A normal hemostatic response to vascular injury requires both factor VIII (FVIII) and von Willebrand factor (VWF). In plasma, VWF and FVIII normally circulate as a noncovalent complex, and each has a critical function in the maintenance of hemostasis. Furthermore, the interaction between VWF and FVIII plays a crucial role in FVIII function, immunogenicity, and clearance, with VWF essentially serving as a chaperone for FVIII. Several novel recombinant FVIII (rFVIII) therapies for hemophilia A have been in clinical development, which aim to increase the half-life of FVIII (∼12 hours) and reduce dosing frequency by utilizing bioengineering techniques including PEGylation, Fc fusion, and single-chain design. However, these approaches have achieved only moderate increases in half-life of 1.5- to 2-fold compared with marketed FVIII products. Clearance of PEGylated rFVIII, rFVIIIFc, and rVIII-SingleChain is still regulated to a large extent by interaction with VWF. Therefore, the half-life of VWF (∼15 hours) appears to be the limiting factor that has confounded attempts to extend the half-life of rFVIII. A greater understanding of the interaction between FVIII and VWF is required to drive novel bioengineering strategies for products that either prolong the survival of VWF or limit VWF-mediated clearance of FVIII.

Factor VIII (FVIII) is a glycoprotein synthesized by sinusoidal and vascular endothelial cells in the liver and lung1,2  that is critical for hemostasis.3  Normal hemostatic responses to vascular injury require both FVIII and its physiologic partner, von Willebrand factor (VWF).4  In plasma, VWF and FVIII circulate as a noncovalent complex that regulates platelet aggregation and clot formation. Each protein is a separate gene product but the processes they regulate are coordinated and critical to maintenance of hemostasis. VWF is required for platelet adhesion to subendothelium, and for normal FVIII survival in circulation.4  Deficiencies or structural defects in FVIII or VWF are responsible for hemophilia A (HA) and von Willebrand disease (VWD), respectively.5,6  Herein, we review the FVIII-VWF interaction and its implications for the treatment of HA, including FVIII half-life, clearance, and immunogenicity. A better understanding of this interaction is critical to facilitate development of improved therapies.

Although it is now well recognized that VWF and FVIII are 2 separate gene products, the presence of a VWF-FVIII protein complex led to early misunderstandings about their relationship. Hemophilia was recognized in the Talmud >1700 years ago7  and was first effectively treated using whole-blood transfusion in the 1840s.8  VWD was first described by Erik von Willebrand and the disorder was termed pseudohaemophilia.9  In the 1960s, antihemophilic factor, which is synonymous with FVIII, was recognized as the missing clotting factor in both HA and severe VWD; it was also recognized that another vascular component was missing in VWD that was present in patients with HA.10  In the 1970s, sometimes heated debates ensued as to whether the clotting factor (FVIII) and the vascular factor (VWF) were on 111  or 212  molecules.

Once VWF was recognized as a distinct protein, a flurry of studies demonstrated that it was synthesized in 2 cell types: megakaryocytes and endothelial cells, processed and then stored in platelet α-granules and endothelial Weibel-Palade bodies, respectively.13-15  The intracellular synthesis and trafficking of VWF is complex and has been best characterized in endothelial cells. Following synthesis of pro-VWF (propeptide and VWF monomer), C-terminal dimerization takes place cotranslationally in the endoplasmic reticulum before transport to the Golgi where the dimers undergo further carbohydrate modification, multimerization, VWF-propeptide cleavage, and storage. The endothelium stores high-molecular-weight VWF multimers16  that can be released by 1-deamino-8-D-arginine vasopressin (DDAVP; desmopressin), which binds to the vasopressin V2 receptor and induces release of VWF multimers into plasma.17  DDAVP releases VWF from endothelial cells but not from platelets.

Synthesis and storage of VWF in megakaryocytes/platelets is more complex. Although α-granules contain VWF, other stored proteins are much more heterogeneous.18,19  Platelet α-granule contents are released during platelet aggregation in response to agonists such as arachidonic acid, adenosine 5′-diphosphate, collagen, and epinephrine. VWF is critical for both recruitment and activation of platelets. At sites of vascular injury, plasma VWF binds to exposed collagen via 2 collagen-binding sites. The A1 domain binds to collagen IV and VI, and the A3 domain binds to collagen I and III.20  The immobilized VWF then undergoes shear-induced rearrangement21  to spontaneously recruit platelets by adhesion through the glycoprotein Ib-IX receptor22  expressed on platelets. The recruited platelets then undergo activation through fibrinogen- and VWF-dependent interactions with αIIbβ3, which induces platelet aggregation.23,24 

The biosynthesis of FVIII is complex, and some questions about its sites of synthesis remain unanswered (Figure 1). Early transplantation studies in dogs (and subsequently in humans) demonstrated that transplantation of a normal liver to a HA patient can normalize plasma FVIII levels.25,26  Although these studies suggested that liver was a major site of FVIII synthesis, other transplantation studies suggested that FVIII was synthesized by spleen and lymphatic tissue.27,28  Although liver transplantation has resulted in normalized plasma FVIII levels and a lifelong cure for many patients, studies of transplant recipients suggest that the DDAVP-releasable pool of FVIII is not restored by liver transplant.29  Furthermore, significant extrahepatic FVIII synthesis was demonstrated in a normal subject who received a liver from a HA patient and who did not then experience reduced plasma FVIII levels.30  It has also been demonstrated that FVIII is synthesized by sinusoidal and microvascular endothelial cells,1,2  and in situ studies identified FVIII messenger RNA in endothelial cells.31  Finally, 2 recent studies using murine conditional knockout approaches indicated endothelial cells are the major, if not exclusive, source of FVIII,32,33  and gene transfer of FVIII to endothelial cells in mice normalized plasma FVIII and restored the releasable FVIII pool.34  Recent studies have also documented the presence of FVIII on the VWF strings that are released from the endothelium.35  However, the documentation of FVIII expression in lymphatic endothelium and certain types of fenestrated endothelium (in the liver sinusoids and renal glomerulus), with little or no coexpression of VWF,36  suggests that there is still more to learn about the precise endothelial sources of production of these 2 proteins. Nevertheless, it seems clear that different endothelial beds show differential expression of VWF and FVIII.

Figure 1

Details of the sites of synthesis and clearance of VWF and FVIII. Although the synthesis of VWF has long been known to be the vascular endothelium and megakaryocytes, the location of FVIII expression has only recently been confirmed in some types of endothelial cell: in fenestrated forms of endothelium (liver sinusoidal endothelium and glomerular endothelium), in lymphatic endothelium, and in some high endothelial venules. In most forms of endothelium, VWF and FVIII are not coexpressed. Clearance of VWF and FVIII occurs most frequently as a complex, in the sinusoids of the liver and spleen where a range of lectin and scavenger receptors expressed on macrophages, sinusoidal endothelium, and hepatocytes bind to and internalize the 2 proteins. Protein clearance is influenced by factors such as shear, desialylation, and protein sequence variants. SCARA5, scavenger receptor class A, member 5.

