Journal articles on the topic 'Facteur VIII (FVIII)'

To see the other types of publications on this topic, follow the link: Facteur VIII (FVIII).

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Facteur VIII (FVIII).'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Labrouche, S., C. Vergnes, J. Constans, A. Berard, and G. Freyburger. "Augmentation du facteur VIII (FVIII) : recherche d’une implication moléculaire du LRP." Journal des Maladies Vasculaires 30, no. 4 (September 2005): 6. http://dx.doi.org/10.1016/s0398-0499(05)86244-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Yada, Koji, Kenichi Ogiwara, Masaru Shibata, Midori Shima, and Keiji Nogami. "Effects of anti-factor VIII inhibitor antibodies on factor VIIa/tissue factor-catalysed activation and inactivation of factor VIII." Thrombosis and Haemostasis 105, no. 06 (2011): 989–98. http://dx.doi.org/10.1160/th10-12-0781.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryFactor (F)VIIa/tissue factor (TF) rapidly activates FVIII activity by proteolysis at Arg372 and Arg740, and subsequently inactivates FVIIIa activity by proteolysis at Arg336, although this activation is weaker than that by thrombin. The effects of anti-FVIII inhibitor antibodies on these reactions remain unknown, however. In this study, 13 of anti-FVIII inhibitor antibodies recognising the A2 or C2 domain were prepared. None of them, irrespective of epitope specificity, significantly affected FVIIa/TFcatalysed FVIII activation in one-stage clotting assays. Anti-A2 and anti-C2 type 2 antibodies had little effect on the inactivation phase. Anti-C2 type 1 antibodies, however, modulated inactivation by 40–60% of that seen with control IgG, suggesting that the activity of FVIIIa generated by FVIIa/TF persisted in the presence of this specific type of inhibitor. SDS-PAGE analysis demonstrated that all antibodies had little effect on FVIIa/TF-catalyzed proteolysis at Arg372 and Arg740. Anti-C2 type 1, however, significantly delayed cleavage at Arg336 in dose-dependent manners. Neither anti-A2 nor anti-C2 type 2 affected this reaction, and the findings were consistent with the results of the functional assays. In addition, anti-C2 monoclonal antibodies with type 1 and 2 demonstrated similar patterns of reaction as the anti-C2 polyclonal antibodies in FVIIa/TF-mediated FVIII mechanisms. We demonstrated that FVIIa/TF activated FVIII even in the presence of anti-FVIII antibodies, but inactivation patterns appeared to depend on inhibitor type. It could be important to determine the characteristic of these inhibitor antibodies for prediction of their effects on FVIIa-related FVIII reactions, and the results could have significant therapeutic implications.Note: An account of this work was presented at the 51st annual meeting of the American Society of Hematology, 2009, New Orleans, LA, USA. This work was supported by grants for MEXT KAKENHI 21591370 in Japan and Bayer Hemophilia Award program.
3

Yada, Koji, Keiji Nogami, Kenichi Ogiwara, Katsumi Nishiya, Masahiro Takeyama, and Midori Shim. "Effects of Anti-FVIII Inhibitors On Factor VIIa/Tissue Factor-Catalyzed Activation and Inactivation of Factor VIII." Blood 114, no. 22 (November 20, 2009): 3169. http://dx.doi.org/10.1182/blood.v114.22.3169.3169.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 3169 Poster Board III-110 Factor (F)VIIa with tissue factor (TF) is a primary trigger of blood coagulation. We have recently demonstrated that FVIIa/TF rapidly activated FVIII by proteolysis of the heavy chain (HCh), and served physiologically as a potent activator for up-regulation of FVIII activity in very early-timed phase (ASH #1036, 2008). FVIII inhibitors develop as alloantibodies in multi-transfused patients with hemophilia A and also arise as autoantibodies in normal individuals. FVIII inactivation by inhibitors is associated with impairment of FVIII(a) cofactor function through the binding to functional crucial epitopes in FVIII. Anti-C2 inhibitors prevent FVIII binding to phospholipid, von Willebrand factor, and FXa. Anti-A2 inhibitors prevent FVIII binding to FIXa and thrombin. However, effects of these inhibitors on FVIIa action for FVIII have remained to be studied. In this study, we prepared 13 of anti-FVIII inhibitor IgGs (2 of anti-A2, 7 of anti-C2 with type 1 behavior, and 4 of anti-C2 with type 2). We first examined FVIIa/TF-catalyzed FVIII activation in the presence of anti-FVIII inhibitors in one-stage clotting assay. The levels of FVIII activity (10 nM) elevated rapidly by ∼2.0-fold within 30 sec after adding of FVIIa/TF (1 nM), and subsequently decreased to the initial level within 20 min. The presence of anti-FVIII inhibitors did not significantly affect FVIIa/TF-catalyzed FVIII activation (by 1.7∼2.2-fold) compared to normal IgG. This action was independent of the difference of inhibitor epitopes. In addition, FVIIa-catalyzed FVIIIa inactivation with anti-A2 or anti-C2 with type 2 inhibitors was little affected, similar to that with normal IgG. However, of note, all of anti-C2 with type 1 significantly inhibited FVIIa-catalyzed inactivation of FVIIIa. Inactivation rates of FVIIa with anti-C2 with type 1 (k ∼0.15) was ∼40% less than that with control IgG (k ∼0.24), supporting that the presence of anti-C2 with type 1 might persist the activity of FVIIIa generated by FVIIa. To clarify this inhibitory mechanism of anti-C2 with type 1, we performed FVIIa-catalyzed FVIII cleavage in Western blotting. FVIIa/TF (1 nM) proteolyzed the HCh of FVIII (10 nM) rapidly by cleavages at Arg372 (and Arg740), whilst cleavage at Arg336 in the A1 domain was appeared at ∼2.5 min, supporting that cleavages at Arg372 and Arg336 by FVIIa contribute to the up- and down-regulation of FVIII(a) activity, respectively. All inhibitors, independent of recognizing epitopes, did not affect FVIIa-catalyzed cleavage at Arg372. However, the presence of anti-C2 type 1 delayed the cleavage at Arg336 in timed- and dose-dependent manners, whilst either anti-A2 or anti-C2 type 2 did not affect, consistent with the functional inactivation results. FVIIa binds to the A2, A3, and C2 domains in FVIII. Based on our findings, FVIIa-interactive sites on FVIII unlikely overlapped with anti-A2 and -C2 inhibitor epitopes, and inhibition of Arg336 cleavage may be due to conformational change caused by antibody binding. Furthermore, FVIIa indeed activates FVIII even in the presence of anti-FVIII inhibitors, different from thrombin, FXa, etc, and it would be important to predict the effect of FVIIa for FVIII to determine the characteristics of anti-FVIII inhibitors. Disclosures No relevant conflicts of interest to declare.
4

Nakajima, Yuto, Koji Yada, Keiji Nogami, and Midori Shima. "A Novel Mechanism of Factor VIIa/Tissue Factor (TF)-Catalyzed Activation and Inactivation of B-Domain-Deleted Factor VIII in the Early Initiation Phases of Coagulation." Blood 132, Supplement 1 (November 29, 2018): 1162. http://dx.doi.org/10.1182/blood-2018-99-115645.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract We have reported that factor (F)VIII was rapidly activated by FVIIa/tissue factor (TF) in vitro by limited proteolysis of the heavy chain (HCh) at Arg372 and Arg740 in the very early-timed coagulation phase and inactivated by proteolysis at Arg336 (JTH 2010). Furthermore, the activation could be observed even in the presence of anti-FVIII inhibitors irrespective of their type of kinetics and the epitope recognized, whilst the inactivation was moderated by anti-C2 inhibitor with type 1 kinetics (Thromb Haemost 2011). A role of FVIII B-domain on FVIIa/TF-catalyzed activation and inactivation remain unknown, however. In this study, focusing on the roles of the B-domain of FVIII, we investigated the mechanism(s) of FVIIa/TF-catalyzed FVIIIa activation and inactivation by utilizing B-domain deleted (BDD)-FVIII as well as full-length (FL)-FVIII. We firstly examined FVIIa/TF-catalyzed activation and inactivation of FL- or BDD-FVIII(a) by a one-stage clotting assay. The FVIII activity (FVIII:C) of FL-FVIII (10nM) rapidly increased by ~1.7-fold within 0.5 min after addition of FVIIa (1nM)/TF (0.1nM), and subsequently decreased to the initial levels within 15 min (k = ~0.03). Interestingly, FVIII:C of BDD-FVIII (10nM), which increased up to ~1.7-fold of the initial level within 0.5 min after addition of FVIIa (1nM)/TF (0.1nM) similar to that of FL-FVIII, demonstrated a slower reduction to the initial level within 30 min (k = ~0.015) than that of FL-FVIII. In order to explore these inhibitory mechanisms of FVIIa/TF-catalyzed inactivation of BDD-FVIIIa, we investigated FVIIa/TF-catalyzed proteolytic cleavage of both BDD-FVIII and FL-FVIII by using SDS-PAGE. A rapid proteolysis in the heavy chain (Hch) of FL-FVIII within 0.5 min after addition of FVIIa/TF was observed by the cleavage at Arg740, followed by the cleavage at Arg372 and the subsequent cleavage at Arg336, consistent with our previous study. In contrast, it was of surprise that the proteolysis in the Hch of BDD-FVIII by cleavage at Arg372 was little observed after addition of FVIIa/TF, whilst that by the cleavage at Arg336 was observed within 0.5 min preceding the elevation of FVIII:C. The initial velocity of Arg336 cleavage at 0.5 min for BDD-FVIII (4.4/min) was ~3.3-times higher than that for FL-FVIII (1.4/min) by a densitometry. To the next, we examined the spontaneous dissociation of A2-domain from FVIIa/TF-catalyzed FL- or BDD-FVIIIa by a one stage clotting assay. In the presence of excess amount of A2-subunit (400nM), more than 50% of FVIIa/TF-catalyzed FVIIIa inactivation was inhibited compared to that in its absence, but no significant difference was observed between FL- and BDD-FVIII, suggesting that the spontaneous dissociation of A2-domain little affected the inhibition of the FVIIa/TF-catalyzed inactivation of BDD-FVIIIa. To further clarify the mechanism of FVIIa/TF-catalyzed BDD-FVIII activation/inactivation, we prepared and stably expressed recombinant BDD-FVIII mutants, R336A and R372A. FVIIa(1nM)/TF(0.1nM)-catalyzed activation and inactivation of R336A and R372A (10nM) was examined by a one stage clotting assay. FVIII:C of R336A and R372A rapidly increased by ~2.0-fold of the initial level within 0.5 min after addition of FVIIa/TF, similarly to that of wild type BDD-FVIII, and that of R336A subsequently decreased to the initial level within 30min (k = ~0.04), whilst little reduction of FVIII:C was observed for R372A (k = ~0.004). Evaluated by SDS-PAGE, FVIIa/TF-catalyzed proteolytic cleavage at Arg336 was predominantly observed for R372A within 0.5min after addition of FVIIa/TF, whilst cleavage at Arg372 was conversely observed for R336A. Taken together, FVIIa/TF-catalyzed activation of BDD-FVIII could be predominantly initiated by the cleavage at Arg336 or secondarily at Arg372 and resistance to the cleavage at Arg372 would hamper the subsequent inactivation. In conclusion, the B-domain of FVIII would regulate the FVIIa/TF-catalyzed activation and inactivation of FVIII by controlling the order of proteolytic cleavage at Arg336 and Arg372. We believe that our findings should also contribute to the development of more effective combination therapy of FVIIa and BDD-FVIII for hemophilia A with inhibitor. Disclosures Yada: Shire Japan Co., Ltd.: Other: Teacher at a endowed course. Nogami:Chugai Pharmaceutical Co., Ltd: Consultancy, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties: Anti-FIXa/X bispecific antibodies , Research Funding, Speakers Bureau. Shima:Chugai Pharmaceutical Co., Ltd: Consultancy, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties: Anti-FIXa/X bispecific antibodies , Research Funding, Speakers Bureau; F. Hoffmann-La Roche Ltd: Membership on an entity's Board of Directors or advisory committees.
5

Ogiwara, Kenichi, Keiji Nogami, Masahiro Okuda, Katsumi Nishiya, Masahiro Takeyama, and Midori Shima. "Interactions of Factor VIII with Tissue Factor Contributes to the Acceleration of Factor Xa Generation in the Initiation Phase of Blood Coagulation." Blood 114, no. 22 (November 20, 2009): 3177. http://dx.doi.org/10.1182/blood.v114.22.3177.3177.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 3177 Poster Board III-116 Activated factor (F)VII complex with tissue factor (FVIIa/TF) initiates the blood coagulation by generating FXa as extrinsic Xase complex (ex-Xase). Although FVIIa/TF also activates FIX, FIXa little functions without its cofactor, FVIIIa. A tiny amount of thrombin generated by FXa activates FV and FVIII, followed by forming of intrinsic Xase complex (in-Xase) and prothrombinase complex, respectively. These formations result in ‘thrombin burst’ and successful hemostasis. Although thrombin is thought to be a unique potent activator of FVIII in vivo, FXa and FVIIa/TF also activate FVIII in vitro. We have recently reported the detailed mechanism by which FVIIa/TF activated FVIII more rapidly in early timed-phase than thrombin (Blood Abst.1036, 2008). In this study, we further developed to examine whether TF affected FVIII(a) function. (1) FVIIa/TF rapidly increased FVIII activity by 4.7-fold of initial in the presence of Ca2+ and phospholipid (PL), following by inactivation, in one-stage clotting assay. However, since even in the presence of TF alone, FVIII activity elevated by 1.8-fold of initial, actual increase of FVIII activity by FVIIa/TF was 2.6-fold. A possibility that TF might bind to FVIIa contained in FVIII-deficient plasmas used, was negligible, since FVIIa-inhibitor used blocked an ex-Xase effect >95%. In the presence of FVIIa-inhibitor, residue FVIII activity with TF was ∼50%, thus TF alone affected FVIII cofactor activity independently of FVIIa. (2) Using SDS-PAGE, the addition of TF accelerated FVIII cleavage by FVIIa, whilst decelerated that by thrombin and FXa. (3) Surface plasmon resonance-based assays showed that FVIII(a) directly bound to TF with high affinity (Kd; ∼3 nM). (4) The effect of FVIIa/TF on in-Xase was evaluated in FXa generation assay. 0.1 nM FVIIa/TF, 1 nM FVIII, 90 nM FIX and 20 μM PL were reacted with 150 nM FX at various combinations. FVIIa/TF and FVIIa/TF/FVIII/FIX generated FXa with 3.9 and 10.4 nM/min, respectively. When FVIIa-inhibitor was added prior to addition of FX, FXa generated by FVIIa/TF and FVIIa/TF/FVIII/FIX were 5% and 46% (0.2 and 4.8 nM/min) of those without FVIIa-inhibitor, respectively. The latter was considered as FXa generated by in-Xase. Therefore, FXa derived from in-Xase was ∼40% of total FXa in this condition. (5) FVIIIa/FIXa (1 nM/2 nM)-dependent FXa generation in the presence of TF was evaluated. FXa generation in the presence of TF (0.02 and 0.3 nM) increased by ∼2 and ∼6-folds, respectively, of that in its absence. Furthermore, the functional affinity of FVIIIa for in-Xase complex in the presence of TF (0.1 nM), showed an ∼1.5-fold greater than that in its absence (Km; 4.9 ± 0.4 and 7.1 ± 0.9 nM, respectively). In conclusion, FVIIa/TF can generate FVIIIa in early timed-phase in vitro as well as FXa and FIXa, and possess potential of forming in-Xase. In addition, TF directly binds to FVIII(a), and functions in-Xase complex more efficiently by enhancing the affinity of FVIIIa for in-Xase. Although TF-dependent these reactions may be terminated rapidly via anticoagulant systems such as tissue factor pathway inhibitor, our data suggest that interactions of FVIII with TF might contribute to the acceleration of FXa generation in the initiation phase of blood coagulation. Disclosures Okuda: Sysmex Corporation: Employment.
6

