cys. The absence of this mutation in 100 normal vWF genes as well as the lack, in these patients, of any other abnormality within the whole exon 28 encoding amino acids 463–921 of mature vWF, provide a strong support that this non-conservative mutation may be at the origin of the disease in this family. The presence of an additional cysteine at position 751 may induce a conformational change of the vWF subunit affecting either its ' in vivo ' sensitivity to proteolytic cleavage or, more likely, its intracellular transport as suggested by the abnormal multimeric pattern of platelet vWF observed in these patients.  相似文献   

17.
Reduced von Willebrand factor (VWF) half-life has been suggested as a new pathogenic mechanism in von Willebrand disease (VWD). The usefulness of VWF propeptide (VWFpp) in exploring VWF half-life was assessed in 22 type 1 and 14 type Vicenza VWD patients, and in 30 normal subjects, by comparing the findings on post-Desmopressin (DDAVP) VWF t(1/2) elimination (t(1/2el)). The VWFpp/VWF antigen ratio (VWFpp ratio) was dramatically increased in type Vicenza VWD (13.02 +/- 0.49) when compared to normal subjects (1.45 +/- 0.06), whereas it appeared to be normal in all type 1 VWD patients (1.56 +/- 0.7), except for the four carrying the C1130F mutation (4.69 +/- 0.67). A very short VWF t(1/2el) was found in type Vicenza VWD (1.3 +/- 0.2 h), while all type 1 VWD patients had a t(1/2el) similar to that of the controls (11.6 +/- 1.4 and 15.4 +/- 2.5 h respectively), except for the four patients carrying the C1130F mutation, who had a significantly shorter VWF survival (4.1 +/- 0.2 h). A significant inverse correlation emerged between VWFpp ratio and VWF t(1/2el) in both VWD patients and normal subjects. The VWFpp ratio thus seemed very useful for distinguishing between type 1 VWD cases with a normal and a reduced VWF survival, as well as for identifying type Vicenza VWD.  相似文献   

18.
Type IIB von Willebrand's disease (vWD) is a distinct form of this disorder in which the largest multimers of the von Willebrand factor (vWF) are lacking in plasma but present in platelets. When the vasopressin analogue, 1-deamino-8-D-arginine vasopressin (DDAVP), is given to patients with type IIB vWD, an abnormal vWF is released to plasma. This vWF causes thrombocytopenia in vivo and platelet aggregation in vitro. Aggregation occurs in the plasma milieu and thus at physiological fibrinogen concentration. In this study we demonstrate that IIB post-DDAVP vWF aggregated only metabolically active platelets. The platelet aggregation was completely inhibited by EDTA and PGE1, and either inhibited or greatly weakened by ASA, demonstrating the role of divalent cations and thromboxane A2 formation. In spite of inhibiting platelet aggregation, EDTA, PGE1 and ASA did not prevent platelet binding of IIB post-DDAVP vWF. An antiserum against GP Ib made normal platelets less responsive to the IIB vWF although neither platelet aggregation nor vWF binding were completely prevented. The aggregation was fibrinogen-dependent and platelets from patients with Glanzmann's thrombasthenia were unresponsive. The studies provide evidence that IIB post-DDAVP vWF is bound to unstimulated platelets and that the interaction between vWF and platelets in type IIB vWD is different from ristocetin-induced as well as thrombin- and epinephrine-induced binding to platelets of normal vWF.  相似文献   

19.
Kravtsov DV  Wu W  Meijers JC  Sun MF  Blinder MA  Dang TP  Wang H  Gailani D 《Blood》2004,104(1):128-134
The bleeding diathesis associated with hereditary factor XI (fXI) deficiency is prevalent in Ashkenazi Jews, in whom the disorder appears to be an autosomal recessive condition. The homodimeric structure of fXI implies that the product of a single mutant allele could confer disease in a dominant manner through formation of heterodimers with wild-type polypeptide. We studied 2 unrelated patients with fXI levels less than 20% of normal and family histories indicating dominant disease transmission. Both are heterozygous for single amino acid substitutions in the fXI catalytic domain (Gly400Val and Trp569Ser). Neither mutant is secreted by transfected fibroblasts. In cotransfection experiments with a wild-type fXI construct, constructs with mutations common in Ashkenazi Jews (Glu117Stop and Phe283Leu) and a variant with a severe defect in dimer formation (fXI-Gly350Glu) have little effect on wild-type fXI secretion. In contrast, cotransfection with fXI-Gly400Val or fXI-Trp569Ser reduces wild-type secretion about 50%, consistent with a dominant negative effect. Immunoprecipitation of cell lysates confirmed that fXI-Gly400Val forms intracellular dimers. The data support a model in which nonsecretable mutant fXI polypeptides trap wild-type polypeptides within cells through heterodimer formation, resulting in lower plasma fXI levels than in heterozygotes for mutations that cause autosomal recessive fXI deficiency.  相似文献   

20.
Endothelial cells (EC) were cultured from the umbilical cord of a male neonate whose mother was previously diagnosed with type IIA von Willebrand's disease (vWd). The diagnosis of type IIA vWd in the proband was confirmed by low ristocetin activity and the absence of the highest molecular weight (MW) forms of von Willebrand factor (vWf) in his platelet poor plasma. The vWf of EC cultured from the neonate's umbilical cord differed from that of control EC and the cell line EA.hy926 in two respects. Firstly, the full range of molecular weight forms was present in the patient EC lysate and, secondly, vWf:Ag expression was approximately seven-fold greater than that of control cells. Platelet lysates prepared from other affected members of the type IIA vWd family in the presence or absence of proteolytic inhibitors demonstrated a near normal vWf multimeric distribution. Resistance of these high MW forms to heat degradation was conferred by the presence of proteolytic inhibitors. Moreover, the full plasma vWf multimeric distribution could not be restored by the inclusion of EDTA. N-ethylmaleimide and leupeptin in the anticoagulant during the rapid preparation of platelet poor plasma. These findings lend support to the heterogeneous nature of type IIA vWd and has possible implications in the understanding of the intracellular processes involved in the biosynthesis and storage of the vWf macromolecular complex as well as the pathogenesis of type IIA vWd.  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