Figure 1

Details of the sites of synthesis and clearance of VWF and FVIII. Although the synthesis of VWF has long been known to be the vascular endothelium and megakaryocytes, the location of FVIII expression has only recently been confirmed in some types of endothelial cell: in fenestrated forms of endothelium (liver sinusoidal endothelium and glomerular endothelium), in lymphatic endothelium, and in some high endothelial venules. In most forms of endothelium, VWF and FVIII are not coexpressed. Clearance of VWF and FVIII occurs most frequently as a complex, in the sinusoids of the liver and spleen where a range of lectin and scavenger receptors expressed on macrophages, sinusoidal endothelium, and hepatocytes bind to and internalize the 2 proteins. Protein clearance is influenced by factors such as shear, desialylation, and protein sequence variants. SCARA5, scavenger receptor class A, member 5.

Close modal

It is reasonable to ask where these 2 proteins first meet to form the noncovalent complex that circulates in plasma (Figure 1). In humans, the DDAVP-releasable FVIII pool is dependent on the in vivo synthesis of both VWF and FVIII.37  If FVIII expression is induced in a cell that synthesizes and stores VWF, FVIII will follow VWF and be stored and released as a noncovalent complex similar to that found in normal plasma.38  This is probably what happens in some but not necessarily all endothelial cells. Type 2N VWD is associated with mutations in its D′D3 domain that interfere with FVIII binding.39,40  When not bound to VWF, FVIII is cleared rapidly, as in type 3 VWD. However, DDAVP induces release of both VWF and FVIII, but following release the FVIII half-life is 2 hours whereas VWF has a normal half-life of 12 hours in type 2N VWD.41  This suggests that intracellularly, in the acidic environment of the late Golgi, there is normal association and storage of the VWF-FVIII noncovalent complex that is disrupted at the neutral pH of plasma following release.42 

The fact that FVIII and VWF are colocalized in storage granules within endothelial cells has implications for the selection of target tissues for FVIII gene therapy studies. Despite the investigation of various strategies to deliver the complementary DNAs (cDNAs) for either FVIII or FIX in human gene therapy trials, success has only been achieved using a liver-targeted recombinant adeno-associated virus vector to deliver the FIX cDNA to individuals with hemophilia B.43  Initial attempts at gene therapy for HA included in vivo delivery of FVIII using retrovirus or “gutless” adenovirus vectors and omental implantation of ex vivo transduced autologous fibroblasts. However, these strategies have not led to sustained FVIII expression or therapeutic FVIII plasma levels. These attempts have been hampered by the large size of the FVIII cDNA and relatively inefficient expression in target tissues. To overcome these challenges, current strategies involve bioengineering the FVIII cDNA to remove the B domain, adding targeted consensus sequences for asparagine-linked glycosylation, and optimizing the transgene codons.44  Investigators have also been exploring alternative target tissues, in an effort to mimic a more physiologic source for FVIII expression in vivo. In 1 study, human umbilical vein endothelial cells were transduced with a retroviral FVIII (B-domain deleted) construct.45  The investigators demonstrated that FVIII colocalized with VWF in Weibel-Palade bodies, and the human umbilical vein endothelial cells, upon agonist-induced stimulation, displayed a parallel release of FVIII and VWF, suggesting that this approach reconstituted the normal physiologic storage pool of FVIII-VWF complexes. Blood outgrowth endothelial cells have also been transduced ex vivo with viral vectors, and the transduced blood outgrowth endothelial cells can be delivered systemically, resulting in their transplantation into several organs.46  This strategy also achieved FVIII/VWF colocalization. In addition, it has been demonstrated that targeting FVIII expression to platelets results in FVIII storage together with VWF in platelet α-granules and corrects the murine hemophilia A phenotype even in the presence of high-titer anti-FVIII inhibitory antibodies (inhibitors).47 

FVIII spends virtually its entire life cycle interacting with other proteins and membrane surfaces. Recent studies have expanded our understanding of the structures and interactions of FVIII and VWF, yielding clinical insights. For example, mutagenesis studies and patient-derived data are defining the roles of specific amino acids in the activities of both proteins. The structures and affinities of FVIII binding to its various partners, though informative, must be interpreted with caution, as interactions between isolated components of, for example, the intrinsic tenase complex may differ somewhat from those that occur physiologically, at wound sites and under shear forces (Figure 2). Therefore, the growing body of FVIII structural studies (crystallographic,48-50  fluorescence resonance energy transfer,51  cryoelectron microscopic,52  etc) are best considered a series of “snapshots” from the ensemble of conformations accessible to this multidomain protein.

Figure 2

Dynamic equilibrium between VWF and FVIII and details of the VWF-FVIII association under normal conditions of synthesis, secretion, and clearance. Whereas the vast majority of VWF circulates as an FVIII-free protein in the circulation, the opposite is true for FVIII with 95% to 98% being in complex with VWF.128  Although of relatively high affinity (KD, 0.2 nM), complex formation is characterized by a temperature-sensitive highly dynamic equilibrium, with rapid association and dissociation rate constants (2-4 × 106 M−1 s−1 and 0.3-6 × 10−3 s−1), respectively.129  The influence of shear on the VWF-FVIII association and configuration is unresolved but may play a role in modulating clearance and immunogenicity.

Figure 2

Dynamic equilibrium between VWF and FVIII and details of the VWF-FVIII association under normal conditions of synthesis, secretion, and clearance. Whereas the vast majority of VWF circulates as an FVIII-free protein in the circulation, the opposite is true for FVIII with 95% to 98% being in complex with VWF.128  Although of relatively high affinity (KD, 0.2 nM), complex formation is characterized by a temperature-sensitive highly dynamic equilibrium, with rapid association and dissociation rate constants (2-4 × 106 M−1 s−1 and 0.3-6 × 10−3 s−1), respectively.129  The influence of shear on the VWF-FVIII association and configuration is unresolved but may play a role in modulating clearance and immunogenicity.

Close modal

Binding of FVIII to VWF is mediated by noncovalent interactions between the FVIII-C1 and FVIII-C2 domains, an acidic peptide (FVIII-ap) at its light-chain N terminus, and the VWF D′D3 region (VWF-D′D3).53  VWF-D′D3 comprises the N-terminal region of monomeric VWF created during proteolysis and assembly to create multimeric VWF. Studies investigating the crystal structure of human FVIII-C2,49  homology-modeled FVIII-C1,54  and mutations in both domains have suggested that certain amino acid substitutions affect residues that directly mediate FVIII-VWF interaction, whereas others result in protein misfolding.54  Medium-resolution crystal structures show no electron density for FVIII-ap, indicating that it is flexible.48,50  A VWF-D′ solution structure55  (Figure 3), as well as modeling based on single-particle electron microscopy of complexes formed by FVIII and VWF-D′D3 domains,56,57  have allowed further analysis of the FVIII-VWF interface and of VWF-D′ mutations associated with reduced FVIII-VWF affinity and type 2N VWD. These studies have shown that the positively charged VWF-D′ surface is probably a binding site for FVIII-ap.56,57  More detailed models are on the horizon, as the 3-dimensional structures suggest additional mutagenesis studies to identify specific residues that mediate FVIII-VWF binding.