Ogiwara, Kenichi, Keiji Nogami, Tetsuhiro Soeda, Tomoko Matsumoto, Katsumi Nishiya, and Midori Shima. "Mechanisms of Factor VIII Activation by Recombinant Factor VIIa Analog through Tissue Factor-Independent Manner." Blood 112, no. 11 (November 16, 2008): 1028. http://dx.doi.org/10.1182/blood.v112.11.1028.1028.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Factor VIIa (FVIIa), complexed with tissue factor (TF), is a trigger of blood coagulation. Analog of recombinant FVIIa (rFVIIa), NN1731 (V158D/E296V/M298Q) possesses a greater hemostatic effect than rFVIIa and has been expected in clinical application. Factor X activation rate of NN1731 compared to rFVIIa was 1.2-fold in the presence of TF (TF(+)), and was 30-fold on activated platelets in its absence (TF(−))(Allen, Arterioscler Thromb Vasc Biol.2007; 27: 683). This TF-independent mechanism likely attributes to excellent effects by NN1731. More recently, we reported the physiological role of FVIIa/TF-dependent FVIII activation in the early phase of blood coagulation. Therefore, we were tempted to investigate the action of NN1731 in FVIII activation. Time-dependent change in FVIII activity after the addition of rFVIIa/NN1731 was examined by one-stage clotting assay under the presence of phospholipids (PS:PC:PE=1:6:3), CaCl2 and TF(+)/TF(−). NN1731 raised FVIII activity up to peak level rapidly within 30 sec (TF(+)), following by inactivation. Peak level of FVIII activity by NN1731 in TF(−) reached to the same peak level of that in TF(+) within 5 min, and this peak level persisted for ~30 min. Whilst, peak FVIII level by rFVIIa in TF(−) showed only ~35% of that in TF(+) even at 30 min. FVIII activating rate of NN1731 was observed to be 1.2-fold (TF(+)) and 3.8-fold (TF(−)) of rFVIIa-catalyzed activation. Kinetics by the Xa generation assay showed the Km values of NN1731 in FVIII activation were ~1.5-fold lower than those of rFVIIa (NN1731/rFVIIa; TF(+) 27.3/49.2 nM and TF(−) 50.5/68.1 nM). Vmax values of NN1731 in FVIII activation, however, showed the obvious difference between TF(+) (2.3-fold; NN1731/rFVIIa 70.0/30.4 nM•min−1) and TF(−) (7.9-fold; 92.5/11.7 nM•min−1), compared to rFVIIa. Inactivation of FVIIIa by NN1731 was somewhat faster than that by rFVIIa. FVIII cleavages by NN1731 were analyzed using SDS-PAGE/Western blotting. The heavy chain of FVIII was proteolyzed at Arg740 (A2-B junction), Arg372 (A1-A2 junction) and Arg336 (within the A1), faster by NN1731 than by rFVIIa. These predominant cleavages by NN1731 were more evident in TF(−). However, little cleavage of the light chain of FVIII was observed by both proteases. FVIII cleavages were correlated with the observations of FVIII activation and/or inactivation. To further localize the binding region for NN1731, we evaluated the interactions between FVIII and Glu-Gly-Arg-active site modified (EGR-) NN1731, lacking enzymatic activity, in a surface plasmon resonance-based assay. The Kd value of EGR-NN1731 with FVIII was similar to that of EGR-rFVIIa (6.3 and 7.8 nM, respectively). Binding was particularly evident with the A2, A3, and C2 domains, whilst the A1 domain failed to bind, similar to the results obtained by rFVIIa. We demonstrated that NN1731 possesses higher potential as an activator for up-regulation of FVIII activity than rFVIIa. Furthermore, catalytic activity of NN1731 in TF(−), rather than binding affinity, likely attributes to this potential of its analog. We concluded that the analog has another novel mechanism in its potent hemostatic effect through FVIII activation in TF-independent manner.
7

Soeda, Tetsuhiro, Keiji Nogami, Tomoko Matsumoto, Kenichi Ogiwara, Katsumi Nishiya, and Midori Shima. "Tissue Factor-Dependent Activation of Factor VIII by Factor VIIa in the Early Phase of Blood Coagulation." Blood 112, no. 11 (November 16, 2008): 1036. http://dx.doi.org/10.1182/blood.v112.11.1036.1036.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Factor VIIa (FVIIa), complexed with tissue factor (TF), is a trigger of blood coagulation through activation of factor X in the initiation phase. FVIIa can catalyze intrinsic clotting factors such as not only factor IX, but also factor VIII (FVIII). However the role and the mechanisms of the FVIIa-catalyzed FVIII are poorly understood. We first examined FVIIa-catalyzed FVIII activation in the presence of phospholipid (PL) using a one-stage clotting assay. The levels of FVIII activity elevated rapidly by ~4-fold within 30 sec after the addition of FVIIa (1 nM)-TF (1 nM)complex, and subsequently decreased to the initial level within 20 min. This time-dependent reaction was enhanced by the presence of TF and PL in dose-dependent manners, but was moderately inhibited (~50%) in the presence of von Willebrand factor at physiological concentration of 10 μg/mL. FVIII cleavage was evaluated using western blotting immediately after the addition of FVIIa-TF complex. The heavy chain of FVIII was proteolyzed more rapidly (at 15 sec) by cleavages at Arg740 (A2-B junction) and Arg372 (A1-A2 junction) by FVIIa-TF complex, whilst the cleavage at Arg336 in the A1 domain was appeared at ~2.5 min. However little cleavage of the light chain of FVIII was observed, supporting that cleavages at Arg740/Arg372 and Arg336 by FVIIa-TF complex contribute to the up- and down-regulation of FVIII(a) activity, respectively. Of interest, no proteolysis of isolated intact heavy chain was observed, indicating that the proteolysis of the heavy chain was governed by the presence of the light chain. Compared to FVIII activation by thrombin (0.1–1 nM), the activation by FVIIa (0.1–1 nM)-TF (1 nM) complex was observed more rapidly. The activation rate observed by the addition of FVIIa-TF complex was ~50-fold greater than that by thrombin. Kinetics by the chromogenic Xa generation assay showed the catalytic efficiency (kcat/Km; 8.9 min−1/32.8 nM) on FVIIa-TF complex-catalyzed FVIII activation showed ~4-fold greater than that on thrombin-catalyzed activation (kcat/Km; 7.5 min−1/86.4 nM). Furthermore, the catalytic efficiencies on cleavages at Arg740 and Arg372 of FVIII by FVIIa-TF complex were ~3- and ~20-fold greater compared to those by thrombin, respectively. These findings suggested that FVIIa-TF complex was a greater FVIII activator than thrombin in very early phase. In order to localize the binding region for FVIIa, we evaluated the interactions between FVIII subunit and Glu-Gly-Arg-active site modified FVIIa, lacking enzymatic activity, in a surface plasmon resonance-based assay. The heavy chain of FVIII bound to EGR-FVIIa with higher affinity than the light chain (Kd; 2.1 and 45 nM, respectively). Binding was particularly evident with the A2, A3, and C2 domains (Kd; 34, 37, and 44 nM, respectively), whilst the A1 domain failed to bind. In conclusion, we demonstrated that FVIIa-TF complex rapidly activated FVIII by proteolysis of the heavy chain and the activation was governed by the presence of the light chain. Furthermore, present results suggested the role of TF-dependent FVIII activation by FVIIa which is responsible for the initiation phase of blood coagulation.
8

Shoko, Furukawa, Keiji Nogami, Kenichi Ogiwara, and Midori Shima. "Mechanism of Tissue Factor (TF) Enhancing Factor (F)VIII Activity on FXa Generation in Initial Phase of Coagulation and Interaction Between TF and FVIII C2 Domain." Blood 128, no. 22 (December 2, 2016): 1391. http://dx.doi.org/10.1182/blood.v128.22.1391.1391.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract In the cell-based coagulation model, factor (F)VIIa complex with tissue factor (TF) initiates the blood coagulation by generating FXa as the extrinsic tenase complex and activates FIX which composes the intrinsic tenase complex. We demonstrated that FVIIa/TF directly activated FVIII in an early coagulation phase (Soeda, JTH, 2010), and TF enhanced the intrinsic tenase activity via possible interaction with FVIIIa (Ogiwara, ASH, 2010). In this study, we clarified the enhancing mechanism of intrinsic tenase activity in the TF-related up-regulation of FVIII, and identified the TF-interactive region on FVIII. To explore the enhancing mechanism of TF for FVIII regulation, we performed the FXa generation assay with various amounts of FVIII or thrombin-mediated FVIIIa, constant FIXa (1 nM), FX (300 nM) and phospholipid vesicles (PL; 20 µM) in the presence of recombinant lipidated TF (rTF, Innovin®). The Km value for FVIII in the presence of rTF was ~2.4-fold lower than that its absence (Km; 5.6±0.5, 13.3±3.7 nM, respectively, p<0.05). Similarly, the Km for FVIIIa in the presence of TF was ~1.5-fold lower than that in its absence (Km; 6.3±0.6 nM, 9.7±1.6 nM, respectively, p<0.05), supporting that the presence of TF could promote the FXa-catalyzed activation of FVIII and FVIIIa-dependent generation of intrinsic FXa. To further evaluate the effect of TF on FVIII-dependent FXa generation, the FXa generation assay with FVIIa/TF-activated FVIIIa (FVIIIa-VIIa/TF) was also performed. The initial velocity on FXa generation with FVIIIa-VIIa/TF was 22.6 nM/min. However, the initial velocity on FXa generation with FVIIIa-VIIa/TF by addition of FVIIa-inhibitor (E-76), not to generate FVIIa/TF-dependent FXa, was 3.4 nM/min, and that with FVIII alone was 0.05 nM/min. In addition, the initial velocity with FVIIa/TF alone was 10.4 nM/min. These findings supported that the TF increased FXa generation greater than the additive effect of FVIII-dependent and FVIIa/TF-dependent FXa generation in early initiation phase of coagulation prior to thrombin generation. Since tissue factor pathway inhibitor (TFPI) is present in physiological circulating whole blood, a similar experiment on FXa generation assay was repeated under the presence of TFPI, estimated to be present at 0.5 nM in the pre-coagulant state or at 15 nM in the coagulant state in circulating blood. The initial velocity on FXa generation with FVIIIa-VIIa/TF was reduced by the presence of 0.5 or 15 nM TFPI (17.9 and 12.6 nM/min, respectively). By contrast, the initial velocity on FVIIIa-VIIa/TF-dependent FXa generation with addition of FVIIa inhibitor was little reduced by 0.5 nM TFPI, whilst was reduced by 15 nM TFPI (3.5 and 2.5 nM/min, respectively). These findings supported that the TFPI possibly didn't inhibit TF on the enhanced intrinsic tenase on association with FVIII in the pre-coagulant state. We further reported that TF enabled FXa to activate FVIII, irrespective of von Willebrand factor (VWF), and the direct association of rTF and non-lipidated TF with FVIII (Furukawa, ISTH, 2015). Since TF is transmembrane protein, however, we performed a surface plasmon resonance (SPR)-based assay (BIAcore®) and solid phase-based ELISA to identify the interactive region(s) on FVIII to recombinant soluble TF (sTF; Altor BioScience), a portion of TF outside of the PL membrane. An SPR-based assay revealed the direct binding of intact FVIII, LCh (a3-A3C1C2, A3C1C2) subunit, C2 domain to immobilized sTF (Kd; 2.3±0.6, 5.8±1.0, 10.5±3.5, 11.8±0.5 nM, respectively). The intact HCh, A1 or A2 domain to sTF failed to bind, however. A non-equilibrium ELISA also revealed that sTF bound to immobilized C2 domain with moderate affinity (Kdapp; 16.9±2.2 nM), and the interaction was dependent on ionic strength and Ca2+. In addition, the presence of VWF significantly competitively inhibited the C2 and sTF binding by ~90% (IC50; 5.7 µg/ml) at the maximal concentration employed, suggesting that the C2 domain-TF interaction could activate FVIII by FXa even in the presence of VWF. We concluded that it might be possible that TF enhanced the FVIII-mediated FXa generation by not only FVIIa but also FXa, additionally this enhancing mechanism might not be suppressed by TFPI in the initiation phase of coagulation. Furthermore, TF might function to FVIII activation, irrespective of presence of VWF, by the binding to C2 domain through the competition with VWF. Disclosures Nogami: Sysmex Corporation: Patents & Royalties, Research Funding; F. Hoffmann-La Roche Ltd.: Honoraria, Membership on an entity's Board of Directors or advisory committees; Chugai Pharmaceutical Co., Ltd.: Honoraria, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties, Research Funding. Shima:Sysmex Corporation: Patents & Royalties, Research Funding; F. Hoffmann-La Roche Ltd.: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees; Chugai Pharmaceutical Co., Ltd.: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Patents & Royalties, Research Funding.
9

Radtke, Klaus-Peter, Dean Chamberlain, John H. Griffin, and Andrew J. Gale. "Whole Blood Thromboelastogram Assays Demonstrate Prolonged Factor VIIIa Potency for Recombinant Disulfide Bond-Stabilized Factor VIII Variants." Blood 104, no. 11 (November 16, 2004): 2976. http://dx.doi.org/10.1182/blood.v104.11.2976.2976.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Following proteolytic activation of factor VIII (FVIII) by thrombin, the FVIIIa A2 domain, A3 domain and light chain (A3-C1-C2 domains) form a non-covalent hetero-trimer. Because spontaneous A2 subunit dissociation causes loss of FVIIIa activity, we previously made two mutants each with two new Cys to form a disulfide bond linking residues 662 (A2) and 1828 (A3) (FVIIIC662-C1828) or residues 664 (A2) and 1826 (A3) (FVIIIC664-C1826). Following thrombin activation, each FVIIIa mutant was stabile compared to wild type (wt) B-domain-deleted (BDD) FVIII. Previous SDS-PAGE data showed that the A2 domain was disulfide linked to the light chain. To show that this is true for undenatured FVIIIa, here we used surface plasmon resonance (SPR) to monitor A2 dissociation from thrombin-activated wild type and variant FVIII species that were bound to the sensor surface via a monoclonal antibody. Following passage of thrombin over sensor-bound FVIII, only wt FVIII showed a characteristic decrease of SPR reflecting A2 subunit dissociation and thrombin-treated FVIIIC662-C1828 and FVIIIC664-C1826 showed only minor decreases in SPR. Thus, SPR data directly demonstrate that engineered inter-domain disulfide bridges between the A2 and A3 domains prevent A2 domain dissociation from FVIIIa. In contrast to simple plasma coagulation assays of FVIIIa, rotational thromboelastogram (RoTEG) assays of whole blood provide multiple parameters reflecting clot formation, clot quality, and clot dissolution. RoTEG assays using fresh severe hemophilia A whole blood that was reconstituted with either wt FVIII, or FVIIIC662-C1828 or FVIIIC664-C1826 were performed to test the hypothesis that the disulfide-stabilized FVIIIa mutants would show improved potency for thrombin generation. After recalcification of hemophilia A blood with added FVIII, we measured the clotting time (CT), the rate of clot-formation, the clot-firmness time (CFT), defined as the time required to reach a specified clot firmness, and the clot firmness at 5 min (CF-A5), defined as the clot firmness at 5 min after the observed CT. Samples reconstituted with disulfide-bridge-stabilized FVIII mutants or wt-FVIII had comparable CTs at similar concentrations. However, in comparison to wild type BDD-FVIII, comparable rates of clot-formation, CFTs and CF-A5 were observed for up to 10-fold lower concentrations of each disulfide-bridge-stabilized FVIII mutant. The differences between wt and FVIII mutants were especially pronounced at very low FVIII concentrations whereas at FVIII concentrations &gt;0.01 U/mL the differences were less apparent. Because clot formation occurs early relative to overall thrombin generation which is better reflected by CFT and CF-A5 values, we interpret these data to indicate that the disulfide-stabilized FVIIIa variants provide sustained thrombin generation in whole blood compared to wt FVIII and speculate that these FVIII variants may prove superior to wt FVIII for stabilizing a hemostatic plug by providing sustained thrombin generation capacity.
10

Meeks, Shannon, Ernest T. Parker, Amy L. Dunn, John F. Healey, and Pete Lollar. "Proteolytically Inactivatable Factor VIII is Less Immunogenic than Factor VIII in a Murine Hemophilia A Model." Blood 114, no. 22 (November 20, 2009): 27. http://dx.doi.org/10.1182/blood.v114.22.27.27.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 27 Patients with hemophilia A have a congenital deficiency of the factor VIII (fVIII) protein due to a mutation in the fVIII gene that frequently leads to absence of detectable expression of fVIII. Accordingly, the therapeutic replacement fVIII protein potentially is recognized as non-self by the immune system. Thirty percent of patients with severe hemophilia A develop detectable inhibitory anti-fVIII antibodies (inhibitors). Additionally, greater than 90 percent of hemophilia A mice treated with human fVIII develop inhibitors using dosing schedule that mimics use in humans. Because fVIII is an immunologically foreign protein, it might be expected that a hemophilia A patient would make a fVIII inhibitor. However, intravenous injection of soluble proteins in either humans or rodents usually results in tolerance rather than a humoral immune response. One major difference between fVIII and other proteins is that it is released from its large carrier protein von Willebrand factor (VWF) and is potentially exposed to the immune system at sites of active hemostasis and inflammation. Heat-inactivated, denatured fVIII, which maintains all T-cell epitopes but lacks several B-cell epitopes, is less immunogenic than native fVIII, suggesting that fVIII-dependent thrombin generation along the intrinsic pathway of blood coagulation may provide co-stimulatory signals necessary for the immune response (Skupsky BS, Zhang A, Scott DW Blood 2008; 112:1220a). We constructed a B domain-deleted human fVIII mutant, designated fVIIIi, which contains alanine substitutions at two critical thrombin cleavage sites, Arg372 and Arg1689, and purified it to homogeneity. FVIIIi does not develop procoagulant activity and is not released from VWF in response to thrombin. Therefore fVIIIi is less likely than wild-type fVIII to be exposed to the immune system at sites of active hemostasis and inflammation. Additionally, VWF binds to the immunodominant fVIII C2 domain and potentially hides part of fVIII from the immune system. FVIIIi was antigenically intact judging from intact binding to a panel of11 mouse anti-fVIII monoclonal antibodies whose epitope specificity was represented by all five domains of BDD fVIII. The immunogenicity of wild-type fVIII and fVIIIi was compared in a murine hemophilia A model in which groups of 25 mice received 8 weekly injections of physiologic doses of fVIII. Plasma was collected weekly for total anti-fVIII antibody titers by ELISA and one week following the last injection for total anti-fVIII antibody titers, inhibitor titers by Bethesda assay and for epitope mapping. Mice treated with fVIIIi had significantly lower levels of inhibitory as well as total anti-fVIII antibodies than mice treated with wild-type fVIII. Domain mapping using single human domain hybrid human/porcine molecules as ELISA antigens revealed that hemophilia A mice broadly recognized all fVIII domains in response to either wild-type or fVIIIi, although fVIIIi produced less anti-light chain antibodies. Mice in both the wild-type fVIII and fVIIIi groups produced antibodies that recognized the phospholipid-binding site of the C2 domain, even though this site overlaps the VWF binding site on fVIII. There was no difference in the isotype spectrum of the antibodies made to fVIII or fVIIIi. This study indicates that inactivatable fVIII is less immunogenic than native fVIII and suggests that the immunogenicity of fVIII is related either to its interaction with VWF or to events triggered by activation of the coagulation mechanism. Disclosures: No relevant conflicts of interest to declare.
11