Figure 3

VWF-D′ solution structure. FVIII and VWF-D′ domain structures. Residues where substitutions affected or would be expected to affect binding affinity are shown in spherical representation (blue, positively charged; red, negatively charged; brown, hydrophobic; turquoise, neutral). (A) Hydrophobic FVIII-C2 residues M2199, F2200, L2251, and L2252 interact with VWF, whereas flanking surface-exposed residues R2215, R2220, and K2249 make this FVIII-C2 region positively charged. Five noncysteine hemophilic FVIII-C1 domain amino acid substitutions that affected FVIII-VWF binding (http://www.factorviii-db.org/) are indicated. The cleavage site for the FVIII light-chain (LC) acidic peptide is also indicated. VWF-D′ is oriented with its flexible TIL′ domain approaching FVIII. (B) VWF-D′ is oriented here with its rigid E domain at the top. Noncysteine amino acids whose substitutions are associated with type 2N VWD are shown in spherical representation, and those affecting charged side chains are labeled. Disulfide bonds are shown in orange stick representation.

Figure 3

VWF-D′ solution structure. FVIII and VWF-D′ domain structures. Residues where substitutions affected or would be expected to affect binding affinity are shown in spherical representation (blue, positively charged; red, negatively charged; brown, hydrophobic; turquoise, neutral). (A) Hydrophobic FVIII-C2 residues M2199, F2200, L2251, and L2252 interact with VWF, whereas flanking surface-exposed residues R2215, R2220, and K2249 make this FVIII-C2 region positively charged. Five noncysteine hemophilic FVIII-C1 domain amino acid substitutions that affected FVIII-VWF binding (http://www.factorviii-db.org/) are indicated. The cleavage site for the FVIII light-chain (LC) acidic peptide is also indicated. VWF-D′ is oriented with its flexible TIL′ domain approaching FVIII. (B) VWF-D′ is oriented here with its rigid E domain at the top. Noncysteine amino acids whose substitutions are associated with type 2N VWD are shown in spherical representation, and those affecting charged side chains are labeled. Disulfide bonds are shown in orange stick representation.

Close modal

FVIII is sulfated at 6 tyrosine residues, and these are key modulators of its extracellular protein-protein interactions. Sulfation of FVIII-Tyr1680 is required for efficient VWF binding, increasing the FVIII-VWF affinity fivefold.58  Full sulfation at Tyr1680 appears to be dependent on the cell line and expression conditions. Nielsen and colleagues examined extracted ion chromatograms showing sulfated and nonsulfated (minor or absent peak) Tyr1680-containing peptides.59  Plasma-derived FVIII and recombinant FVIII (rFVIII) from human HEK293 cells demonstrated only a sulfated peak,60  whereas rFVIII expressed by rodent cells exhibited both a sulfated peak and nonsulfated peaks. Likewise, a B-domain–deleted rFVIII expressed in Chinese hamster ovary (CHO) cells showed full Tyr1680 sulfation, suggesting that culture conditions may influence the efficiency of this modification. These studies highlight that choices made in cell production and design of FVIII constructs can have important implications with regard to optimizing the FVIII-VWF interaction.

Development of an immune response against FVIII complicates the treatment of ∼30% of patients with HA.61  Neutralizing antibodies (“inhibitors”) are a more significant problem in HA than in VWD because approximately half of severe HA cases are caused by mutations that result in no circulating FVIII protein, whereas type 3 VWD (complete lack of VWF) is extremely rare. Patients with a circulating dysfunctional FVIII or VWF (eg, with a missense substitution) are far more likely to have functional immune tolerance to therapeutic FVIII. Patients who develop inhibitors typically have antibodies that bind to the FVIII-A2 and FVIII-C2 domains,62  and there has been substantial progress in characterizing FVIII B-cell epitopes.63-67 

Risk factors for development of inhibitors in HA include genetic factors such as the type of F8 gene mutation, the severity of hemophilia, HLA haplotype, and polymorphisms in genes involved in immunoregulation68  as well as environmental factors such as the intensity of FVIII exposure and concomitant immunologic “danger signals.”69  Identification of risk factors related to rFVIII vs plasma-derived FVIII has been more controversial. Product-specific differences that have been examined include amino acid sequence alterations, including B-domain deletion/truncation, glycosylation patterns, and a potential protective role of VWF.70,71 

Glycosylation patterns for rFVIII depend on the cell line from which the protein is produced, and they tend to differ from those of plasma-derived products. These differences may contribute to immunogenicity, although other factors cannot be ruled out, including an immunomodulatory role for VWF.72  A rFVIII has been produced in a human cell line with the aim of retaining human-specific posttranslational modifications, including its glycosylation profile.60  To date, clinical studies with this human cell line-derived rFVIII, involving 59 and 32 previously treated children and adult hemophiliacs, respectively, have shown no FVIII inhibitors.73,74 

As the chaperone of FVIII in plasma, the role of VWF in FVIII immunogenicity cannot be ignored. It has been proposed that VWF may serve an immunoprotective role in 2 ways. First, VWF reduces FVIII endocytosis by antigen-presenting cells and a recent study has shown that VWF also modulates FVIII peptide presentation by dendritic cells.75-77  Second, by increasing the FVIII half-life in circulation, VWF may increase contact time and processing of FVIII by B cells in the splenic marginal zone, promoting immunoregulatory mechanisms, and resulting in tolerance. Immunodominant B-cell epitopes have been described in the FVIII A2 and C2 domains.78  Binding of FVIII to VWF is postulated to shield 1 or more C2 domain epitopes and decrease uptake of FVIII by antigen-presenting cells. However, a study by Meeks and colleagues found that neutralizing anti-C2 antibodies still formed when FVIII was unable to dissociate from VWF.79  The same study also found evidence that in the absence of VWF, FVIII may be cleared by alternative receptors through which it is degraded in a process that does not include antigen presentation. Thus, it appears that FVIII immunogenicity may be both positively and negatively affected by its association with VWF. FVIII protein content in FVIII products may also affect immunogenicity. One study showed that rFVIII preparations contained significantly more FVIII protein per international unit (IU) than plasma-derived FVIII concentrates, and that only ∼80% of the rFVIII could bind to VWF.80  This unbound rFVIII may have been denatured or damaged, which could cause increased immunogenicity.

Preclinical and clinical data further support the notion that the presence of VWF in FVIII preparations is associated with reduced FVIII immunogenicity.72,81,82  In a systematic review of previously untreated patients (PUPs) with HA, the risk of inhibitor development was greater with rFVIII than with VWF-containing plasma-derived FVIII (27.4% [95% confidence interval (CI), 23.6-31.5] vs 14.3% [95% CI, 10.4-19.4], respectively).83  In addition, using a combination of retrospective and prospective data, the relative risk for inhibitor development in PUPs treated with rFVIII was higher than for PUPs who received plasma-derived FVIII (relative risk = 2.4; 95% CI, 1.0-5.8; P = .049).84  A note of caution in the interpretation of PUP studies is that, in many instances, some of the PUPs have received a small number of blood-product infusions prior to study inclusion (minimally treated patients). This factor may confound interpretation of the origin of the FVIII immunogenic stimulus.

It should be noted that other studies have reported conflicting results. A large retrospective analysis of data on 316 PUPs enrolled in the Concerted Action on Neutralizing Antibodies in severe haemophilia A (CANAL) cohort study found that plasma-derived FVIII with considerable quantities of VWF carried the same inhibitor risk as rFVIII products.85  A meta-analysis of 28 prospective studies investigating inhibitor development in 1421 PUPs with severe HA reached similar conclusions, finding no statistically significant difference in inhibitor development between plasma-derived FVIII and rFVIII or between different classes of rFVIII.71  Most recently, the Study on Inhibitors in Plasma-Product Exposed Toddlers (SIPPET) trial, the first prospective randomized study to address this issue, has reported a 1.87-fold increase in inhibitors in PUPs treated with rFVIII compared with plasma-derived FVIII.86  The study includes 251 previously untreated HA boys (<6 years of age) randomly assigned to receive a VWF-containing plasma-derived FVIII product or a rFVIII product and followed for 50 exposure days for inhibitor development. This study presents the first clear evidence of a differential immunogenicity risk associated with rFVIII products. Although several issues have been raised by this study, including whether the results from a geographically unusual study population (73% of subjects derive from India, Egypt, and Iran) can be extrapolated more widely, the limit of follow-up to 50 exposure days, and the choice of a lower-than-usual inhibitor titer threshold (0.4 Bethesda units), the study results have nevertheless caused a major debate about the optimal FVIII product to begin therapy in PUPs. Importantly, the mechanistic basis of a differential immunogenic tendency for rFVIII products remains without explanation.

Inhibitor development requires activation of CD4+ T-effector cells, and several immunodominant T-cell epitopes have been identified in FVIII.87-89  Characterization of FVIII-specific T-cell clones and lines, and their use in studies exploring cell-based immunotherapies to promote immune tolerance to FVIII, are suggesting novel approaches to reduce the incidence of inhibitors.90,91  Animal models remain essential for mechanistic studies,92  and preclinical examination has demonstrated potential immune-regulatory properties of Fc fusion proteins93  and the possible mitigation of FVIII immunogenicity with the long-term persistence of circulating FVIII levels achieved by gene transfer either alone94  or through the generation of T-regulatory cells.95 

Because FVIII-VWF circulates as a protein complex, assessment of how the individual constituents are removed from circulation is complex. Human and animal studies have demonstrated that FVIII half-life is reduced about sixfold in the absence of VWF, whereas the VWF half-life is unaffected by the presence of FVIII. This is compatible with the view that most FVIII molecules are eliminated while complexed to VWF, and further suggests that VWF protects FVIII from premature clearance. To understand FVIII clearance it is thus essential to understand VWF catabolism.

Interestingly, VWF clearance varies significantly between individuals. This is not only true when assessing endogenous VWF survival following desmopressin treatment (range, 4-26 hours), but also after infusing exogenous VWF concentrates (range, 4-157 hours).96  This variability likely originates from both VWF and non-VWF-related factors. Three categories of potential parameters that could modulate VWF clearance are summarized in Table 1: patient-related, receptor-related, and VWF-related. To illustrate how these parameters influence VWF catabolism, the interaction between VWF and some of its potential clearance receptors (Table 2) will be discussed.

Table 1

Parameters that potentially modulate clearance of VWF

ParametersAffects VWF interaction with
Patient-related parameters  
 • Body weight • ASGPR 
 • Age • ASGPR 
 • Diet • ASGPR 
 • ABO antigen status • ? 
 • Vascular integrity • LRP1 
 • Inflammatory state • ASGPR 
Receptor-related parameters  
 • Repertoire, expression level, activity • ? 
 • Sequence variations • CLEC4M, LRP1 
VWF-related parameters  
 • Sequence variations (polymorphisms, mutations) • LRP1 
 • Glycosylation variations • ASGPR, Siglec-5 
 • Proteins associated with VWF (FVIII, osteoprotegerin, galectins, etc) • ? 
 • Activation state, unfolding • LRP1 
ParametersAffects VWF interaction with
Patient-related parameters  
 • Body weight • ASGPR 
 • Age • ASGPR 
 • Diet • ASGPR 
 • ABO antigen status • ? 
 • Vascular integrity • LRP1 
 • Inflammatory state • ASGPR 
Receptor-related parameters  
 • Repertoire, expression level, activity • ? 
 • Sequence variations • CLEC4M, LRP1 
VWF-related parameters  
 • Sequence variations (polymorphisms, mutations) • LRP1 
 • Glycosylation variations • ASGPR, Siglec-5 
 • Proteins associated with VWF (FVIII, osteoprotegerin, galectins, etc) • ? 
 • Activation state, unfolding • LRP1 
Table 2

Potential clearance receptors for FVIII and VWF

Receptor*FVIIIClearance?VWFClearance?
LRP1 Yes Yes Yes Yes (requires VWF unfolding) 
LDL-R Yes Yes Unknown Unknown 
ASGPR Yes Yes (hyposialylated FVIII) Yes Yes (hyposialylated VWF) 
CD206 Yes Probably not (antigen presentation) No No 
Siglec-5 Yes Yes Yes Yes 
CLEC4M Unknown Unknown Yes Yes 
STAB2 Yes Yes Yes Yes 
SCARA5 Unknown Unknown Yes Unknown 
HSPGs Yes Possibly Yes Unlikely 
Receptor*FVIIIClearance?VWFClearance?
LRP1 Yes Yes Yes Yes (requires VWF unfolding) 
LDL-R Yes Yes Unknown Unknown 
ASGPR Yes Yes (hyposialylated FVIII) Yes Yes (hyposialylated VWF) 
CD206 Yes Probably not (antigen presentation) No No 
Siglec-5 Yes Yes Yes Yes 
CLEC4M Unknown Unknown Yes Yes 
STAB2 Yes Yes Yes Yes 
SCARA5 Unknown Unknown Yes Unknown 
HSPGs Yes Possibly Yes Unlikely 

HSPG, heparin sulfate proteoglycan; SCARA5, scavenger receptor class A, member 5; STAB2, stabilin 2.

*

The 9 receptors listed have been demonstrated at some level to bind to and, in some cases, mediate clearance of FVIII and/or VWF.

Responses indicate evidence of receptor binding to the protein.

Responses indicate evidence for FVIII and VWF clearance by each receptor.

ASGPR

Like many proteins, VWF glycan structures are well sialylated,97,98  suggesting that the contribution of asialoglycoprotein receptor (ASGPR) to basal VWF clearance is limited. However, age, diet, and body weight can all reduce the extent of sialylation,99,100  consequently favoring ASGPR-mediated clearance. Furthermore, several infectious pathogens are known to release sialidases (or neuramidases in the case of influenza infections). These sialidases could remove sialic acids from VWF, thereby promoting binding to ASGPR. Indeed, infection with Streptococcus pneumoniae reduces VWF sialylation and is associated with increased VWF clearance.101  Thus, the patient’s inflammatory state may potentially affect VWF clearance. Finally, the extent of sialylation highly depends on the activity of the enzymatic glycosylation machinery. For instance, reduced activity of the sialyltransferase ST3Gal-IV results in decreased VWF sialylation and subsequently increased clearance.102  Conversely, increased sialylation activity may promote interactions with sialic acid–recognizing receptors, such as sialic acid binding immunoglobulin-like lectin 5 (Siglec-5).103 

CLEC4M

Genome-wide association studies have shown that the CLEC4M gene is associated with VWF plasma levels,104  and later studies confirmed that the CLEC4M lectin receptor can bind to VWF and modulate VWF plasma levels.105  Interestingly, a relationship between the number of tandem repeats in the CLEC4M gene and levels of VWF activity and antigen was observed when analyzing plasma samples from VWD patients. Apparently, sequence variations in VWF receptors may modulate the efficiency by which they remove VWF from circulation.