Yada, Koji, Keiji Nogami, Kenichi Ogiwara, and Midori Shima. "A novel mechanism of Enhancing the Haemostatic Effect in the Combination with Recombinant Factor VIII and Activated Prothrombin Complex Concentrate(APCC) in Hemophilia A Patients with Inhibitor." Blood 118, no. 21 (November 18, 2011): 1178. http://dx.doi.org/10.1182/blood.v118.21.1178.1178.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 1178 We have reported that factor (F)VIIa/tissue factor (TF) rapidly activated FVIII by proteolysis of the heavy chain (HCh), and that appeared to serve physiologically as an activator for up-regulation of FVIII activity in very early-timed coagulation phase (J Thromb Haemost. 2010;8:2494). Furthermore, the activation phase could be observed even in the presence of anti-FVIII inhibitors, independently of their types of kinetic and epitope of inhibitors, whilst the inactivation was moderated by anti-C2 with type 1 behavior (Thromb Haemost. 2011;105:989). More recently, the other group has reported that the combination of FVIII and bypassing agent, APCC, also potentiated the thrombin generation in hemophilia A plasmas with inhibitors (Klintman et al. Br J Haematol. 2010;151:381), but the mechanisms have remained unknown. In this study, we investigated the hemostatic effect of the combination therapy with APCC and FVIII. We first examined FVIII activation catalyzed by APCC in one-stage clotting assay. The activity level of FVIII (10 nM) elevated rapidly by ∼3-fold within 1 min after addition of APCC (0.05 U/mL)/TF (0.5 nM), and subsequently decreased to the initial level within 10 min. However, the addition of APCC without TF little affected FVIII activity within 10 min, but after then gradually elevated its activity. The presence of E-76, FVIIa-specific inhibitor, significantly moderated the reaction triggered by APCC/TF, but hirudin, FIIa-specific inhibitor, little affected this activation. We further evaluated the velocity of APCC-induced thrombin generation in the presence or absence of FVIII. The increase rate of thrombin production triggered by APCC/TF in the presence of FVIII was greater than that by APCC/TF in the absence of FVIII by ∼1.6-fold, whilst the increase was little observed in the absence of TF. To clarify its enhancing effects, we performed APCC-catalyzed FVIII cleavage in SDS-PAGE and Western blot. APCC contains FVII (mainly active form) and FII, FIX and FX (mainly non-active forms). In general, FVIIa/TF, FIIa, and FXa rapidly proteolyze the HCh at Arg372 (and Arg740), and FVIIa/TF and FXa proteolyze at Arg336. FVIII proteolysis by FVIIa/TF is dependent on the presence of PL, but not FIIa. Interestingly, APCC/TF proteolyzed the HCh at Arg372 and Arg740, followed by at Arg336 in the presence of PL, whilst did not proteolyze in its absence. The presence of TF accelerated the proteolysis by ∼6.6-fold compared to the absence of TF. However, the addition of E-76 significantly delayed these cleavages. These findings supported that APCC as well as rFVIIa possessed a potential to activation/inactivation of FVIII in early-timed coagulation phase, and that FVIIa in APCC appeared to play a major role in APCC-catalyzed FVIII activation/inactivation. Furthermore, to confirm this enhancing effect in the presence of anti-FVIII inhibitors, we prepared the anti-FVIII inhibitor IgGs (3 of anti-A2, 4 of anti-C2 with type 1, and 2 of anti-C2 with type 2). The presence of anti-FVIII inhibitors did not significantly affect the APCC-catalyzed FVIII activation (by ∼3-fold), independently of their epitopes, in one-stage clotting assay. Of surprise, anti-C2 with type 1 significantly moderated APCC-catalyzed FVIIIa inactivation, and the peak level of FVIIIa retained over 30 min. In contrast, the other inhibitors little affected this inactivation, similar to FVIIa/TF-catalyzed FVIII reaction. In conclusion, we demonstrated the putative mechanism of enhancing hemostatic effects in the combination therapy using FVIII and APCC. In addition, only a small amount of APCC relative to the standard dosage (1–2 U/mL) for clinical use could activate FVIII even in the presence of anti-FVIII inhibitors, and this combination therapy would provide new therapeutic strategy in congenital hemophilia A with inhibitor and/or acquired hemophilia A. Disclosures: Yada: Baxter Hemophilia Scientific Research and Education Fund 2011: Research Funding. Nogami:Bayer Award 2009: Research Funding.
12

Messer, Amanda S., Barbara Ulmasov, Yogesh Kumar, Kanagasabai Vadivel, Degang Zhong, Philip Fay, and S. Paul Bajaj. "Epitope Mapping of a Monoclonal Antibody to Factor VIII That Inhibits Factor IXa:Factor VIIIa Interaction and Thrombin Activation of Factor VIII." Blood 118, no. 21 (November 18, 2011): 1177. http://dx.doi.org/10.1182/blood.v118.21.1177.1177.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 1177 Factor VIII (FVIII) circulates in plasma as a noncovalent heterodimer consisting of a heavy chain (HC, A1-a1-A2-a2-B domains) and a light chain (LC, a3-A3-C1-C2 domains) in a noncovalent complex with von Willebrand factor (wVF). Thrombin (IIa) cleaves FVIII between the A1-a1/A2 domains at Arg372, A2-a2/B domains at Arg740 and B-a3/A3 domains at Arg1689 generating FVIIIa that consists of an A1-a1/A2-a2/A3-C1-C2 heterotrimer. FVIIIa increases the efficiency of Factor IXa (FIXa) catalyzed activation of Factor X (FX) in a Ca2+ and phospholipid (PL) dependent manner. The A3-C1-C2 segment of FVIIIa plays an important role in FIXa:FVIIIa interaction. Here, we describe a series of experiments to map the epitope of a monoclonal antibody (mAb) that is reported to inhibit FVIII clotting activity in a one stage clotting assay (Brown et al; J Lab Clin Med, 101: 793–805, 1983). The binding of mAb to FVIII, B-domain deleted FVIII and isolated LC was assessed using surface plasmon resonance. In these experiments, mAb captured on a protein A/G coupled CM5 sensor chip served as the ligand, and FVIII and its isolated fragments served as the analytes. The Kd of binding of LC (∼40 nM) was similar to FVIII and the B-domain deleted FVIII. No binding was observed for isolated A1 and A2 domains. Further, in plasma based inhibition assays, the Kd of binding of mAb to FVIII-vWF complex and to FVIII was ∼30 nM. This suggests that the mAb epitope does not significantly overlap with the vWF binding site in the acidic a3 region of LC. Western blot analysis confirmed that the mAb is specific for the LC of FVIII. Moreover, IIa-cleaved LC starting at residue 1690 gave only a weak signal and FXa-cleaved LC starting at residue 1721 did not react with the mAb in Western blots. These data suggest that the epitope for this mAb spans the IIa-cleavage site in the LC. Consistent with these observations, the A3-C1-C2 fragment but not the C1-C2 fragment expressed in COS cells reacted with the mAb. To further define a part of the epitope in the IIa-cleaved LC, twelve A3 domain deletion fragments were constructed and expressed in E. coli. Western blot analysis of these fragments restricted the partial epitope to 1690–1710 residues of the IIa-cleaved LC. In additional experiments, the mAb did not inhibit mouse, rabbit or canine plasma FVIII in a one stage clotting assay. It did however inhibit porcine plasma FVIII with ∼40 nM Kd, sheep plasma FVIII with ∼ 68 nM Kd, and bovine plasma FVIII with ∼300 nM Kd. Analysis of the sequence alignment of residues 1680 to 1710 of FVIII from each species indicated that residues 1681 to 1694 of human FVIII most likely constitute the epitope of this mAb. The dissimilarity and the charge differences in amino acids suggest that residues Asp1681, Glu1684, Asn1685, and Ser1687 on the N terminal side and Lys1693 on the C terminal side of the IIa-cleavage site Arg1689-Ser1690 may be important for this epitope. Fluorescence energy transfer (FRET) experiments indicated that the mAb inhibits FIXa interaction with the IIa-cleaved LC consisting of A3-C1-C2 domains. In these experiments, A3-C1-C2 subunit was labeled with acrylodan (fluorescence donor) and FIXa was labeled with fluorescein-Phe-Phe-Arg-chloromethylketone (fluorescence acceptor). In the presence of FIXa, the acrylodan fluorescence was quenched indicating a biomolecular complex formation. Addition of 1.2 μM mAb abolished the acrylodan fluorescence quenching suggesting inhibition of the FIXa:LC interaction. Notably, the mAb did not inhibit activation of FX by FIXa/Ca2+/PL and FXa-cleaved FVIIIa (instead of IIa-cleaved FVIIIa). This suggests that the mAb inhibits FIXa:LC interaction by a steric hindrance and not by a direct blockage of the FIXa:LC interactive sites. In summary, the mAb inhibits clotting by preventing FVIII activation by IIa. The epitope of the mAb appears to be restricted to residues 1681–1694 of FVIII. Notably, in some of the hemophilia A patients, the epitope of the inhibitory antibodies is confined to the IIa-cleavage site including the a3 acidic domain of LC. To locate the epitope for such antibodies, one of the approaches used was to construct porcine and human FVIII hybrids. Our strategy may represent a simplified approach to locate the epitope of similar antibodies in hemophilia A patients. Such antibodies may bind strongly to LC and weakly to IIa-cleaved LC. Further, these antibodies may not bind to FXa-cleaved LC or A1/A2 subunits. Disclosures: No relevant conflicts of interest to declare.
13

Zakas, Philip, Kristopher Knight, Ernest T. Parker, H. Trent Spencer, Eric Gaucher, and Christopher B. Doering. "Bioengineering Coagulation Factor VIII through Ancestral Protein Reconstruction." Blood 126, no. 23 (December 3, 2015): 123. http://dx.doi.org/10.1182/blood.v126.23.123.123.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract The development of transformative hemophilia A therapeutics has been hindered by the size, instability, immunogenicity and biosynthetic inefficiency of coagulation factor VIII (FVIII). Through the study of FVIII orthologs from existing vertebrate species, we discovered unique molecular, cellular and biochemical properties that can overcome the limitations of human FVIII. This approach facilitated the development of recombinant porcine FVIII for acquired hemophilia A and has enabled low resolution mapping and bioengineering of functional sequence determinants into human FVIII. To further extend this bioengineering approach, we employed a novel methodology termed ancestral protein reconstruction that provides certain advantages over 'rational design' approaches including a priori confidence that each ancestral FVIII is hemostatically functional. First, a mammalian FVIII phylogenetic tree with corresponding ancestral node (An) sequences was constructed through Bayesian inference using both DNA and amino acid-based models in PAML Version 4.1 (Figure 1). The limited availability of non-mammalian sequences precluded accurate ancestor prediction outside of this class. Initially, we selected 14 An-FVIII sequences for reconstruction and subsequent molecular, cellular, biochemical and immunological characterization. Each An-FVIII displayed activity in coagulation assays utilizing human hemophilia A plasma as a substrate thus demonstrating evolutionary mammalian compatibility. Infusion of highly purified preparations of several An-FVIIIs into hemophilia A mice also corrected the bleeding phenotype following a tail transection bleeding challenge confirming in vivo functionality. To study biosynthetic efficiency, secreted FVIII activity and mRNA transcript levels were analyzed following transfection of An-FVIII plasmids into HEK293 and BHK-M cell lines. An-53, common ancestor to rodents and primates, and An-68, ancestor to a subset of current rodents, displayed the highest FVIII biosynthetic efficiencies that were 12 and 15 fold greater than human FVIII, respectively (P = 0.002; Mann Whitney U test). These two An-FVIII sequences share 95 and 87% amino acid identity to human FVIII, respectively. In contrast, intermediate ancestors between An-53 and human FVIII, designated An-55, -56 and -57, do not display enhanced biosynthetic efficiency suggesting that the functional sequence determinant of high expression was lost during primate evolution. Predicting that high expression ancestral FVIIIs would be enabling to gene therapy approaches, An-53, An-68 and human FVIII cDNAs were placed in an AAV expression cassette under the control of a potent liver-specific promoter and the resulting plasmid DNA was infused hydrodynamically into hemophilia A mice. An-53 and An-68, but not human FVIII vector treated animals, achieved sustained, therapeutic plasma FVIII activity levels over 4 weeks (0.1 - 0.6 IU/ml versus <0.01 IU/ml, respectively). Recombinant An-FVIIIs were expressed, purified and biochemically characterized by SDS-PAGE, specific activity, decay following thrombin activation and inhibitor recognition. Early mammalian and all primate lineage thrombin-activated An-FVIII(a) displayed half-lives between 1.5 - 2.2 min that were not distinguishable from human FVIII. We have shown previously that modern murine, porcine, and ovine FVIIIa display significantly longer half-lives and thus this property may have evolved under positive selection. Supporting this conclusion, An-68 and An-78 display prolonged half-lives of 16 and 7 min, respectively. Lastly, the immune recognition of An-FVIIIs by a panel of A2 and C2 domain targeting inhibitory murine monoclonal antibodies as well as hemophilia A inhibitor patient plasmas was examined and many examples of reduced reactivity were revealed, which may enable the development of less immunogenic FVIII products. Herein, we report molecular discoveries that enhance our understanding of FVIII structure/function and provide a blueprint for bioengineering novel FVIII molecules with enhanced properties. These studies also show 'proof of concept' for ancestral protein reconstruction as a powerful approach to studying the biochemistry, molecular biology and evolution of the vertebrate coagulation system, which should enable identification of other new hematological drug targets and candidate biotherapeutics. Figure 1. Figure 1. Disclosures Spencer: Expression Therapeutics: Equity Ownership. Doering:Expression Therapeutics: Equity Ownership; Bayer Healthcare: Consultancy, Honoraria, Research Funding.
14

Okayama, Yusuke, Masato Bingo, Kazuki Sakatoku, Hiroshi Okamura, Satoru Nanno, Mitsutaka Nishimoto, Yasuhiro Nakashima, Hideo Koh, Masayuki Hino, and Hirohisa Nakamae. "The safety of the combination therapy of recombinant factor VIIa and plasma-derived factor VIIa and factor X for refractory hemorrhage in acquired hemophilia A." Blood Coagulation & Fibrinolysis 34, no. 6 (July 19, 2023): 419–22. http://dx.doi.org/10.1097/mbc.0000000000001243.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Acquired hemophilia A (AHA) is a rare, life-threatening hemorrhagic disease caused by autoantibodies against factor VIII (FVIII), and bypassing agents (BPA) are used to control bleeding. However, some cases need a change of BPA or BPAs given sequentially or in combination for refractory bleeding. A 71-year-old man was admitted with subcutaneous hemorrhage. Laboratory investigations showed prolongation of activated partial thromboplastin time (APTT) and low-coagulation FVIII activity and FVIII inhibitor; we, therefore, diagnosed AHA. He was treated with recombinant factor VIIa (rFVIIa) BPA and prednisolone. However, his symptoms did not improve sufficiently, thus we switched BPA to activated prothrombin complex concentrate. Unfortunately, this was not effective and he suffered hemorrhagic shock. Therefore, we selected rFVIIa, with plasma-derived FVIIa and factor X (pd-FVIIa/FX) as combination therapy, and hemostasis was achieved without thrombosis. This case suggests that the combination of rFVIIa and pd-FVIIa/FX short-term can be well tolerated for refractory hemorrhage in AHA.
15