LRP1

Although initially recognized as a FVIII clearance receptor,106,107  LRP1 has recently also been identified as a clearance receptor for VWF.108  Interestingly, VWF clearance by LRP1 depends strictly on the presence of increased hydrodynamic forces, suggesting that VWF needs to be unfolded in order to interact with LRP1. It has been hypothesized that changes in vascular integrity alter the shear stress to which VWF is exposed, thereby disturbing the clearance equilibrium between VWF and LRP1.108  Notably, although the contribution of LRP1 to VWF catabolism is modest (as for FVIII), it seems physiologically relevant, given that polymorphisms in the LRP1 gene are associated with VWF (and FVIII) plasma levels.109 

Parameters that modulate VWF clearance undoubtedly also affect FVIII clearance, either indirectly or directly. Indeed, there is a large overlap between potential clearance receptors that recognize both FVIII and VWF (Table 2). Interestingly, VWF often interferes with FVIII binding to its proper receptor in vitro, as has been shown for ASGPR,110  Siglec-5,103  and LRP1.108  This does not mean that FVIII never interacts alone with these receptors. It is important to realize that the FVIII-VWF interaction is highly dynamic, with fast association and dissociation rates (Figure 2). This dynamic equilibrium implies that FVIII is frequently dissociated from VWF and therefore can bind to its proper clearance receptor. Moreover, it allows FVIII to move from 1 VWF molecule to another, a phenomenon that Yee and coworkers have elegantly shown occurs physiologically.111 

Understanding VWF clearance is of relevance with regard to the development of long-acting FVIII variants (Table 3). These variants display limited prolongation of their half-life, most likely because they still interact with endogenous VWF and are cleared as part of the FVIII-VWF complex. Efforts to prolong FVIII half-life by modifying VWF to modulate its clearance have also met with limited success because FVIII can distribute to unmodified, endogenous VWF. Therefore, longer-acting FVIII variants may require modifications that exclude association with endogenous VWF.

Table 3

Novel, structurally altered rFVIII concentrates in development or recently approved, and the advantages and disadvantages of the technology used in their development

CharacteristicPEGylated FVIIIFc fusion platform
Novel FVIII design
N8-GPBAX 855*BAY 94-9027rFVIIIFcrVIII-SingleChain
Product description B-domain–modified 40K O-glycoPEGylated rFVIII concentrate 20-kDa PEGylated full-length rFVIII BDD rFVIII with a site-specific branched 60-kDa PEG side chain Recombinant BDD FVIII-Fc fusion protein Recombinant single-chain FVIII construct 
Spec.Act. 11 200 IU/mg Spec.Act. 8000 IU/mg Spec.Act. 9717 IU/mg Spec.Act. 9348 IU/mg Spec.Act. 12 000 IU/mg 
Advantages of technology • Well-established technology • Established technology • Increases the intrinsic stability of the FVIII molecule by reducing the potential dissociation of the heavy and light chains 
• Improves solubility of proteins • Extends half-life of proteins • Increased affinity for and binding to VWF 
• Extends half-life of non-PEG protein/drug • May mitigate immunogenicity of the protein 
• Reduced kidney clearance because of larger hydrodynamic size 
• Protects the drug from proteolysis 
• May protect the drug from the immune system 
• Reduces aggregation 
Disadvantages of technology • Random PEGylation (BAX 855) may result in loss or change in protein activity • Potential for activation of the immune system (antibody-dependent or complement-dependent) • Not specifically designed for extension of half-life 
• Preexisting anti-PEG antibodies identified in the healthy population, which may result in the rapid clearance of PEGylated compounds124  
Half-life (patients >12 y), h 19.04125  14.3-16126  ∼18.2-19.5127  ∼19114,116  — 
Percentage increase in half-life vs rFVIII 60%125 § 40%-50%126  28.1%-40.5%127  53%-70%114,116  — 
Clearance, mL/h/kg 1.79125  — — 1.68 33%-35% decrease vs rFVIII114  — 
CharacteristicPEGylated FVIIIFc fusion platform
Novel FVIII design
N8-GPBAX 855*BAY 94-9027rFVIIIFcrVIII-SingleChain
Product description B-domain–modified 40K O-glycoPEGylated rFVIII concentrate 20-kDa PEGylated full-length rFVIII BDD rFVIII with a site-specific branched 60-kDa PEG side chain Recombinant BDD FVIII-Fc fusion protein Recombinant single-chain FVIII construct 
Spec.Act. 11 200 IU/mg Spec.Act. 8000 IU/mg Spec.Act. 9717 IU/mg Spec.Act. 9348 IU/mg Spec.Act. 12 000 IU/mg 
Advantages of technology • Well-established technology • Established technology • Increases the intrinsic stability of the FVIII molecule by reducing the potential dissociation of the heavy and light chains 
• Improves solubility of proteins • Extends half-life of proteins • Increased affinity for and binding to VWF 
• Extends half-life of non-PEG protein/drug • May mitigate immunogenicity of the protein 
• Reduced kidney clearance because of larger hydrodynamic size 
• Protects the drug from proteolysis 
• May protect the drug from the immune system 
• Reduces aggregation 
Disadvantages of technology • Random PEGylation (BAX 855) may result in loss or change in protein activity • Potential for activation of the immune system (antibody-dependent or complement-dependent) • Not specifically designed for extension of half-life 
• Preexisting anti-PEG antibodies identified in the healthy population, which may result in the rapid clearance of PEGylated compounds124  
Half-life (patients >12 y), h 19.04125  14.3-16126  ∼18.2-19.5127  ∼19114,116  — 
Percentage increase in half-life vs rFVIII 60%125 § 40%-50%126  28.1%-40.5%127  53%-70%114,116  — 
Clearance, mL/h/kg 1.79125  — — 1.68 33%-35% decrease vs rFVIII114  — 

—, not available; BDD, B-domain deleted.

*

Adynovate (Baxalta US Inc), approved by the US Food and Drug Administration (FDA) in November 2015.

Eloctate (Biogen Idec Inc), approved by the FDA in June 2014.

Specific activities (Sp.Act.) have been derived using either chromogenic or 1-stage functional assays.

§

Versus patients’ previous treatment of plasma-derived FVIII or rFVIII.

PEGylation/GlycoPEGylation

PEGylation is being used to prolong the half-lives of many coagulation factors, and there are 3 PEGylated FVIII concentrates currently in development or recently approved: N8-GP, BAX 855 (Adynovate; Baxalta US Inc, Westlake Village, CA), and BAY 94-9027 (Table 1). Polyethylene glycol (PEG) can be selectively added to increase half-life by shielding therapeutic proteins from proteolytic enzymes, clearance receptors, and immune effector cells. Preclinical data in 2 animal models demonstrated that the FVIII-VWF interaction contributes to the longer half-life of PEGylated FVIII.112  The PEGylated FVIII products in development have half-lives that are all ∼1.5-fold longer than that of non-PEGylated FVIII in patients with hemophilia A (Table 3).