Ogiwara, Kenichi, Midori Shima, and Keiji Nogami. "Factor VIII activation by factor VIIa analog (V158D/E296V/M298Q) in tissue factor-independent mechanisms." Thrombosis and Haemostasis 106, no. 10 (2011): 665–74. http://dx.doi.org/10.1160/th11-04-0264.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryFactor (F)VIIa with tissue factor (TF) is a primary trigger of blood coagulation. The recombinant (r)FVIIa analog, NN1731 (V158D/E296V/ M298Q) containing a thrombin/FIXa-mimicking catalytic domain, is ~30-fold more effective on activated platelets without TF, but ~1.2-fold with TF, than rFVIIa for FX activation. We have recently demonstrated the FVIIa/TF-dependent FVIII activation in the early coagulation phase. We assessed the action of NN1731 on FVIII activation. NN1731/TF increased FVIII activity ~2.9-fold within 30 seconds, followed by rapid inactivation, and was slightly more active than rFVIIa/TF. NN1731-catalysed activation, however, was enhanced ~6-fold at 5 minutes (min), and its peak level persisted for ~30 min. NN1731/TF proteolysed FVIII at Arg740, Arg372, and Arg336, similar to rFVIIa/TF, but cleavage by NN1731 alone was much slower at Arg336 than at Arg740 and Arg372. The Km and Vmax for NN1731/TF-catalysed activation were ~1.8-fold lower and ~2.3-fold greater than rFVIIa/TF. The Km for NN1731 alone was ~1.3-fold lower than rFVIIa, whilst the Vmax was ~7.9-fold greater, indicating that the efficiency of FVIII activation by NN1731 and NN1731/TF was ~11- and ~4-fold greater, respectively, than equivalent reactions with rFVIIa. In SPR-based assays, NN1731 bound to FVIII and the heavy chain (Kd; 0.62 and 1.9 nM) with ~1.4- and ~3.1-fold higher affinity than rFVIIa, and the A2 domain contributed to this increase. Von Willebrand factor moderated NN1731-catalysed activation more significantly than NN1731/TF. In conclusion, NN1731 was a greater potential than rFVIIa in up-regulating FVIII activity, and the TF-independent FVIII activation might represent a potential extra mode of its enhanced haemostatic effect.
16

Rao, LV, and SI Rapaport. "Factor VIIa-catalyzed activation of factor X independent of tissue factor: its possible significance for control of hemophilic bleeding by infused factor VIIa." Blood 75, no. 5 (March 1, 1990): 1069–73. http://dx.doi.org/10.1182/blood.v75.5.1069.1069.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Infusing factor VIIa (FVIIa) has been reported to control bleeding in hemophilic patients with factor VIII (FVIII) inhibitors. This is difficult to attribute to an enhanced FVIIa/tissue factor (TF) activation of factor X, since in vitro studies suggest that infusion of FVIIa should neither increase substantially the rate of formation of FVIIa/TF complexes during hemostasis (Proc Natl Acad Sci USA 85:6687, 1988) nor bypass the dampening of TF-dependent coagulation by the extrinsic pathway inhibitor (EPI) (Blood 73:359, 1989). Partial thromboplastin times have also been reported to shorten after infusion of FVIIa. The experiments reported herein establish that shortening of partial thromboplastin times after adding FVIIa to hemophilic plasma in vitro stems from an FVIIa-catalyzed activation of factor X independent of possible trace contamination of reagents with TF. Experiments in purified systems confirmed that FVIIa can slowly activate factor X in a reaction mixture containing Ca2+ and phospholipid but no source of TF. The rate of activation was sufficient to account for the shortening of partial thromboplastin times observed. EPI, which turned off continuing FVIIa/TF activation of factor X, was unable to prevent continuing FVIIa/phospholipid activation of factor X. Because circulating plasma contains only a trace, if any, free FVIIa, such a reaction could never occur physiologically. However, infusing FVIIa creates a nonphysiologic circumstance in which a continuing slow FVIIa/phospholipid catalyzed activation of factor X could conceivably proceed in vivo unimpeded by EPI. Such a mechanism of factor X activation might compensate for an impaired factor IXa/FVIIIa/phospholipid activation of factor X during hemostatis, and therefore control bleeding in a hemophilic patient.
17

Rao, LV, and SI Rapaport. "Factor VIIa-catalyzed activation of factor X independent of tissue factor: its possible significance for control of hemophilic bleeding by infused factor VIIa." Blood 75, no. 5 (March 1, 1990): 1069–73. http://dx.doi.org/10.1182/blood.v75.5.1069.bloodjournal7551069.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Infusing factor VIIa (FVIIa) has been reported to control bleeding in hemophilic patients with factor VIII (FVIII) inhibitors. This is difficult to attribute to an enhanced FVIIa/tissue factor (TF) activation of factor X, since in vitro studies suggest that infusion of FVIIa should neither increase substantially the rate of formation of FVIIa/TF complexes during hemostasis (Proc Natl Acad Sci USA 85:6687, 1988) nor bypass the dampening of TF-dependent coagulation by the extrinsic pathway inhibitor (EPI) (Blood 73:359, 1989). Partial thromboplastin times have also been reported to shorten after infusion of FVIIa. The experiments reported herein establish that shortening of partial thromboplastin times after adding FVIIa to hemophilic plasma in vitro stems from an FVIIa-catalyzed activation of factor X independent of possible trace contamination of reagents with TF. Experiments in purified systems confirmed that FVIIa can slowly activate factor X in a reaction mixture containing Ca2+ and phospholipid but no source of TF. The rate of activation was sufficient to account for the shortening of partial thromboplastin times observed. EPI, which turned off continuing FVIIa/TF activation of factor X, was unable to prevent continuing FVIIa/phospholipid activation of factor X. Because circulating plasma contains only a trace, if any, free FVIIa, such a reaction could never occur physiologically. However, infusing FVIIa creates a nonphysiologic circumstance in which a continuing slow FVIIa/phospholipid catalyzed activation of factor X could conceivably proceed in vivo unimpeded by EPI. Such a mechanism of factor X activation might compensate for an impaired factor IXa/FVIIIa/phospholipid activation of factor X during hemostatis, and therefore control bleeding in a hemophilic patient.
18

Minami, Hiroaki, Keiji Nogami, Takehisa Kitazawa, Kunihiro Hattori, and Midori Shima. "FVIII Heavy Chain Enhances Tenase Activity Induced By FVIIIa Mimicking Bispesific Antibody, ACE910." Blood 124, no. 21 (December 6, 2014): 1481. http://dx.doi.org/10.1182/blood.v124.21.1481.1481.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Background: ACE910, asymmetric bispecific monoclonal antibodies to activated factor IX (IXa) and factor X, mimics the cofactor function of activated factor VIII (VIIIa) by modulating an optimal position on the tenase assembly. The estimated therapeutic range of ACE910 shows ~30% of thrombin generation in native tenase assembly, supporting that the structure on ACE910-mimicking tenase assembly is different from that on native tenase. Being close to physiological structure consisting from factor IXa, factor X, and factor VIIIa is important for potentiating the clotting function. We examined the effects of factor VIII subunits (light chain, heavy chain, A1 and A2, C2) on ACE910-tenase. Materials/Methods: The factor VIII light chain and heavy chain were isolated from EDTA-treated recombinant factor VIII following chromatography on SP- and Q- Sepharose columns. The A2 and A1 subunits were purified from thrombin-cleaved factor VIII heavy chain by Heparin-, SP- Sepharose columns. Purified factor Xa generation assays was examined with (i) factor VIII subunit (0-40 nM), ACE910 (10 µg/ml), phospholipid (PL) (40 µM), factor IXa (1 nM) and factor X (200 nM), (ii, iii) the A2 or heavy chain (40 nM), ACE910 (10 µg/ml), PL (40 µM), factor IXa and factor X (1 or 0-80 nM, and 0-300 or 200 nM, respectively). These mixtures were reacted for five minutes (i, ii) or one minute (iii). These assays were conducted at 37 °C. Results: (i) The factor Xa generation in ACE910-tenase complex in the absence of factor VIIIa was 10.1±2.2 nM. With the intact heavy chain and A2, amounts of factor Xa were increased dose-dependently, resulting in 1.3- and 1.2-fold increases, respectively. While, the light chain and A1 subunit failed to increase at all. (ii) Vmax for factor X in ACE910-tenase was 173.0±7.0 nM and Km was 31.2±3.9 nM. Vmax obtained with the heavy chain or A2 was 175.9±6.1 or 159.0±6.1 nM, whilst Km was 17.0±2.2 or 31.9±3.5 nM, respectively, indicating that the heavy chain enhanced the binding affinity for factor X in ACE910-tenase. (iii) Vmax for factor IXa in ACE910-tenase was 43.8±2.7 nM and Km was 36.9±4.8 nM. With the heavy chain or A2, Vmax was 46.8±3.0 or 45.0±3.1 nM, and Km was 36.4±3.0 or 32.1±4.9 nM, respectively, indicating that either the heavy chain or A2 did not enhance the catalytic activity and the binding affinity for factor IXa in ACE910-tenase. Conclusion: ACE910-tenase assembly seems to be close to physiological structure by the presence of intact heavy chain interacting with factor X. In addition, ACE910 may substitute the position such as the factor VIII(a) light chain associated with FIXa and FX on ACE910-tenase assembly defecting factor VIII. Disclosures Minami: Chugai Pharmaceutical Co., Ltd.: Research Funding. Nogami:Chugai Pharmaceutical Co., Ltd.: Membership on an entity's Board of Directors or advisory committees, Research Funding. Kitazawa:Chugai Pharmaceutical Co., Ltd.: Employment, Equity Ownership, Patents & Royalties. Hattori:Chugai Pharmaceutical Co., Ltd.: Employment, Equity Ownership, Patents & Royalties. Shima:Chugai Pharmaceutical Co., Ltd.: Consultancy, Honoraria, Membership on an entity's Board of Directors or advisory committees, Research Funding.
19

Doshi, Bhavya, Courtney Cox, Bagirath Gangadharan, Christopher B. Doering, and Shannon L. Meeks. "Factor VIII Supplementation Improves Recombinant VIIa Initiated Thrombin Generation in Hemophilia A Inhibitor Patient Plasmas." Blood 118, no. 21 (November 18, 2011): 28. http://dx.doi.org/10.1182/blood.v118.21.28.28.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 28 Hemophilia A is an X-linked recessive disorder that is caused by a deficiency or defect of factor VIII (fVIII) coagulant protein. Approximately 20–30% of patients with severe hemophilia A develop antibodies (Abs) against fVIII (inhibitors) following fVIII replacement therapy, which makes bleeding episodes more difficult to control. Patients with inhibitors are treated with fVIII-bypassing agents such as recombinant factor VIIa (rfVIIa) or activated prothrombin-complex concentrate. However for unknown reasons, some patients display poor hemostatic response to bypass therapy and improved treatment options are needed. Thrombin generation assays provide an in vitro methodology for monitoring bypass therapy in hemophilia (Turecek PL et al. Pathophysiol Haemost Thromb 2003; Varadi K et al. Haemophilia 2004). Recently, it was demonstrated by us and others that combination of fVIII and by-passing agents potentiates in vitro thrombin production in hemophilia A inhibitor plasma (Klintman J et al. Br J Haematol 2010). In our study we investigated the potentiation fVIII confers to fVIIa initiated in vitro thrombin generation using a panel of anti-fVIII Abs with known epitopes. We showed that kinetics of inhibition and Ab epitope were the dominant factors influencing ability of fVIII to potentiate in vitro thrombin production. Specifically, monoclonal Abs targeting only 2 of 11 epitopes, 1 of 3 non-overlapping A2 epitopes and 1 of 2 non-overlapping C2 epitopes, inhibited thrombin generation in a manner that could not be recovered by fVIII supplementation. Here, we analyzed in vitro thrombin generation in epitope-mapped plasmas from 10 patients with hemophilia A and long-standing inhibitors after addition of fVIIa alone or in conjunction with fVIII. Methods: FVIII inhibitor plasmas from 10 patients with hemophilia A were obtained as part of an IRB approved study at the Emory Comprehensive Hemophilia Center. FVIII inhibitor titers and inhibitor kinetics were determined using a modified Bethesda assay. Samples were classified as having type II inhibitors if undiluted plasma resulted in incomplete inhibition of residual fVIII activity (Meeks SL et al. Blood 2007). Thrombin generation assays were carried out in the presence of 2.25 μg/ml recombinant fVIIa in the presence or absence of 1 U/ml recombinant full-length fVIII using reagents purchased from DiaPharma (West Chester, OH). The parameters analyzed include endogenous thrombin potential (area under thrombin generation curve), peak thrombin concentration, time to peak thrombin, lag time (time to 1/6th of peak thrombin) and index velocity (Vi-peak thrombin divided by time to peak minus lag time). Domain specific epitope mapping was carried out using direct ELISA and human/porcine domain hybrid fVIII proteins. Results: Domain mapping of the Abs in the plasmas identified 2 plasmas with predominantly anti-A2 Abs, 4 with predominantly anti-C2 Abs, 2 with both anti-A2 and anti-C2 Abs, and 2 with antibodies that were porcine fVIII cross-reactive (see Table). Plasmas with inhibitor titers less than 25 BU/ml were more responsive to fVIII supplementation with 6 of 7 having increased thrombin generation. Plasmas harboring even trace anti-A2 Abs were more resistant to increased thrombin generation with fVIII supplementation than plasmas with anti-C2 Abs alone. Conclusion: This study suggests a more favorable response to fVIII supplementation of rfVIIa may be predicted by the presence of anti-C2 Abs or inhibitory titers less than 25 BU/ml. In conjunction with our previous monoclonal Ab data, further mapping of epitopes within the fVIII A2 and C2 domains may help improve the ability to predict positive responses to fVIII supplementation of by-passing agents.PatientInhibitor Titer (BU/ml)DomainFVIII InhibitorThrombin Generation (fVIII + fVIIa vs. fVIIa)122A2Type IIIncreased242A2Type IIEqual384C2, small A2Type IEqual47C2Type IIncreased58C2Type IIIncreased620C2Type IEqual78C2Type IIncreased842C2, small A2Type IEqual922Porcine cross-reactiveType IIIncreased105.2Porcine cross-reactiveType IIncreased Disclosures: No relevant conflicts of interest to declare.
20

Wilhelm, Amelia R., Nicole A. Parsons, Benjamin J. Samelson-Jones, Robert J. Davidson, Charles T. Esmon, Rodney M. Camire, and Lindsey A. George. "Activated protein C has a regulatory role in factor VIII function." Blood 137, no. 18 (May 6, 2021): 2532–43. http://dx.doi.org/10.1182/blood.2020007562.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Mechanisms thought to regulate activated factor VIII (FVIIIa) cofactor function include A2-domain dissociation and activated protein C (APC) cleavage. Unlike A2-domain dissociation, there is no known phenotype associated with altered APC cleavage of FVIII, and biochemical studies have suggested APC plays a marginal role in FVIIIa regulation. However, the in vivo contribution of FVIIIa inactivation by APC is unexplored. Here we compared wild-type B-domainless FVIII (FVIII-WT) recombinant protein with an APC-resistant FVIII variant (FVIII-R336Q/R562Q; FVIII-QQ). FVIII-QQ demonstrated expected APC resistance without other changes in procoagulant function or A2-domain dissociation. In plasma-based studies, FVIII-WT/FVIIIa-WT demonstrated dose-dependent sensitivity to APC with or without protein S, whereas FVIII-QQ/FVIIIa-QQ did not. Importantly, FVIII-QQ demonstrated approximately fivefold increased procoagulant function relative to FVIII-WT in the tail clip and ferric chloride injury models in hemophilia A (HA) mice. To minimize the contribution of FV inactivation by APC in vivo, a tail clip assay was performed in homozygous HA/FV Leiden (FVL) mice infused with FVIII-QQ or FVIII-WT in the presence or absence of monoclonal antibody 1609, an antibody that blocks murine PC/APC hemostatic function. FVIII-QQ again demonstrated enhanced hemostatic function in HA/FVL mice; however, FVIII-QQ and FVIII-WT performed analogously in the presence of the PC/APC inhibitory antibody, indicating the increased hemostatic effect of FVIII-QQ was APC specific. Our data demonstrate APC contributes to the in vivo regulation of FVIIIa, which has the potential to be exploited to develop novel HA therapeutics.
21

Gale, Andrew J., Diana Rozenshteyn, and Justin Riceberg. "The Neutrophil Proteases, Elastase and Cathepsin G, May Modulate the Activity of Factor VIII." Blood 108, no. 11 (November 16, 2006): 1716. http://dx.doi.org/10.1182/blood.v108.11.1716.1716.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Neutrophils and monocytes express cathepsin G and elastase and also can bind to activated platelets, thus they can be localized to the site of active coagulation. Early studies suggested that cathepsin G and elastase inactivated factor VIII (FVIII) and were thus anticoagulant. But other studies have suggested procoagulant functions for cathepsin G and elastase in activation of factor V or activation of platelets among other possible mechanisms. Therefore, we investigated the effects of human neutrophil elastase and human neutrophil cathepsin G on FVIII/VIIIa. Elastase does inactivate both FVIII and FVIIIa but cathepsin G activates FVIII while having very little effect on FVIIIa. Cathepsin G activation of FVIII is enhanced by phospholipid vesicles, apparently due to enhanced rate of cleavage and stabilization of the resulting molecule. The maximum level of activation is less than that of thrombin, but it is still four-fold as measured in an APTT assay. Cleavage sites for both proteases in FVIII were identified by Edman degradation and gel analysis. FVIII cleavages are limited to a few specific sites that are mostly located near known activating and inactivating cleavage sites. A notable exception is a cleavage site for elastase after valine 26 in the A1 domain. Cathepsin G cleavage sites near to thrombin cleavage sites likely contribute to the partial activation of FVIII. The unique elastase cleavage site at valine 26 likely contributes to the inactivation of FVIII and FVIIIa. Therefore, it is possible that neutrophils and monocytes may provide some pro-coagulant effect by activating FVIII and may also provide negative feedback by inactivating FVIIIa as well.
22