Fc fusion proteins

Another technique to increase half-lives of proteins and decrease potential immunogenicity is fusion of the therapeutic protein to a human immunoglobulin Fc fragment. The half-life is thereby extended through Fc interaction with the salvage neonatal Fc receptor.113  A recombinant Fc-FVIII construct (rFVIIIFc; Eloctate, Biogen Inc, Cambridge, MA) has been developed for the treatment of HA (Table 3). Clearance of rFVIIIFc was reduced, resulting in an extended half-life (∼1.5-fold) compared with rFVIII (Table 3).114-116  Furthermore, rFVIIIFc maintains normal FVIII interactions with other proteins necessary for its activity and prolonged in vivo half-life.117  A strong correlation was observed between VWF levels and rFVIIIFc clearance.114,115  For both PEGylated rFVIII and rFVIIIFc, clearance was decreased and half-life was prolonged as endogenous VWF levels increased. Thus, for both PEGylated and Fc-fusion forms of FVIII, clearance and half-life are still modulated by interactions with VWF, thereby limiting the degree to which these strategies can extend the FVIII half-life.

Single-chain design

The heterodimeric 2-chain form of FVIII is the physiologic configuration, but this is a labile structure that can dissociate spontaneously. Thus, a covalently linked single-chain version of FVIII (rVIII-SingleChain) was developed to improve the protein’s stability.

A clinical program for rVIII-SingleChain is under way, but prior preclinical studies have shown improved pharmacodynamic efficacy comparable to full-length and BDD human rFVIII.118  An unexpected observation was that the half-life of rVIII-SingleChain was approximately twofold greater than that of full-length rFVIII.119  This half-life extension may be attributed to the threefold higher affinity of rVIII-SingleChain for plasma-derived VWF compared with full-length rFVIII,120  thus resulting in a smaller proportion of unbound and rapidly cleared rFVIII molecules. This pharmacokinetic behavior has been replicated in a recent phase 1/3 clinical trial and was stable with repeated infusions.121 

Despite investigation of numerous approaches, the rFVIII products in late-stage clinical development have achieved only moderate increases in half-life compared with currently marketed rFVIII products (Table 3). The half-life of VWF (∼15 hours) seems to be the limiting factor, suggesting that FVIII may be subject to a dominant VWF-dependent clearance. To overcome the limitation of VWF dependence, 2 novel strategies are currently being investigated. One seeks to prolong the VWF half-life, and the other uses a strategy to limit VWF-dependent clearance of FVIII.

A VWF-albumin fusion protein has been expressed in mammalian cells and preliminary findings show this approach leads to a significantly longer VWF half-life in vivo.120  The second novel strategy to potentially circumvent endogenous VWF-dependent clearance of FVIII is the development of a FVIII molecule comprising 2 polypeptides: a single-chain BDD FVIII with XTEN inserted within the FVIII sequence, and the D’D3 FVIII-binding region of VWF. These polypeptides are fused to the Fc region of immunoglobulin G1 to enable the D′D3 fragment to correctly align to bind the FVIII moiety. This approach has shown favorable pharmacokinetics in animal models, including a fourfold increase in half-life compared with BDD-FVIII.122 

Lastly, it should be stated that another novel hemophilia therapeutic agent, emicizumab, a bispecific antibody that possesses FVIII mimetic activity (and is not intended to increase FVIII half-life), appears to have no interaction with VWF, and does not interfere with the treatment of breakthrough bleeding with infused FVIII.123 

The critical physiological importance of FVIII and VWF is highlighted by the inherited bleeding disorders HA and VWD. After 35 years of intensive investigation, we are beginning to understand some of the basic biological features of these 2 proteins and their respective and intertwined life cycles. Nevertheless, much still remains to be learned. Why is FVIII expression limited to certain types of endothelium? How are the 2 proteins cleared from plasma? Is there a role for VWF in modulating FVIII immunogenicity? These are just 3 of many unanswered questions. Solving these unknowns will not only advance our basic knowledge of hemostasis physiology, but may also provide enhanced opportunities to treat patients with HA and VWD.

D.L. is the recipient of research funding relating to FVIII and VWF from the Canadian Institutes of Health Research and holds a Canada Research Chair in Molecular Hemostasis. P.J.L. is a recipient of a grant from the Agence Nationale de la Recherche (ANR-13-BSV1-0014). K.P.P. acknowledges startup funding from Uniformed Services University of the Health Sciences.

This review was initiated at a face-to-face meeting for which travel support was provided by CSL. CSL did not select the coauthor group and there was no commercial involvement in considering the content or in the writing of this review.

Contribution: S.W.P., R.R.M., K.P.P., P.J.L., and D.L. wrote the manuscript and edited versions of the document.

Conflict-of-interest disclosure: S.W.P. has served as a consultant to Baxalta, CSL, Biogen, Novo Nordisk, Roche/Genentech, and Bayer. D.L. receives research support from Biogen, Bayer, CSL, Octapharma, and Sangamo. The remaining authors declare no competing financial interests.

Correspondence: David Lillicrap, Department of Pathology and Molecular Medicine, Richardson Laboratory, Queens University, Kingston, ON, Canada; e-mail: dpl@queensu.ca.

1
Do
 
H
Healey
 
JF
Waller
 
EK
Lollar
 
P
Expression of factor VIII by murine liver sinusoidal endothelial cells.
J Biol Chem
1999
, vol. 
274
 
28
(pg. 
19587
-
19592
)
2
Jacquemin
 
M
Neyrinck
 
A
Hermanns
 
MI
, et al. 
FVIII production by human lung microvascular endothelial cells.
Blood
2006
, vol. 
108
 
2
(pg. 
515
-
517
)
3
Schenone
 
M
Furie
 
BC
Furie
 
B
The blood coagulation cascade.
Curr Opin Hematol
2004
, vol. 
11
 
4
(pg. 
272
-
277
)
4
Terraube
 
V
O’Donnell
 
JS
Jenkins
 
PV
Factor VIII and von Willebrand factor interaction: biological, clinical and therapeutic importance.
Haemophilia
2010
, vol. 
16
 
1
(pg. 
3
-
13
)
5
Mannucci
 
PM
Duga
 
S
Peyvandi
 
F
Recessively inherited coagulation disorders.
Blood
2004
, vol. 
104
 
5
(pg. 
1243
-
1252
)
6
Schneppenheim
 
R
Budde
 
U
von Willebrand factor: the complex molecular genetics of a multidomain and multifunctional protein.
J Thromb Haemost
2011
, vol. 
9
 
suppl 1
(pg. 
209
-
215
)
7
Rosner
 
F
Hemophilia in the Talmud and rabbinic writings.
Ann Intern Med
1969
, vol. 
70
 
4
(pg. 
833
-
837
)
8
Lane
 
S
Successful transfusion of blood.
Lancet
1840
, vol. 
1
 (pg. 
185
-
188
)
9
Von Willebrand
 
EA
Hereditary pseudohaemophilia.
Haemophilia
1999
, vol. 
5
 
3
(pg. 
223
-
231
)
10
Barrow
 
EM
Roberts
 
HR
Pons
 
K
Graham
 
JB
Studies of the antihemophilic factor (Ahf, factor Viii) produced in von Willebrand’s disease.
Proc Soc Exp Biol Med
1964
, vol. 
115
 (pg. 
760
-
763
)
11
Switzer
 
ME
McKee
 
PA
Studies on human antihemophilic factor. Evidence for a covalently linked subunit structure.
J Clin Invest
1976
, vol. 
57
 
4
(pg. 
925
-
937
)
12
Zimmerman
 
TS
Edgington
 
TS
Factor VIII coagulant activity and factor VIII-like antigen: independent molecular entities.
J Exp Med
1973
, vol. 
138
 
4
(pg. 
1015
-
1020
)
13
Michaux
 
G
Cutler
 
DF
How to roll an endothelial cigar: the biogenesis of Weibel-Palade bodies.
Traffic
2004
, vol. 
5
 
2
(pg. 
69
-
78
)
14
Sporn
 
LA
Chavin
 
SI
Marder
 
VJ
Wagner
 
DD
Biosynthesis of von Willebrand protein by human megakaryocytes.
J Clin Invest
1985
, vol. 
76
 
3
(pg. 
1102
-
1106
)
15
Wagner
 
DD
Olmsted
 
JB
Marder
 
VJ
Immunolocalization of von Willebrand protein in Weibel-Palade bodies of human endothelial cells.
J Cell Biol
1982
, vol. 
95
 
1
(pg. 
355
-
360
)
16
Ruggeri
 
ZM
Zimmerman
 
TS
Variant von Willebrand’s disease: characterization of two subtypes by analysis of multimeric composition of factor VIII/von Willebrand factor in plasma and platelets.
J Clin Invest
1980
, vol. 
65
 
6
(pg. 
1318
-
1325
)
17
Mannucci
 
PM
Ruggeri
 
ZM
Pareti
 
FI
Capitanio
 
A
1-Deamino-8-d-arginine vasopressin: a new pharmacological approach to the management of haemophilia and von Willebrands’ diseases.
Lancet
1977
, vol. 
1
 
8017
(pg. 
869
-
872
)
18
Rendu
 
F
Brohard-Bohn
 
B
The platelet release reaction: granules’ constituents, secretion and functions.
Platelets
2001
, vol. 
12
 
5
(pg. 
261
-
273
)
19
Kaplan
 
KL
Owen
 
J
Plasma levels of platelet secretory proteins.
Crit Rev Oncol Hematol
1986
, vol. 
5
 
3
(pg. 
235
-
255
)
20
Flood
 
VH
Gill
 
JC
Christopherson
 
PA
, et al. 
Critical von Willebrand factor A1 domain residues influence type VI collagen binding.
J Thromb Haemost
2012
, vol. 
10
 
7
(pg. 
1417
-
1424
)
21
Ruggeri
 
ZM
Orje
 
JN
Habermann
 
R
Federici
 
AB
Reininger
 
AJ
Activation-independent platelet adhesion and aggregation under elevated shear stress.
Blood
2006
, vol. 
108
 
6
(pg. 
1903
-
1910
)
22
Coller
 
BS
Peerschke
 
EI
Scudder
 
LE
Sullivan
 
CA
Studies with a murine monoclonal antibody that abolishes ristocetin-induced binding of von Willebrand factor to platelets: additional evidence in support of GPIb as a platelet receptor for von Willebrand factor.
Blood
1983
, vol. 
61
 
1
(pg. 
99
-
110
)
23
Ruggeri
 
ZM
De Marco
 
L
Gatti
 
L
Bader
 
R
Montgomery
 
RR
Platelets have more than one binding site for von Willebrand factor.
J Clin Invest
1983
, vol. 
72
 
1
(pg. 
1
-
12
)
24
Schullek
 
J
Jordan
 
J
Montgomery
 
RR
Interaction of von Willebrand factor with human platelets in the plasma milieu.
J Clin Invest
1984
, vol. 
73
 
2
(pg. 
421
-
428
)
25
Bontempo
 
FA
Lewis
 
JH
Gorenc
 
TJ
, et al. 
Liver transplantation in hemophilia A.
Blood
1987
, vol. 
69
 
6
(pg. 
1721
-
1724
)
26
Marchioro
 
TL
Hougie
 
C
Ragde
 
H
Epstein
 
RB
Thomas
 
ED
Hemophilia: role of organ homografts.
Science
1969
, vol. 
163
 
3863
(pg. 
188
-
190
)
27
Groth
 
CG
Hathaway
 
WE
Gustafsson
 
A
, et al. 
Correction of coagulation in the hemophilic dog by transplantation of lymphatic tissue.
Surgery
1974
, vol. 
75
 
5
(pg. 
725
-
733
)
28
Norman
 
JC
Covelli
 
VH
Sise
 
HS
Transplantation of the spleen: experimental cure of hemophilia.
Surgery
1968
, vol. 
64
 
1
(pg. 
1
-
14
)
29
Lamont
 
PA
Ragni
 
MV
Lack of desmopressin (DDAVP) response in men with hemophilia A following liver transplantation.
J Thromb Haemost
2005
, vol. 
3
 
10
(pg. 
2259
-
2263
)
30
Madeira
 
CL
Layman
 
ME
de Vera
 
RE
Fontes
 
PA
Ragni
 
MV
Extrahepatic factor VIII production in transplant recipient of hemophilia donor liver.
Blood
2009
, vol. 
113
 
21
(pg. 
5364
-
5365
)
31
Hollestelle
 
MJ
Thinnes
 
T
Crain
 
K
, et al. 
Tissue distribution of factor VIII gene expression in vivo--a closer look.
Thromb Haemost
2001
, vol. 
86
 
3
(pg. 
855
-
861
)
32
Everett
 
LA
Cleuren
 
AC
Khoriaty
 
RN
Ginsburg
 
D
Murine coagulation factor VIII is synthesized in endothelial cells.
Blood
2014
, vol. 
123
 
24
(pg. 
3697
-
3705
)
33
Fahs
 
SA
Hille
 
MT
Shi
 
Q
Weiler
 
H
Montgomery
 
RR
A conditional knockout mouse model reveals endothelial cells as the principal and possibly exclusive source of plasma factor VIII.
Blood
2014
, vol. 
123
 
24
(pg. 
3706
-
3713
)
34
Shi
 
Q
Fahs
 
SA
Kuether
 
EL
Cooley
 
BC
Weiler
 
H
Montgomery
 
RR
Targeting FVIII expression to endothelial cells regenerates a releasable pool of FVIII and restores hemostasis in a mouse model of hemophilia A.
Blood
2010
, vol. 
116
 
16
(pg. 
3049
-
3057
)
35
Turner
 
NA
Moake
 
JL
Factor VIII is synthesized in human endothelial cells, packaged in Weibel-Palade bodies and secreted bound to ULVWF strings.
PLoS One
2015
, vol. 
10
 