Yada, Koji, and Keiji Nogami. "Novel Insights and New Developments Regarding Coagulation Revealed by Studies of the Anti-Factor IXa (Activated Factor IX)/Factor X Bispecific Antibody, Emicizumab." Arteriosclerosis, Thrombosis, and Vascular Biology 40, no. 5 (May 2020): 1148–54. http://dx.doi.org/10.1161/atvbaha.120.312919.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Emicizumab is a humanized anti-FIXa/FX (factor IXa/X) bispecific monoclonal antibody that mimics FVIIIa (activated factor VIII) cofactor function. The hemostatic efficacy of emicizumab has been confirmed in clinical studies of patients with hemophilia A, irrespective of the presence of FVIII inhibitors. Emicizumab differs in some properties from FVIIIa molecule. Emicizumab requires no activation by thrombin and is not inactivated by activated protein C, but emicizumab-mediated coagulation is regulatable and maintains hemostasis. A small amount of FIXa (activated factor IX) is required to initiate emicizumab-mediated hemostasis, whereas tissue factor/FVIIa (activated factor VII)-mediated FXa (activated factor X) and thrombin activation initiates FVIIIa-mediated hemostasis. Fibrin formation, followed by fibrinolysis, appears to be similar between emicizumab- and FVIIIa-mediated hemostasis. These results suggest possible future uses of emicizumab for treating hemorrhagic diseases other than hemophilia A and reveal previously unobservable behaviors of procoagulation and anticoagulation factors in conventional hemostasis. Here, we have reviewed novel insights and new developments regarding coagulation highlighted by emicizumab.
23

Addiego, Joseph E., Edward Gomperts, Liu Shu-Len, Patricia Bailey, Suzanne G. Courter, Martin L. Lee, Gerald G. Neslund, Henry S. Kingdon, and Michael J. Griffith. "Treatment of Hemophilia A with a Highly Purified Factor VIII Concentrate Prepared by Anti-FVIIIc Immunoaffinity Chromatography." Thrombosis and Haemostasis 67, no. 01 (1992): 019–27. http://dx.doi.org/10.1055/s-0038-1648373.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryTo reduce the risk of pathogenic virus transmission associated with the therapeutic administration of plasma-derived antihemophilic factor (FVIIIc), a process utilizing anti-FVIIIc immunoaffinity chromatography to isolate FVIIIc has been developed. In addition, the starting cryoprecipitate solution has been treated with an organic solvent/detergent mixture to inactivate lipid-enveloped viruses. A final ion exchange chromatography step is used to further remove contaminants, e.g., anti-FVIIIc antibody, potentially leached with FVIIIc during the immunoaffinity step. The purified FVTII is stabilized for lyophili-zation and storage by the addition of human albumin. The monoclonal anti-FVIIIc antibody used in the immunoaffinity step of the process is not detectable in the final preparation. Viral reduction studies performed at specific steps of the process demonstrate that 11 logs of human immunodeficiency virus (HIV) and greater than 4-5 logs of other lipid-enveloped viruses are inactivated within the first 30 s of exposure to the solvent/ detergent mixture and 4-5 logs of various model viruses, e. g. Endomyocarditis virus (EMC), are physically removed during washing of the immunoaffinity column. The lyophilized product is reconstituted using sterile water in a matter of seconds.The pharmacokinetics of Hemofil® M were compared to those obtained using a standard heat-treated concentrate (Hemofil® CT) in five severe factor VIII deficient hemophiliacs in a randomized, cross-over study. No statistically significant differences were observed in mean half life (p >0.6) or median recovery (p = 0.4) between the two preparations. No clinically significant adverse effects were observed in patients receiving either FVIII preparation.In addition, 43 patients at 18 different centers underwent pharmacokinetic studies, with a nominal dose of 50 u/kg FVIIIc Hemofil® M. The mean recovery was 103.6%, and the t 1/2 was 14.6 h. The recovery of FVIII in this group was as expected, providing an increase of assayed FVIII of approximately 2% per unit of FVTII/kg infused.Clinical trials using Hemofil® M have been initiated in 124 hemophilia A patients. The safety and efficacy of Hemofil® M has been established. To date, 0 of 60 patients tested have seroconverted to HIV. None of the previously untreated patients show clinical or laboratory evidence of Non-A, Non-B hepatitis (NANB), with 21 patients remaining negative as far as presence of antibodies to the Hepatitis C virus (a-HCV negative) at least 6 months after the initial infusion. There is no evidence of neoantigenicity, evidenced by seroconversion to murine antibody. An 8.7% (2 of 23) prevalence of anti-FVIIIc inhibitor development has been observed in previously untreated patients with FVIIIc⩽3%, receiving only the monoclonally purified solvent/ detergent treated FVIII concentrate while on study and on poststudy surveillance. All patients demonstrated clinical hemostasis following product use for either on demand bleeding or surgical prophylaxis.
24

Neuenschwander, Pierre F. "Evidence of Abnormal Coagulation: Production of Factor VIII α-Fragment In Vivo." Blood 106, no. 11 (November 16, 2005): 1020. http://dx.doi.org/10.1182/blood.v106.11.1020.1020.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Treatment of thrombosis typically involves the administration of coagulation inhibitors that must be carefully monitored and balanced so as to reduce unwanted coagulation (thrombosis) while maintaining normal or near-normal hemostasis. This balancing is necessary since the anticoagulants used alter enzymatic activities that are involved in both processes. This therapeutic strategy is based entirely on the view that thrombosis occurs by the same general pathways as normal hemostasis. While the enzymatic cascade of blood coagulation is well described and well accepted, numerous other minor reactions have been shown to occur in vitro but have not been examined in great detail due to the belief that they do not occur significantly during normal coagulation in vivo. We postulate that in certain pathological environments some of these minor procoagulant reactions may in fact become significant and lead to thrombogenic situations. If true, this could potentially allow novel targets for anticoagulation to be identified. In addition, the inhibition of these abnormal reactions could attenuate pathological coagulation whilst having limited or no effect on normal hemostatic reactions. One candidate reaction is the proteolysis of factor VIII (fVIII) by the factor VIIa-tissue factor (fVIIa-TF) complex, which results in a mixture of active and inactive fVIII molecules. We have previously shown that this reaction occurs in vitro using purified plasma components and in situ in a plasma-based system. Both of these systems produce a low level of fVIII activation with sustained (albeit low) fVIIIa activity. While it remains possible that this reaction is important in early hemostasis the elevated levels of TF in many pathological situations raises the possibility that this reaction may be more pronounced under certain circumstances in disease states. Examination of the importance of this reaction in vivo is an extremely important issue, but very difficult to address due to the inability to ascertain if fVIII activity or fragments found in vivo derive from fVIIa-TF proteolysis or proteolysis by other enzymes such as thrombin, factor Xa, or activated protein C. With this in mind we have developed an antibody reagent that can specifically detect a fVIII fragment that is a unique product of fVIII proteolysis by the fVIIa-TF complex. This antibody detects only fVIIa-TF proteolyzed fVIII (fVIII cleaved at Arg336) and its major product (α-fragment) on Western blots but not intact (unactivated) fVIII or thrombin-activated fVIII. Using this antibody we screened samples of pulmonary lavage and pleural fluid from normal patients as well as patients with acute respiratory distress syndrome, interstitial lung disease, pneumonia and lung cancer—all of which have associated procoagulant pathologies. Sandwich ELISAs of patient samples showed variably elevated levels of α-fragment (from 100 – 2000 pM) compared to normal controls (~5 pM). Western blots of lavage samples confirmed the presence of α-fragment in samples as well as the elevated levels compared to normals. These data strongly support the notion that alternative “abnormal” coagulation products can be and are generated in vivo in certain pathological settings. The data are also strongly suggestive that the fVIIa-TF complex is the most likely source of fVIII α-fragment. Although it remains unclear if fVIII α-fragment is one of the causative agents in the procoagulant pathologies of these disorders or merely an indicator of the abnormal procoagulant state, its presence in vivo indicates that the role of abnormal coagulation reactions should be further investigated.
25

Bloem, Esther, Henriet Meems, Maartje van den Biggelaar, Koen Mertens, and Alexander B. Meijer. "Factor VIII Region 1811–1818 Is Involved in Factor IXa Binding and Factor VIIIa Stability." Blood 120, no. 21 (November 16, 2012): 3357. http://dx.doi.org/10.1182/blood.v120.21.3357.3357.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 3357 Previously, we identified a role for the lysine residue couple 1967/1968 in the stability of activated factor VIII (FVIIIa). Using tandem mass tags (TMT 126/127) in combination with mass spectrometry, we identified lysine residues involved in the interaction between the A2 domain and the rest of heterodimer (A1/A3-C1-C2) of FVIIIa (Bloem et al., J Biol Chem 2012;287:5575–83). Upon FVIII activation and A2 domain dissociation, the highest increase in surface exposure occurred for the lysine couple 1967/1968, and functional studies confirmed the role thereof in FVIIIa stability. In addition to lysines 1967/1968 also other lysines displayed an increased surface exposure, including those in positions 1804, 1808, 1813 and 1818. The A3 domain region 1803–1818 has previously been implicated in interactions with ligands such as activated factor IX (FIXa). As such, one might expect increased surface exposure due to FVIII activation. On the other hand, A2 domain dissociation may have rearranged the A3 domain surface in this region. The relation between FIXa assembly and A2 domain retention was therefore explored in the present study. To unravel the role of region 1803–1818 in FVIIIa stability and FIXa binding, either region 1803–1810 or 1811–1818 was replaced by the corresponding regions of the homologous factor V. Additionally, as Asn1810 is N-linked glycosylated and this glycan is maintained in both chimeras, a variant that lacks this glycan (N1810C) was investigated. Factor × activation kinetics were used to investigate the apparent FIXa binding affinity of the FVIII variants. FXa generation was assessed on 15% phosphatidyl serine (PS) containing membranes. FIXa titration experiments showed that the affinity for the 1811–1818 variant is reduced (apparent Kd from 1.3 nM to 2.4 nM). Removal of the glycan and substitution of 1803–1810 has little or no effect on the apparent FIXa binding. FVIIIa-FIXa assembly on membranes containing 15% PS was studied using lipid- coated glass beads (lipospheres). Lipospheres were incubated with fluorescein-labeled FIXa and different FVIIIa concentrations. FIXa did only display liposphere binding in the presence of FVIIIa. Therefore, the mean fluorescent intensity on the lipospheres at increasing FVIIIa concentration could be used as a measure for FVIIIa-FIXa assembly on lipids. Results from these experiments showed that the 1811–1818 variant fails to promote FVIIIa-FIXa assembly, whereas the other variants behave like WT. To investigate FVIIIa stability, and thereby the role of the mutations on A2 domain dissociation, the activity of the variants was followed over time. Results showed that the 1811–1818 variant has a decreased half life of 1.5 min, compared to 6.9 min for WT. Also substitution of region 1803–1810 results in a lower half life, although to a lesser extent (2.8 min), whereas the glycan lacking variant behaves like WT (6.8 min). Incubation of FVIIIa variants with FIXa is known to stabilize FVIIIa and leads to an increased half life for all variants. However, the 1803–1810 variant is most efficiently stabilized by FIXa, shown by a 3-fold increase in half life, instead of 1.6-fold as seen for both WT and N1810C. The 1811–1818 variant is stabilized by 2.2-fold and therefore remains significantly less stable than WT. Together these results show that the 1811–1818 region is not only involved in FIXa binding, but additionally plays a major role in A2 domain retention. Region 1803–1810 also plays a role in FVIIIa stability, although to a lesser extent. Remarkably, the glycan at position Asn1810 does not influence neither FIXa binding nor FVIIIa stability, and apparently serves some other function. Disclosures: No relevant conflicts of interest to declare.
26

Wilhelm, Amelia R., Nicole A. Parsons, Charles T. Esmon, Rodney M. Camire, and Lindsey A. George. "In Vivo hemostatic Significance of Activated Protein C in Factor VIIIa Regulation." Blood 134, Supplement_1 (November 13, 2019): 93. http://dx.doi.org/10.1182/blood-2019-124493.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Activated factor VIII (FVIIIa) is an essential cofactor in the intrinsic tenase (Xase) enzyme complex that generates factor Xa and propagates clot formation. The FVIIIa heterotrimer is comprised of a metal ion linked dimer (A1/A3-C1-C2 domains) that is associated with the A2 domain by weak non-covalent interactions. Regulation of FXa formation by the intrinsic Xase enzyme complex occurs by FIXa inhibition and mechanisms contributing to FVIIIa inactivation, including: 1) rapid A2 domain dissociation and 2) activated protein C (APC) cleavage of FVIIIa. While FVIIIa inactivation by APC is considered important, there are surprisingly no in vivo studies documenting the hemostatic role of APC in FVIIIa regulation. Further, published data demonstrate APC cleavage of FVIIIa at physiologic protein concentrations occurs over hours while A2 dissociation occurs rapidly over minutes. Thus, it is thought that the predominant mechanism of FVIIIa inactivation is A2 dissociation and APC likely plays a marginal role in FVIIIa regulation. Additionally, unlike described A2 mutations that enhance dissociation and cause hemophilia A (HA), there is no known disease state attributed to altered FVIIIa cleavage by APC. This is in contrast to FVIII's homologous protein, FVa, whereby resistance to APC cleavage is the most common inherited thrombophilia (FV-Leiden [FVL]). Understanding the physiologically relevant mechanisms of FVIIIa inactivation has immediate clinical applicability for understanding safety considerations in HA therapeutics that bypass FVIIIa regulation (FVIII mimetic antibodies, e.g. emicizumab). Further, as evidenced by successful hemophilia B gene therapy trials using a gain of function FIX variant (FIX-Padua), altering FVIIIa inactivation could be exploited for therapeutic benefit in the setting of gene transfer. We aimed to determine the in vivo hemostatic role of APC in FVIIIa regulation and pair these studies with purified system analysis. We introduced Arg to Gln mutations at FVIII APC cleavage sites (R336Q and R562Q, herein called FVIII-QQ) on a B-domain deleted FVIII (FVIII-WT) backbone and produced recombinant FVIII-QQ and FVIII-WT. Unlike FVIII-WT, western blot analysis of FVIII-QQ incubated with APC and phospholipids had no evidence of cleavage. Enzyme kinetic studies using purified components demonstrated no appreciable difference in the Km or Vmax for FX within the intrinsic Xase enzyme complex or A2 dissociation of FVIII-QQ relative to FVIII-WT. These data confirmed no unexpected differences in FVIII-QQ relative to FVIII-WT. To begin to evaluate the role of APC in FVIIIa regulation, we measured thrombin generation in murine and human HA plasma reconstituted with FVIII-QQ or FVIII-WT in the presence of increasing APC concentrations. The IC50 of APC was 2-3-fold higher for FVIII-QQ than FVIII-WT. To evaluate the in vivo hemostatic effect of APC in FVIIIa regulation, HA mice were infused with FVIII-QQ or FVIII-WT and evaluated by tail clip injury and 7.5% FeCl3 carotid artery occlusion models. Required doses of FVIII-QQ to normalize blood loss from a tail clip assay and time to vessel occlusion in a FeCl3 assay were 4-5 fold lower than necessary FVIII-WT doses; the superior hemostatic effect of FVIII-QQ supported the physiologic significance of APC in FVIIIa inactivation. To isolate the role of APC in FVIIIa regulation from APC inactivation of FVa, we backcrossed HA mice with FVL mice to create homozygous HA/FVL mice. HA/FVL mice were infused with FVIII-QQ or FVIII-WT and underwent tail clip assay analysis. Doses of FVIII-QQ required to normalize blood loss were again less than FVIII-WT. To further isolate the enhanced hemostatic effect of FVIII-QQ to APC resistance, we performed the tail clip assay in HA/FVL mice infused with FVIII-QQ or FVIII-WT in the presence or absence of MPC1609, an antibody that blocks murine APC function (Xu et al. J Thromb Haemost 2008). In the presence of MPC1609, the same dose of FVIII-WT and FVIII-QQ was required to normalize blood loss (Figure 1). Collectively, our in vitro and in vivo data support the physiologic significance of APC in FVIIIa regulation. To our knowledge these data are the first to demonstrate the in vivo hemostatic effect of APC in FVIIIa inactivation. Our data may be translated to rationally exploit APC regulation of FVIIIa to develop novel HA therapeutics or further delineate safety considerations in therapies that bypass FVIIIa regulation. Figure 1 Disclosures Camire: Pfizer: Research Funding. George:University of Pennyslvania: Employment; Pfizer: Consultancy; Avrobio: Membership on an entity's Board of Directors or advisory committees.
27

Healey, John F., Ernest Parker, and John (Pete) S. Lollar. "Comparative Decay Rates of Human, Rhesus Macaque, Cynomolgus, and Porcine Activated Factor VIII." Blood 114, no. 22 (November 20, 2009): 3164. http://dx.doi.org/10.1182/blood.v114.22.3164.3164.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Abstract 3164 Poster Board III-104 The proteolytic conversation by thrombin of factor VIII (fVIII) to fVIIIa produces a A1/A2/A3-C1-C2 heterotrimer that spontaneously dissociates into inactive A1/A3-C1-C2 and A2 species. Human mutations that increase the rate of A2 subunit dissociation produce hemophilia A, indicating that A2 subunit dissociation is physiologically relevant and is an important regulatory feature of the blood coagulation mechanism. The A2 subunit dissociation rate from human fVIIIa is significantly faster than the corresponding dissociation rates from porcine or murine fVIIIa. The fast decay rate of human fVIIIa raises the question whether the f8 gene is under positive selection for this trait. To determine whether fast A2 dissociation occurs elsewhere in the primate lineage, we cloned cDNAs encoding B-domain deleted (BDD) fVIII from rhesus macaque and cynomolgus monkey liver. The deduced BDD amino acid sequences of rhesus and cynomolgus fVIII were 97.9 % and 98% identical to human fVIII, respectively, and were 99.9% identical to each other. The expression of rhesus and cynomolgus fVIII from baby hamster kidney-derived cells was similar to human fVIII and ten-fold lower than porcine fVIII. BDD human, rhesus, cynomolgus, and porcine fVIII molecules were purified to homogeneity by tandem ion-exchange chromatography. Concentrations of the purified constructs were calculated using a molar extinction coefficient at 280 nm based on their predicted tyrosine, tryptophan and cysteine compositions. Human, rhesus, and cynomolgus fVIII displayed similar specific coagulant activities by one-stage coagulation assay (6800, 4500, and 5200 units per mg, respectively). The kinetics of decay of human, rhesus, cynomolgus and porcine fVIIIa were measured following rapid activation of 1 nM fVIII by thrombin using a chromogenic substrate assay of purified intrinsic fXase complex under conditions in which fVIIIa was limiting. Decay curves were fit using nonlinear least-squares regression to a first-order model (Fig. 1). Decay rate constants for rhesus and cynomolgus fVIIIa were similar (0.31 and 0.27 min-1, respectively) and were slightly, but significantly lower than human fVIIIa (0.40 min-1). In contrast, the decay rate constant for porcine fVIIIa, 0.17 min-1, was 2.3-fold lower than human fVIIIa, consistent with previous observations. These results suggest that fast A2 subunit dissociation rates evolved before evolution of the primate lineage. Disclosures No relevant conflicts of interest to declare.
28

Rozenshteyn, Diana, and Andrew J. Gale. "Cathepsin G, a leukocyte protease, activates coagulation factor VIII." Thrombosis and Haemostasis 99, no. 01 (2008): 44–51. http://dx.doi.org/10.1160/th07-08-0495.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryNeutrophils and monocytes express cathepsin G and can also bind to activated platelets, thus they can be localized to the site of active coagulation. Previous studies have suggested that cathepsin G inactivated coagulation factorVIII (FVIII) and was thus anticoagulant. But other studies have indicated procoagulant functions for cathepsin G in activation of coagulation factorV or activation of platelets among other possible mechanisms. Therefore, it remains unclear if cathepsin G is anticoagulant or procoagulant. We investigated the effects of human neutrophil cathepsin G on FVIII/VIIIa. Cathepsin G activates FVIII to a partially active form while having only a minor inactivating effect on thrombin- activated FVIIIa. This inactivation is mostly due to decreased stability of FVIIIa since a disulfide bond that prevents A2 subunit dissociation from FVIIIa prevents any loss of activity due to cathepsin G proteolysis. FVIII that has been cleaved by cathepsin G can still be activated by thrombin if A2 subunit dissociation is prevented. Cathepsin G cleavages of FVIII are limited to a few specific sites that are mostly located near known activating and inactivating cleavage sites. Cathepsin G cleavage sites near to thrombin cleavage sites likely contribute to the partial activation of FVIII. Therefore, it is possible that cathepsin G from neutrophils and monocytes may provide some pro-coagulant effect by activating FVIII.
29

van 't Veer, C., TM Hackeng, C. Delahaye, JJ Sixma, and BN Bouma. "Activated factor X and thrombin formation triggered by tissue factor on endothelial cell matrix in a flow model: effect of the tissue factor pathway inhibitor." Blood 84, no. 4 (August 15, 1994): 1132–42. http://dx.doi.org/10.1182/blood.v84.4.1132.1132.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract The procoagulant subcellular matrix of stimulated endothelial cells that contains tissue factor (TF) was used to investigate the mechanism by which TF pathway inhibitor (TFPI) inhibits thrombin formation initiated by TF/factor VIIa (FVIIa) under flow conditions. Purified coagulation factors VII, X, and V and prothrombin were perfused at a wall shear rate of 100 s-1 through a flow chamber containing a coverslip covered with matrix of cultured human umbilical vein endothelial cells. This resulted in a TF- and FVII-dependent FXa and thrombin generation as measured in the effluent at the outlet of the system. Inhibition of this TF/FVIIa-triggered thrombin formation by TFPI purified from plasma was dependent on the amount of TF present on the endothelial cell matrix. The rate of prothrombinase assembly and steady-state levels of thrombin formation were decreased by TFPI. Because persistent albeit decreased steady-state levels of thrombin formation occurred in the presence of TFPI, we conclude that plasma- TFPI does not inhibit FXa present in the prothrombinase complex. The addition of FIX and FVIII to perfusates containing FVII and FX increased the FXa generation on endothelial matrices, and counteracted the inhibition of thrombin formation on endothelial cell matrices by TFPI. Our data provide further evidence for the hypothesis that the rapid inactivation of TF/FVIIa by TFPI in combination with the absence of either FVIII or FIX causes the bleeding tendency of patients with hemophilia A or B.
30

van 't Veer, C., TM Hackeng, C. Delahaye, JJ Sixma, and BN Bouma. "Activated factor X and thrombin formation triggered by tissue factor on endothelial cell matrix in a flow model: effect of the tissue factor pathway inhibitor." Blood 84, no. 4 (August 15, 1994): 1132–42. http://dx.doi.org/10.1182/blood.v84.4.1132.bloodjournal8441132.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
The procoagulant subcellular matrix of stimulated endothelial cells that contains tissue factor (TF) was used to investigate the mechanism by which TF pathway inhibitor (TFPI) inhibits thrombin formation initiated by TF/factor VIIa (FVIIa) under flow conditions. Purified coagulation factors VII, X, and V and prothrombin were perfused at a wall shear rate of 100 s-1 through a flow chamber containing a coverslip covered with matrix of cultured human umbilical vein endothelial cells. This resulted in a TF- and FVII-dependent FXa and thrombin generation as measured in the effluent at the outlet of the system. Inhibition of this TF/FVIIa-triggered thrombin formation by TFPI purified from plasma was dependent on the amount of TF present on the endothelial cell matrix. The rate of prothrombinase assembly and steady-state levels of thrombin formation were decreased by TFPI. Because persistent albeit decreased steady-state levels of thrombin formation occurred in the presence of TFPI, we conclude that plasma- TFPI does not inhibit FXa present in the prothrombinase complex. The addition of FIX and FVIII to perfusates containing FVII and FX increased the FXa generation on endothelial matrices, and counteracted the inhibition of thrombin formation on endothelial cell matrices by TFPI. Our data provide further evidence for the hypothesis that the rapid inactivation of TF/FVIIa by TFPI in combination with the absence of either FVIII or FIX causes the bleeding tendency of patients with hemophilia A or B.
31

Mazurier, Claudine, Armelle Parquet-Gernez, and Maurice Goudemand. "Validation of a Procedure for Potency Assessing of a High Purity Factor VIII Concentrate -Comparison of Different Factor VIII Coagulant Assays and Effect of Prediluent." Thrombosis and Haemostasis 64, no. 02 (1990): 251–55. http://dx.doi.org/10.1055/s-0038-1647295.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryThe assessment of factor VIII coagulant activity (FVTII: C) in recently available highly purified and concentrated FVTII therapeutic products calls for careful evaluation of assay methodologies. We assayed more than 130 batches of a concentrate with a specific activity of about 150 FVTII :C units/mg protein, using one-stage and two-stage clotting and chromogenic methods. There was good agreement between the potency estimates obtained with the different methods. We also compared the FVTII :C potencies obtained after predilution in buffer or FVIII-deficient plasma using either calibrated plasma or FVTII concentrate as references. With the one-stage assay we found a marked discrepancy between the potency values obtained with buffer and with FVTII-deficient plasma used as prediluents. In order to validate our “in vitro” data we performed 6 “in vivo” analyses in severe haemophilia A patients. On the basis of the overall data obtained we chose to label FVIII potency by using FVIII-deficient plasma as prediluent, reference plasma as standard and the chromogenic assay method.
32

Takeyama, Masahiro, Keiji Nogami, Tomoko Matsumoto, Tetsuhiro Soeda, Tsukasa Suzuki, Kunihiro Hattori, and Midori Shima. "Characterisation of an antibody specific for coagulation factor VIII that enhances factor VIII activity." Thrombosis and Haemostasis 103, no. 01 (2010): 94–102. http://dx.doi.org/10.1160/th09-05-0338.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryMany reports have identified factor (F)VIII inhibitory antibodies with epitopes located in all subunits of the FVIII molecule. Antibodies that promote FVIII activity do not appear to have been reported. We characterised, for the first time, a unique anti-FVIII monoclonal antibody, mAb216, that enhanced FVIII coagulant activity. The mAb216 shortened the activated partial thromboplastin time and specifically increased FVIII activity by ~1.5-fold dose-dependently. FXa generation and thrombin generation were similarly increased by ~1.4- and ~2.5-fold, respectively. An A2 epitope, not overlapping the common A2 epitope, was identified and the antibody was shown to enhance thrombin (and FXa)-catalysed activation of FVIII by modestly accelerating cleavage at Arg372. The presence of mAb216 mediated an ~1.5-fold decrease in Km for the FVIII-thrombin interaction. Enhanced FVIII activity was evident to an equal degree, even the presence of anti-FVIII neutralising antibodies with epitopes in each subunit. In addition, mAb216 depressed the rates of heat-denatured loss of FVIII activity and FVIIIa decay by 2 to ~2.5-fold. We have developed an anti-A2, FVIII mAb216 that augmented procoagulant activity. This enhancing effect could be attributed to an increase in thrombin-induced activation of FVIII, mediated by cleavage at Arg372 and a tighter interaction of thrombin with the A2 domain. The findings may cast new light on new principles for improving the treatment of haemophilia A patients.
33

Peerlinck, Kathelijne, Marc G. Jacquemin, Jef Arnout, Marc F. Hoylaerts, Jean Guy G. Gilles, Renaud Lavend’homme, Karen M. Johnson, et al. "Antifactor VIII Antibody Inhibiting Allogeneic but not Autologous Factor VIII in Patients With Mild Hemophilia A." Blood 93, no. 7 (April 1, 1999): 2267–73. http://dx.doi.org/10.1182/blood.v93.7.2267.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Two unrelated patients with the same Arg2150His mutation in the factor VIII (FVIII) C1 domain, a residual FVIII activity of 0.09 IU/mL, and inhibitor titres of 300 and 6 Bethesda Units, respectively, were studied. Further analysis of patient LE, with the highest inhibitor titer, showed that (1) plasma or polyclonal IgG antibodies prepared from LE plasma inhibited the activity of allogeneic (wild-type) but not of self FVIII; (2) the presence of von Willebrand factor (vWF) increased by over 10-fold the inhibitory activity on wild-type FVIII; (3) the kinetics of FVIII inhibition followed a type II pattern, but in contrast to previously described type II inhibitors, LE IgG was potentiated by the presence of vWF instead of being in competition with it; (4) polyclonal LE IgG recognized the FVIII light chain in enzyme-linked immunosorbent assay and the recombinant A3-C1 domains in an immunoprecipitation assay, indicating that at least part of LE antibodies reacted with the FVIII domain encompassing the mutation site; and (5) LE IgG inhibited FVIII activity by decreasing the rate of FVIIIa release from vWF, but LE IgG recognized an epitope distinct from ESH8, a murine monoclonal antibody exhibiting the same property. We conclude that the present inhibitors are unique in that they clearly distinguish wild-type from self, mutated FVIII. The inhibition of wild-type FVIII by LE antibody is enhanced by vWF and is associated with an antibody-dependent reduced rate of FVIIIa release from vWF.
34

Peerlinck, Kathelijne, Marc G. Jacquemin, Jef Arnout, Marc F. Hoylaerts, Jean Guy G. Gilles, Renaud Lavend’homme, Karen M. Johnson, et al. "Antifactor VIII Antibody Inhibiting Allogeneic but not Autologous Factor VIII in Patients With Mild Hemophilia A." Blood 93, no. 7 (April 1, 1999): 2267–73. http://dx.doi.org/10.1182/blood.v93.7.2267.407k21_2267_2273.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Two unrelated patients with the same Arg2150His mutation in the factor VIII (FVIII) C1 domain, a residual FVIII activity of 0.09 IU/mL, and inhibitor titres of 300 and 6 Bethesda Units, respectively, were studied. Further analysis of patient LE, with the highest inhibitor titer, showed that (1) plasma or polyclonal IgG antibodies prepared from LE plasma inhibited the activity of allogeneic (wild-type) but not of self FVIII; (2) the presence of von Willebrand factor (vWF) increased by over 10-fold the inhibitory activity on wild-type FVIII; (3) the kinetics of FVIII inhibition followed a type II pattern, but in contrast to previously described type II inhibitors, LE IgG was potentiated by the presence of vWF instead of being in competition with it; (4) polyclonal LE IgG recognized the FVIII light chain in enzyme-linked immunosorbent assay and the recombinant A3-C1 domains in an immunoprecipitation assay, indicating that at least part of LE antibodies reacted with the FVIII domain encompassing the mutation site; and (5) LE IgG inhibited FVIII activity by decreasing the rate of FVIIIa release from vWF, but LE IgG recognized an epitope distinct from ESH8, a murine monoclonal antibody exhibiting the same property. We conclude that the present inhibitors are unique in that they clearly distinguish wild-type from self, mutated FVIII. The inhibition of wild-type FVIII by LE antibody is enhanced by vWF and is associated with an antibody-dependent reduced rate of FVIIIa release from vWF.
35

Ponder, Katherine P. "FIXing Factor VIII inhibitors." Blood 119, no. 2 (January 12, 2012): 325–26. http://dx.doi.org/10.1182/blood-2011-11-389486.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
In this issue of Blood, Milanov and colleagues demonstrate that a Factor IX (FIX) variant that does not require activated Factor VIII (FVIIIa) for activity induces coagulation in hemophilia A mice with FVIII inhibitors.1 This protein might be developed as a bypass agent.
36

Zhong, Degang, Evgueni L. Saenko, Midori Shima, Matthew Felch, and Dorothea Scandella. "Some Human Inhibitor Antibodies Interfere With Factor VIII Binding to Factor IX." Blood 92, no. 1 (July 1, 1998): 136–42. http://dx.doi.org/10.1182/blood.v92.1.136.413k35_136_142.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Factor VIII (fVIII) functions as a cofactor of factor IXa in the intrinsic pathway of blood coagulation. Its absence or abnormality causes the bleeding disorder hemophilia A. About 23% of hemophiliacs who receive therapeutic fVIII infusions develop antibodies that inhibit its activity. We previously showed by inhibitor neutralization assays that the fVIII A2 and C2 domain polypeptides contain common inhibitor epitopes. Often hemophilic inhibitor plasmas were partially neutralized by C2 and more completely neutralized by fVIII light chain (A3-C1-C2), suggesting the presence of an additional major inhibitor epitope(s) within the A3-C1 domains. In immunoprecipitation assays, 17 of 18 inhibitor IgGs bound to recombinant 35S-A3-C1. Amino acids 1811-1818 of the A3 domain comprise a binding site for factors IX and IXa. Three inhibitor IgGs prevented binding of factor IXa to fVIII light chain, and the binding of each IgG to light chain was competed by A3 peptide 1804-1819. The generation of factor Xa by the fVIIIa/fIXa complex in a chromogenic assay was prevented by these inhibitors. Therefore, we propose that another important mechanism of fVIII inactivation by human inhibitors is the prevention of fVIIIa/fIXa association.
37

Hsu, Ting-Chang, Kathleen P. Pratt, and Arthur R. Thompson. "The factor VIII C1 domain contributes to platelet binding." Blood 111, no. 1 (January 1, 2008): 200–208. http://dx.doi.org/10.1182/blood-2007-01-068957.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Activated factor VIII (FVIIIa) forms a procoagulant complex with factor IXa on negatively charged membranes, including activated platelet surfaces. Membrane attachment involves the FVIII C2 domain; involvement of the adjacent C1 domain has not been established. Binding of recombinant FVIII C1C2 and C2 proteins to platelets was detected by flow cytometry using (1) anti-C2 monoclonal antibody ESH8 followed by a phycoerythrin-labeled secondary antibody; (2) biotinylated C1C2 detected by phycoerythrin-labeled streptavidin, and (3) C1C2 and C2 site-specifically labeled with fluorescein. Highest binding and lowest background were obtained using fluorescein-conjugated proteins. More than 90% of activated platelets bound C1C2, compared with approximately 50% for equimolar C2. Estimates using fluorescent microbeads indicated approximately 7000 C1C2-binding sites per platelet, approximately 1400 for C2, and approximately 3000 for fluorescein-labeled FVIIIa. Unlike C2 or FVIII(a), C1C2 bound to approximately 700 sites/platelet before activation. C1C2 binding to activated platelets appeared independent of von Willebrand factor and was competed effectively by FVIII(a), but only partially by excess C2. Fluorescein-labeled FVIIIa was competed much more effectively by C1C2 than C2 for binding to activated platelets. Two monoclonal antibodies that inhibit C2 binding to membranes competed platelet binding of C2 more effectively than C1C2. Thus, the C1 domain of FVIII contributes to platelet-binding affinity.
38

Grushin, Kirill, Jaimy Miller, Daniela Dalm, and Svetla Stoilova-McPhie. "Factor VIII organisation on nanodiscs with different lipid composition." Thrombosis and Haemostasis 113, no. 04 (July 2015): 741–49. http://dx.doi.org/10.1160/th14-09-0725.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryNanodiscs (ND) are lipid bilayer membrane patches held by amphiphilic scaffolding proteins (MSP) of ~10 nm in diameter. Nanodiscs have been developed as lipid nanoplatforms for structural and functional studies of membrane and membrane associated proteins. Their size and monodispersity have rendered them unique for electron microscopy (EM) and single particle analysis studies of proteins and complexes either spanning or associated to the ND membrane. Binding of blood coagulation factors and complexes, such as the Factor VIII (FVIII) and the Factor VIIIa - Factor IXa (intrinsic tenase) complex to the negatively charged activated platelet membrane is required for normal haemostasis. In this study we present our work on optimising ND, specifically designed to bind FVIII at close to physiological conditions. The binding of FVIII to the negatively charged ND rich in phosphatidylserine (PS) was followed by electron microscopy at three different PS compositions and two different membrane scaffolding protein (MSP1D1) to lipid ratios. Our results show that the ND with highest PS content (80 %) and lowest MSP1D1 to lipid ratio (1:47) are the most suitable for structure determination of the membrane-bound FVIII by single particle EM. Our preliminary FVIII 3D reconstruction as bound to PS containing ND demonstrates the suitability of the optimised ND for structural studies by EM. Further assembly of the activated FVIII form (FVIIIa) and the whole FVIIIa-FIXa complex on ND, followed by EM and single particle reconstruction will help to identify the protein-protein and protein-membrane interfaces critical for the intrinsic tenase complex assembly and function.
39

Hakeos, William, Hongzhi Miao, Nongnuch Sirachainan, Geoffrey Kemball-Cook, Evgueni Saenko, Randal Kaufman, and Steven Pipe. "Hemophilia A Mutations within the Factor VIII A2-A3 Subunit Interface Destabilize Factor VIIIa and Cause One-stage/ Two-stage Activity Discrepancy." Thrombosis and Haemostasis 88, no. 11 (2002): 781–87. http://dx.doi.org/10.1055/s-0037-1613302.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryThrombin-activated factor VIII (FVIIIa) is a heterotrimer with the A2 subunit in a weak ionic interaction with the A1 and A3-C1-C2 subunits. Dissociation of the A2 subunit correlates with inactivation of FVIIIa. A homology model (Blood 89:2413, 1997) of the triplicated A domains of factor VIII (FVIII) predicts a pseudo-threefold axis at the tightly packed hydrophobic core with several interdomain interactions. These lie at the interface of A1-A2, A2-A3 and A1-A3. We have previously demonstrated that hemophilia A mutations (R531H, A284E, S289L) within the predicted A1-A2 and A1-A3 interface disrupt potential intersubunit hydrogen bonds and have the molecular phenotype of increased rate of inactivation of FVIIIa due to increased rate of A2 subunit dissociation. Patients with these mutations exhibit a clinical phenotype where the FVIII activity by one-stage(1-st) assay is at least two-fold higher than by two-stage(2-st) assay. We have now also explored mutations within the predicted A2-A3 interface (N694I, R698W and R698L) that also have the phenotype of 1-st/2-st activity discrepancy. These mutations exhibit the same molecular mechanism of increased instability of FVIIIa as those mutations described along the A1-A2 and A1-A3 interfaces. This suggests that the entire tightly packed hydrophobic core within the predicted pseudo-threefold axis contributes to stabilization of FVIIIa.
40

Strijbis, Viola J. F., Ka Lei Cheung, Pavlina Konstantinova, Ying Poi Liu, Sander J. van Deventer, and Mettine H. A. Bos. "Evaluation of a Blood Coagulation Factor IX Variant That Functions Independently of Factor VIII As an Alternative Treatment for Hemophilia A." Blood 134, Supplement_1 (November 13, 2019): 1110. http://dx.doi.org/10.1182/blood-2019-124120.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
The serine protease factor IXa (FIXa) serves an important role in coagulation by catalyzing the proteolytic activation of factor X (FX) together with its cofactor VIIIa (FVIIIa). Being a critical protease in coagulation, the FIXa structure has evolved to be subjected to strict regulatory mechanisms. While FIXa displays considerable structural homology with other coagulation serine proteases, its active site is uniquely controlled by the 99-loop that blocks access to the active site pocket. Cofactor-mediated interaction of FIXa with its substrate FX induces a conformational change that allows for active site engagement and substrate catalysis. Previously, the molecular constraints of the 99-loop were lifted due to specific modifications in both the 99-loop (K265A), the S1 active site subpocket (V181I, I383V), and the L6F substitution, thereby generating FIX-FIAV [Quade-Lyssy et al. J. Thromb. Haemost. 2014]. As a result, this variant is capable of functioning independently of factor VIII (FVIII). Moreover, FIX-FIAV was demonstrated to ameliorate the hemophilia A phenotype both in vitro and in vivo. To further evaluate its therapeutic potential, FIX-FIAV was stably expressed in HEK293 cells and purified by ion-exchange and hydrophobic interaction chromatography. Evaluation of the kinetics of tissue factor-factor VIIa (TF-FVIIa) activation of FIX-FIAV revealed kinetic parameters similar to those of human wild-type FIX(-WT). Analysis of FIX activation intermediates that are formed upon proteolysis by TF-FVIIa or factor XIa demonstrated prolonged formation of FIX-FIAVα, while no FIXa-WTα could be observed. This is consistent with delayed cleavage at position 180, likely resulting from the V181I substitution in FIX-FIAV. Given that the activation mechanism of FIX-FIAV is unperturbed, we next assessed the specific FVIII clotting activity and demonstrated that FIX-FIAV exhibited significant FVIII-like clotting activity (56 ± 4 U/mg) as opposed to FIX-WT (<13 U/mg). These values correlate with up to 28% of FVIII-independent activity for FIX-FIAV at FIX plasma levels (5 ug/mL), confirming that FIX-FIAV has the potential to enhance thrombin generation in FVIII deficiency. To validate this, tissue factor-initiated (0.5 or 1.0 pM) thrombin generation was assessed in FVIII-immunodepleted plasma, leading to a severely reduced thrombin peak (88% or 81% reduction, respectively) relative to conditions with 100% FVIII. Addition of FIX-FIAV (5 ug/mL) partially restored thrombin generation, demonstrated by an up to ~30% increase in both thrombin peak and endogenous thrombin potential. Evaluation of the FVIII-independent activity of FIX-FIAV in severe hemophilia A patient plasma with or without an inhibitor resulted in an up to 18% or 32% FVIII-like activity, respectively, demonstrating efficacy of FIX-FIAV in the presence of FVIII inhibitors. Although unlikely, it remains to be determined whether specific FVIII-inhibitors may impact FIX-FIAV function. Adding 100% FVIII or low- to mid-range therapeutic concentrations of the bispecific antibody emicizumab to FVIII-deficient plasma incubations with FIX-FIAV resulted in a synergistic enhancement of thrombin generation, demonstrated by a 9-fold increase in thrombin peak. This is consistent with the previously demonstrated hyperactivity of FIX-FIAV in a cofactor-dependent system. In contrast, no synergistic effect on thrombin generation was observed when combining FIX-FIAV with physiologically relevant concentrations of FEIBA or NovoSeven. Summarizing, FIX-FIAV is characterized by a preserved mechanism of activation in addition to being capable of sustaining therapeutic levels of coagulation activity in FVIII deficiency. This provides support for the use of FIX-FIAV as an alternative treatment for hemophilia A. Disclosures Strijbis: uniQure Biopharma B.V.: Research Funding. Konstantinova:uniQure Biopharma B.V.: Employment. Liu:uniQure Biopharma B.V.: Employment. van Deventer:uniQure Biopharma B.V.: Employment. Bos:uniQure Biopharma B.V.: Membership on an entity's Board of Directors or advisory committees, Research Funding.
41

Sen, Prosenjit, Curtis A. Clark, Ramakrishnan Gopalakrishnan, Ulla Hedner, Charles T. Esmon, Usha R. Pendurthi, and L. Vijaya Rao. "Factor VIIa binding to endothelial cell protein C receptor: Differences between mouse and human systems." Thrombosis and Haemostasis 107, no. 05 (2012): 951–61. http://dx.doi.org/10.1160/th11-09-0672.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryRecent in vitro studies have shown that the zymogen and activated form of factor (F)VII bind to endothelial cell protein C receptor (EPCR). At present, there is no evidence that FVIIa binds to EPCR on vascular endothelium in vivo in the presence of circulating protein C, a primary ligand for EPCR. The present study was carried out to investigate the interaction of murine and human ligands with murine EPCR both in vivo and in vitro. Measurement of endogenous plasma levels of FVII in wild-type, EPCR-deficient and EPCR-over expressing mice showed slightly lower levels of FVII in EPCR-over expressing mice. However, infusion of high concentrations of competing ligands, either human APCi or FVIIai, to EPCR-over expressing mice failed to increase plasma levels of mouse FVII whereas they increased the plasma levels of protein C by two- to three-fold. Examining the association of exogenously administered mouse FVIIa or human FVIIa by immunohistochemistry revealed that human, but not murine FVIIa, binds to the murine endothelium in an EPCR-dependent manner. In vitro binding studies performed using surface plasmon resonance and endothelial cells revealed that murine FVIIa binds murine EPCR negligibly. Human FVIIa binding to EPCR, particularly to mouse EPCR, is markedly enhanced by availability of Mg2+ ions. In summary, our data show that murine FVIIa binds poorly to murine EPCR, whereas human FVIIa binds efficiently to both murine and human EPCR. Our data suggest that one should consider the use of human FVIIa in mouse models to investigate the significance of FVIIa and EPCR interaction.
42

Lannert, Kerry W., Hilary S. Gammill, Barbara A. Konkle, and Jill M. Johnsen. "The Ability of Von Willebrand Factor (VWF) to Bind Factor VIII (FVIII) Can Decrease during Pregnancy." Blood 124, no. 21 (December 6, 2014): 4244. http://dx.doi.org/10.1182/blood.v124.21.4244.4244.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Background: The coagulation protein von Willebrand Factor (VWF) and its circulating partner protein, Factor VIII, are known to be elevated in pregnancy. However, the nature of changes in VWF which occur during pregnancy are not well understood. We previously presented results from a study of 46 healthy pregnancies in which we measured VWF antigen (VWF:Ag) and Factor VIII (FVIII determined both by activity and antigen measurements). We found that VWF:Ag and FVIII increased during pregnancy, consistent with less FVIII relative to VWF. We have also observed that VWF multimer size can decrease during pregnancy. We hypothesized that pregnancy can induced molecular changes in VWF, and that such changes could impact the interaction of VWF with its ligands, specifically FVIII. Methods: We adapted a Type 2N von Willebrand Disease (VWD) assay to test the ability of plasma VWF to bind recombinant FVIII. In brief, serial dilutions of plasma VWF were captured using a polyclonal anti-VWF antibody (DAKO), endogenous FVIII was stripped using a CaCl2 wash, recombinant FVIII (Recombinate) was applied, and bound FVIII was detected using an anti-FVIII antibody (Cedarlane). In parallel, a VWF:Ag ELISA was performed using the same anti-VWF capture antibody. The VWF-FVIII binding ratio (VWF-FVIIIb) was then determined by dividing the quantity of rFVIII detected by the amount of VWF:Ag immobilized on the plate. The range, reproducibility, and variance of the assay was first determined in a healthy blood donor repository. We then tested VWF-FVIIIb in healthy pregnancies for which 3rd trimester and non-pregnant baseline samples were available. Additional time points during pregnancy were also studied, when available. VWF-FVIIIb results for each trimester and the non-pregnant baseline were analyzed by one-way ANOVA (All Pairwise Multiple Comparison Procedures, Holm-Sidak method). Results: Forty-six healthy pregnancies with non-pregnant baseline samples were studied. Samples tested were drawn in the first (n=6), second (n=38), and third (n=43) trimesters, at 38 weeks gestational age (n=13), and at a non-pregnant time point (baseline). In non-pregnant samples, the VWF-FVIIIb was close to 1.0 (0.988 +/- 0.084), similar to a normal blood donor repository. We identified a significant decrease in VWF-FVIIIb in the second trimester (0.931 +/- 0.82; p = 0.013), which became more pronounced in the third trimester (0.899 +/- 0.75; p < 0.001) and at 38 weeks gestational age (0.88 +/- 0.072; p < 0.001). The means, quartiles, and standard deviations for each pregnancy time point are shown in Figure 1. Conclusions: These data support the hypothesis that VWF can acquire a decreased ability to interact with FVIII during pregnancy. A decreased capacity to carry FVIII could contribute to the increased VWF:FVIII ratio we and others have observed during pregnancy, although other factors, such as increased VWF production relative to FVIII, could also play a role. This work provides further evidence for a new model of acquired VWF changes in the setting of pregnancy. Disclosures No relevant conflicts of interest to declare.
43

Rolli, Veronique, Nathalie Enjolras, Cecile Ducasse, Marie-Helene Rodriguez, Thomas Weimer, Hans-Peter Hauser, Claude Negrier, and Jean-Luc Plantier. "Generation of Factor VIII Molecules Partially Resistant to Activated Protein C." Blood 104, no. 11 (November 16, 2004): 1729. http://dx.doi.org/10.1182/blood.v104.11.1729.1729.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Following a vascular injury, factor VIII (FVIII) is rapidly activated by thrombin cleavage at arginine (Arg) 372, 740 and 1689. Activated FVIII, an heterotrimer composed by the association of A1, A2 and A3-C1-C2 domains, is rapidly degraded to limit thrombosis risk. Two main phenomenons that account for the disappearence of FVIIIa consist of an intrinsic dissociation of the trimer due to the loss of A2 domain, and the cleavage of the molecule by activated protein C (APC). APC cleaves FVIIIa at Arg 336 and 562. Mutant FVIII molecules were already generated, with one or two arginines substituted, and a subsequent APC cleavage diminished. Since the thrombotic potential of high plasma levels of FVIII has been described, we aimed to generate new factor VIII molecules where the APC cleavage was only modulated. The ultimate goal being to increase in vivo the FVIIIa half-life. Among the two APC cleavage sites, the sequence around the site 562 was the most conserved between species. This region was therefore chosen to be modified with a prior verification that the amino-acids to be modified were not described in any hemophilic phenotype. Subsequently, the six following mutations were realized in a BDD-FVIII cDNA: Q561N, G563A, N564D, N564A, N564Q and I566M as well as the controls R336I, R562K and R336I+R562K. The constructs were transiently expressed in BHK cells to determine the specific activities of the corresponding molecules. The mutant specific activities, as determined by one- and two-stage clotting assays, ranged from 40 to 94 % of the wild-type except for G563A that was almost inactive. CHO clones expressing FVIII molecules were obtained. The mutants and control FVIII molecules were produced, partially purified on heparin column and further analyzed. The comparative specific activities in a two-stage clotting assay were the following (+/− SD; n=6): BDD-FVIII 100 %, Q561N 105 % +/− 45, G563A 6 % +/− 6, N564D 50 % +/− 23, N564A 45 % +/− 21, N564Q 80 % +/− 26, I566M 100 % +/− 41. The one-stage clotting assay gave identical results than the two-stage assay for each mutant. The mutants were then activated by thrombin (1:1) and the occurrence of FVIII activity was monitored. A 20- to 25-fold increase in FVIII activity was measured within 2 minutes following the addition of thrombin for all mutants. The mutants, except for the inactive G563A, were then analyzed for their resistance to APC by three different assays: an APC resistance kit (Coatest, Chromogenix), an in vitro assay that measured APC sensitivity of FVIIIa and an immunoblot assay that visualized the cleavage efficiency of the A2 fragment. These three assays confirmed the APC resistance of the previously published R336I, R562K and R336I+R562K, as compared with wild-type FVIII. They also revealed that the mutants N564D, N564A and I566M behaved similarly to the wild-type FVIII whereas Q561N and N564Q mutants were partially resistant to APC. The APC resistance ratio were the following 2.3 +/− 0.3 for BDD-FVIII, 2.1 +/− 0.4 for N564Q, 1.7 +/− 0,1 for Q561N, 1.6 +/− 0.3 for R562K and 1.3 +/− 0.3 for R336I+R562K. The N564Q and Q561N mutants exhibited a profile intermediate between wild type FVIII and R562K regarding of the loss of FVIIIa activity that was confirmed on the immunoblot profile. In conclusion we have generated new factor VIII molecules that retained their full procoagulant function while possessing a reduced sensitivity to APC cleavage.
44

Markovitz, Rebecca C., John F. Healey, Ernest T. Parker, Shannon L. Meeks, and Pete Lollar. "The diversity of the immune response to the A2 domain of human factor VIII." Blood 121, no. 14 (April 4, 2013): 2785–95. http://dx.doi.org/10.1182/blood-2012-09-456582.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Key Points The Abs to the human fVIII A2 domain in a murine hemophilia A model inhibit fVIIIa and activation of fVIII Epitopes targeted by hemophilia A mouse Abs cover nearly the entire surface of the human fVIII A2 domain
45

Golino, Paolo, Massimo Ragni, Plinio Cirillo, Annalisa Scognamiglio, Amelia Ravera, Chiara Buono, Angela Guarino, et al. "Recombinant human, active site-blocked factor VIIa reduces infarct size and no-reflow phenomenon in rabbits." American Journal of Physiology-Heart and Circulatory Physiology 278, no. 5 (May 1, 2000): H1507—H1516. http://dx.doi.org/10.1152/ajpheart.2000.278.5.h1507.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Oxygen free radicals induce de novo synthesis of tissue factor (TF), the initiator of the extrinsic pathway of coagulation, within the coronary vasculature during postischemic reperfusion. In the present study we wanted to assess whether TF expression might cause myocardial injury during postischemic reperfusion. Anesthetized rabbits underwent 30 min of coronary occlusion followed by 5.5 h of reperfusion. At reperfusion the animals received 1) saline ( n = 8), 2) human recombinant, active site-blocked activated factor VII (FVIIai, 1 mg/kg, n = 8), or 3) human recombinant activated FVII (FVIIa, 1 mg/kg, n = 8). FVIIai binds to TF as native FVII, but with the active site blocked it inhibits TF procoagulant activity. The area at risk of infarction (AR), the infarct size (IS), and the no-reflow area (NR) were determined at the end of the experiment. FVIIai resulted in a significant reduction in IS and NR with respect to control animals (28.1 ± 11.3 and 11.1 ± 6.1% of AR vs. 59.8 ± 12.8 and 24.4 ± 2.7% of AR, respectively, P < 0.01), whereas FVIIa resulted in a significant increase in IS and NR to 80.1 ± 13.1 and 61.9 ± 13.8% of AR, respectively ( P < 0.01). In conclusion, TF-mediated activation of the extrinsic coagulation pathway makes an important contribution to myocardial injury during postischemic reperfusion.
46

Jenkins, P. Vincent, Jan Freas, Kyla M. Schmidt, Qian Zhou, and Philip J. Fay. "Mutations associated with hemophilia A in the 558-565 loop of the factor VIIIa A2 subunit alter the catalytic activity of the factor Xase complex." Blood 100, no. 2 (July 15, 2002): 501–8. http://dx.doi.org/10.1182/blood-2001-12-0361.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract The 558-565 loop region in the A2 subunit of factor (F) VIIIa forms a direct interface with FIXa. We have expressed and purified B-domainless FVIII (FVIIIWT) and B-domainless FVIII containing the hemophilia A–associated mutations Ser558Phe, Val559Ala, Asp560Ala, Gln565Arg, and the activated protein C cleavage site mutant Arg562Ala. Titration of FVIIIa in FXa generation assays showed that the mutant and wild-type proteins had similar functional affinities for FIXa (dissociation constant [Kd] values ∼5 nM-20 nM and ∼100 nM-250 nM in the presence and absence of phospholipid, respectively). The catalytic activities of the factor Xase complex composed of the hemophilia A–associated FVIII species were markedly reduced both in the presence and absence of phospholipid. FVIIIWT and FVIIIArg562Ala showed catalytic rate constant (kcat) values of approximately 60 minute−1 in the presence of phospholipid, whereas the hemophilia A–associated mutants showedkcat values ranging from 3.3 minute−1 to 7.5 minute−1. In the absence of phospholipid, all kcat values were reduced but FVIIIWT and FVIIIArg562Ala retained higher activities as compared with the hemophilic mutant FVIII forms. Fluorescence anisotropy experiments using fluorescein-modified FIXa confirmed that all FVIII forms interacted with FIXa. However, the presence of factor X yielded minimal increases in anisotropy observed with the mutant factor VIII forms, consistent with their reduced activity. These results show that residues within the 558-565 loop are critical in modulating FIXa enzymatic activity but do not contribute significantly to the affinity of FVIIIa for FIXa.
47

Wakabayashi, Hironao, Jennifer Wintermute, and Philip J. Fay. "Stabilizing Factor VIIIa By Combining Mutations That Modulate Inter-Subunit Interactions and Proteolytic Inactivation." Blood 122, no. 21 (November 15, 2013): 3572. http://dx.doi.org/10.1182/blood.v122.21.3572.3572.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Factor (F) VIIIa serves as a cofactor for FIXa forming intrinsic FXase complex. The activity of FVIIIa is labile due to the tendency for A2 subunit of the cofactor to dissociate, thereby inactivating FVIIIa and dampening FXase activity. Recently we have shown that a combination of hydrophobic point mutations generated FVIII(a) variants with improved FVIII thermal stability and enhanced A2 subunit retention in FVIIIa (Wakabayashi et al. J. Thromb. Haemost. 2012, 10: 492-495). Furthermore, FVIIIa is a target for proteolytic inactivation catalyzed by FXa, the product of FXase complex. We have shown this proteolysis further contributes to the down-regulation of FXase and that altering the sequence flanking the primary FXa cleavage site in FVIIIa (at Arg336) yields marked reductions in the rates of proteolytic inactivation of FVIIIa catalyzed by FXa (DeAngelis et al. J. Biol. Chem. 2012, 287: 15409-15417). In this study we prepared various combinations of the above mutations to obtain a panel of novel FVIII molecules and examined the attributes of these reagents in assays monitoring FVIII thermal stability, FVIIIa spontaneous decay rates, rates of proteolytic inactivation by FXa, and thrombin generation potential. The mutants we prepared include 336(P4-P3’)562 [where residues 333-339 (PQLRMKN) that flank the fast FXa-cleavage site at Arg336 are replaced with residues flanking the slow cleavage site at Arg562 (residues 559-565, VDQRGNQ)] are combined with D519V/E665V (336-V) or D519/VE665V/A108I (336-VI). Specific activity values for the combined mutation variants ranged from 85% to 170% the WT value using a one-stage clotting assay and 95% to 110% using a two-stage chromogenic assay. FVIII thermal stability was tested by monitoring FVIII activity remaining during incubation at 57ºC over a 20 min time course. Thermal decay rates for 336-V and 336-VI variants were reduced by ∼1.7 and ∼5-fold as compared with the WT FVIII value. These rate values reflected additive effects of the individual mutations since rate values for controls FVIII D519V/E665V and FVIII A108I were reduced ∼1.7, and ∼3-fold, respectively, relative to WT, while the thermal decay rate for FVIII 336(P4-P3’)562 was WT-like. FVIIIa spontaneous decay rates were determined following activation of FVIII by thrombin and these values were reduced by ∼25-fold for the 336-V and 336-VI variants as compared with the WT FVIIIa value. Interestingly, this magnitude of rate reduction suggested a synergistic effect since rate values were reduced ∼14-fold for the D519V/E665V control and were essentially unaffected in the 336(P4-P3’)562 and A108I controls. FVIIIa inactivation by FXa was monitored by a one-stage clotting assay after FVIIIa was incubated with 5 nM FXa at 37ºC for a 30 min time course. FVIII 336-V and 336-VI variants showed similar resistance to inactivation by FXa (∼10-fold reduced rate compared with WT FVIIIa) as the 336(P4-P3’)562 control. FVIII D519V/E665V and A108I variants showed slightly reduced inactivation rates (∼1.1 and 1.6 fold) as compared with WT FVIII. Thrombin generation assays were performed using FVIII deficient plasma. Assays were run using 0.25 nM FVIII and 4 µM phospholipid vesicles, initiated with 0.25 pM tissue factor, and the amount of generated thrombin was calculated overtime by monitoring the development of fluorescent thrombin substrate peptide. FVIII 336-V and 336-VI variants generated comparable amounts of thrombin as FVIII D519V/E665V, showing ∼70% increases in endogenous thrombin potential (ETP), while the A108I and 336(P4-P3’)562 controls showed WT-like ETP values. Taken together, these results indicate that it is possible to combine the above gain-of-function FVIII mutations to yield FVIII variants such as the 336-VI form in order to generate a more stable procofactor as judged by improved thermal stability (∼5-fold relative to WT), enhanced retention of A2 subunit increasing FVIIIa stability (∼25 fold) and increased resistance to proteolytic inactivation (∼10 fold). The latter two attributes would potentially prolong FXase activity during clotting and this effect is suggested by the improved thrombin generation parameters for this variant. Disclosures: No relevant conflicts of interest to declare.
48

Gangadharan, Bagirath, H. Trent Spencer, Ernest T. Parker, and Christopher B. Doering. "High-Level Expression of Porcine Factor VIII from Murine Mesenchymal Stem Cells." Blood 104, no. 11 (November 16, 2004): 5281. http://dx.doi.org/10.1182/blood.v104.11.5281.5281.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
Abstract Low-level expression of human factor VIII (fVIII) has limited the success of clinical trials for hemophilia A using both ex vivo and in vivo gene transfer methods. Ex vivo genetic modification provides increased control of gene transfer and permits thorough characterization of transgene copy number, chromosomal integration site(s), and expression levels prior to transplantation. A subpopulation of adherent bone marrow-derived cell types, termed mesenchymal stem cells (MSCs), comprise an attractive target cell type for ex vivo gene transfer due to their accessibility, ability to differentiate into multiple cell types, and long-term survival following transplantation. Recently we demonstrated, in vitro, that recombinant B-domain-deleted porcine fVIII (rp-fVIII) is expressed at levels 10 – 14-fold greater than recombinant B-domain-deleted human fVIII (rh-fVIII) due to an enhanced rate of secretion from baby hamster kidney-derived (BHK-M) cells. Additionally we found that only the A1 and activation peptide-A3 domain sequences of porcine fVIII are necessary to retain high-level expression of hybrid human/porcine fVIII constructs. Here we report the expression of rp-fVIII from murine MSCs isolated from exon-16 fVIII knockout mice following ex vivo retroviral transduction. MSCs were transduced with ecotropic envelope-pseudotyped murine stem cell virus containing a rp-fVIIII transgene at an multiplicity of infection of 2 – 5 functional viral particles/target MSC. Following this transduction regimen the cell population harbored an average of 1.8 proviral genomes/cell as determined by quantitative real-time PCR. During culture in growth medium supplemented with 20% fetal bovine serum (FBS) rp-fVIII-transduced MSCs demonstrated a steady-state level of 2,400 fVIII mRNA transcripts/cell and an apparent fVIII production rate of 174 units/106 cells/24 hr as determined by quantitative real-time RT-PCR and one-stage clotting assay, respectively. However, when cultured in serum-free medium the mRNA levels decreased to 950 transcripts/cell and apparent fVIII production was reduced to 14 units/106 cells/24 hr. The latter mRNA/fVIII expression ratio is in agreement with that previously reported from stably transduced BHK-M clonal cell lines. In contrast the former mRNA levels are lower than predicted from the observed fVIII activity levels. The activation quotient (ratio of apparent fVIII activity following thrombin pre-treatment to non-thrombin treated baseline fVIII activity) increased 17-fold when the cells were cultured in serum-free versus serum-containing medium. Additionally, SDS-PAGE analysis of immunoprecipitated fVIII protein from serum-containing and serum-free medium revealed the presence of activated fVIII (fVIIIa) in conditioned medium samples containing FBS. Highly purified rp-fVIII from transduced-MSCs cultured in serum-free medium displayed similar relative mobility to BHK-M produced rp-fVIII upon SDS-PAGE analysis. These data warn against the determination of fVIII activity in serum-containing medium using the one-stage coagulation assay due to the presence of activated fVIII with high specific activity. Based on these results it is reasonable to predict that hemophilia A could be cured using high-level expression fVIII constructs by transplantation of a feasible number (106 - 108) of cells containing a single copy of rp-fVIII or other high-level expression hybrid human/porcine fVIII transgenes.
49

Cooper, Adrian, Zhong Liang, Francis Castellino, and Elliot Rosen. "Cloning and Characterization of the Murine Coagulation Factor X Gene." Thrombosis and Haemostasis 83, no. 05 (2000): 732–35. http://dx.doi.org/10.1055/s-0037-1613901.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryThe gene encoding murine coagulation factor X (fX) was isolated and characterized from a λFIX II library generated from murine genomic DNA. The 20130 bp sequence contains 18049 nucleotides that extend from the initiating methionine to the polyadenylation site. 1056 nucleotides 5’ of the start codon were determined and contain putative start sites for the FX mRNA as well as sites for binding of putative transcription factors. The sequence extends 1024 3’ of the polyadenylattion site.The gene contains 8 exons and 7 introns which were determined by comparing the mouse FX cDNA and gene sequences. The exonic structure of the gene is similar to that of the other mammalian vitamin K-dependent serine proteases of the coagulation system. These include an exon encoding the prepropepetide, the gladomain, a short helical stack, two exons for the two EGF domains, the activation pepetide, and two exons encoding the serine protease domain. The 5’ sequence of the mouse FX gene overlaps with the 3’ region of the FVII gene indicating that the murine FVII and FX gene are arranged in a head to tail arrangement as they are in humans. Abbreviations: fVII, coagulation factor VII; fIX, coagulation factor IX; fX, coagulation factor X; PC, Protein C; fV, coagulation factor V; fVa, activated coagulation factor V; fVIII, coagulation factor VIII; fVIIIa, activated coagulation factor VIII.
50

Ogiwara, Kenichi, Tomoko Matsumoto, Katsumi Nishiya, Masahiro Takeyama, Midori Shima, and Keiji Nogami. "Mechanisms of human neutrophil elastase-catalysed inactivation of factor VIII(a)." Thrombosis and Haemostasis 105, no. 06 (2011): 968–80. http://dx.doi.org/10.1160/th10-12-0777.

Full text
APA, Harvard, Vancouver, ISO, and other styles
Abstract:
SummaryMechanisms of inflammation and coagulation are linked through various pathways. Human neutrophil elastase (HNE), can bind to activated platelets, might be localised on platelet membranes that provide negatively-charged phospholipid essential for the optimum function of tenase complex. In this study, we examined the effect of HNE on factor (F)VIII. FVIII activity was rapidly diminished in the presence of HNE and was undetectable within 10 minutes. The inactivation rate waŝ8-fold greater than that of activated protein C (APC). This time-dependent inactivation was moderately affected by von Willebrand factor. HNE proteolysed the heavy chain (HCh) of FVIII into two terminal products, A11–358 and A2375–708, by limited proteolysis at Val358, Val374, and Val708. Cleavage at Val708 was much slower than that at Val358 in the >90-kDa A1-A2-B compared to the 90-kDa A1-A2. The 80-kDa light chain (LCh) was proteolysed to 75-kDa product by cleavage at Val1670. HNE-cata- lysed FVIIIa inactivation was markedly slower than that of native FVIII (by ~25-fold), due to delayed cleavage at Val708 in FVIIIa. The inactivation rate mediated by HNE was ~8-fold lower than that by APC. Cleavages at Val358 and Val708 were regulated by the presence of LCh and HCh, respectively. In conclusion, HNE-catalysed FVIII inactivation was associated with the limited-proteolysis that led to A11–358, A2375–708, and A3-C1-C21671–2332, and subsequently to critical cleavage at Val708. HNE-related FVIII(a) reaction might play a role in inactivation of HNE-induced coagulation process, and appeared to depend on the amounts of inactivated FVIII and active FVIIIa which is predominantly resistant to HNE inactivation.Note: An account of this work was presented at the 51st annual meeting of the American Society of Hematology, December 10, 2009, New Orleans, LA, USA.

To the bibliography