10
pg. 
e0140740
 
36
Pan
 
J
Dinh
 
TT
Rajaraman
 
A
, et al. 
Patterns of expression of factor VIII and von Willebrand factor by endothelial cell subsets in vivo.
Blood
2016
, vol. 
128
 
1
(pg. 
104
-
109
)
37
Montgomery
 
RR
Gill
 
JC
Interactions between von Willebrand factor and factor VIII: where did they first meet.
J Pediatr Hematol Oncol
2000
, vol. 
22
 
3
(pg. 
269
-
275
)
38
Rosenberg
 
JB
Foster
 
PA
Kaufman
 
RJ
, et al. 
Intracellular trafficking of factor VIII to von Willebrand factor storage granules.
J Clin Invest
1998
, vol. 
101
 
3
(pg. 
613
-
624
)
39
Montgomery
 
RR
Hathaway
 
WE
Johnson
 
J
Jacobson
 
L
Muntean
 
W
A variant of von Willebrand’s disease with abnormal expression of factor VIII procoagulant activity.
Blood
1982
, vol. 
60
 
1
(pg. 
201
-
207
)
40
Tuley
 
EA
Gaucher
 
C
Jorieux
 
S
Worrall
 
NK
Sadler
 
JE
Mazurier
 
C
Expression of von Willebrand factor “Normandy”: an autosomal mutation that mimics hemophilia A.
Proc Natl Acad Sci USA
1991
, vol. 
88
 
14
(pg. 
6377
-
6381
)
41
Mazurier
 
C
Gaucher
 
C
Jorieux
 
S
Goudemand
 
M
Collaborative Group
Biological effect of desmopressin in eight patients with type 2N (‘Normandy’) von Willebrand disease.
Br J Haematol
1994
, vol. 
88
 
4
(pg. 
849
-
854
)
42
van den Biggelaar
 
M
Meijer
 
AB
Voorberg
 
J
Mertens
 
K
Intracellular cotrafficking of factor VIII and von Willebrand factor type 2N variants to storage organelles.
Blood
2009
, vol. 
113
 
13
(pg. 
3102
-
3109
)
43
Nathwani
 
AC
Reiss
 
UM
Tuddenham
 
EG
, et al. 
Long-term safety and efficacy of factor IX gene therapy in hemophilia B.
N Engl J Med
2014
, vol. 
371
 
21
(pg. 
1994
-
2004
)
44
McIntosh
 
J
Lenting
 
PJ
Rosales
 
C
, et al. 
Therapeutic levels of FVIII following a single peripheral vein administration of rAAV vector encoding a novel human factor VIII variant.
Blood
2013
, vol. 
121
 
17
(pg. 
3335
-
3344
)
45
Rosenberg
 
JB
Greengard
 
JS
Montgomery
 
RR
Genetic induction of a releasable pool of factor VIII in human endothelial cells.
Arterioscler Thromb Vasc Biol
2000
, vol. 
20
 
12
(pg. 
2689
-
2695
)
46
Matsui
 
H
Shibata
 
M
Brown
 
B
, et al. 
Ex vivo gene therapy for hemophilia A that enhances safe delivery and sustained in vivo factor VIII expression from lentivirally engineered endothelial progenitors.
Stem Cells
2007
, vol. 
25
 
10
(pg. 
2660
-
2669
)
47
Shi
 
Q
Schroeder
 
JA
Kuether
 
EL
Montgomery
 
RR
The important role of von Willebrand factor in platelet-derived FVIII gene therapy for murine hemophilia A in the presence of inhibitory antibodies.
J Thromb Haemost
2015
, vol. 
13
 
7
(pg. 
1301
-
1309
)
48
Ngo
 
JC
Huang
 
M
Roth
 
DA
Furie
 
BC
Furie
 
B
Crystal structure of human factor VIII: implications for the formation of the factor IXa-factor VIIIa complex.
Structure
2008
, vol. 
16
 
4
(pg. 
597
-
606
)
49
Pratt
 
KP
Shen
 
BW
Takeshima
 
K
Davie
 
EW
Fujikawa
 
K
Stoddard
 
BL
Structure of the C2 domain of human factor VIII at 1.5 A resolution.
Nature
1999
, vol. 
402
 
6760
(pg. 
439
-
442
)
50
Shen
 
BW
Spiegel
 
PC
Chang
 
CH
, et al. 
The tertiary structure and domain organization of coagulation factor VIII.
Blood
2008
, vol. 
111
 
3
(pg. 
1240
-
1247
)
51
Wakabayashi
 
H
Fay
 
PJ
Molecular orientation of factor VIIIa on the phospholipid membrane surface determined by fluorescence resonance energy transfer.
Biochem J
2013
, vol. 
452
 
2
(pg. 
293
-
301
)
52
Stoilova-McPhie
 
S
Lynch
 
GC
Ludtke
 
S
Pettitt
 
BM
Domain organization of membrane-bound factor VIII.
Biopolymers
2013
, vol. 
99
 
7
(pg. 
448
-
459
)
53
Yee
 
A
Kretz
 
CA
Von Willebrand factor: form for function.
Semin Thromb Hemost
2014
, vol. 
40
 
1
(pg. 
17
-
27
)
54
Liu
 
ML
Shen
 
BW
Nakaya
 
S
, et al. 
Hemophilic factor VIII C1- and C2-domain missense mutations and their modeling to the 1.5-angstrom human C2-domain crystal structure.
Blood
2000
, vol. 
96
 
3
(pg. 
979
-
987
)
55
Shiltagh
 
N
Kirkpatrick
 
J
Cabrita
 
LD
, et al. 
Solution structure of the major factor VIII binding region on von Willebrand factor.
Blood
2014
, vol. 
123
 
26
(pg. 
4143
-
4151
)
56
Yee
 
A
Oleskie
 
AN
Dosey
 
AM
, et al. 
Visualization of an N-terminal fragment of von Willebrand factor in complex with factor VIII.
Blood
2015
, vol. 
126
 
8
(pg. 
939
-
942
)
57
Chiu
 
PL
Bou-Assaf
 
GM
Chhabra
 
ES
, et al. 
Mapping the interaction between factor VIII and von Willebrand factor by electron microscopy and mass spectrometry.
Blood
2015
, vol. 
126
 
8
(pg. 
935
-
938
)
58
Leyte
 
A
van Schijndel
 
HB
Niehrs
 
C
, et al. 
Sulfation of Tyr1680 of human blood coagulation factor VIII is essential for the interaction of factor VIII with von Willebrand factor.
J Biol Chem
1991
, vol. 
266
 
2
(pg. 
740
-
746
)
59
Nielsen
 
PF
Bak
 
S
Vandahl
 
B
Characterization of tyrosine sulphation in rFVIII (turoctocog alfa) expressed in CHO and HEK-293 cells.
Haemophilia
2012
, vol. 
18
 
5
(pg. 
e397
-
e398
)
60
Kannicht
 
C
Ramström
 
M
Kohla
 
G
, et al. 
Characterisation of the post-translational modifications of a novel, human cell line-derived recombinant human factor VIII.
Thromb Res
2013
, vol. 
131
 
1
(pg. 
78
-
88
)
61
Scharrer
 
I
Bray
 
GL
Neutzling
 
O
Incidence of inhibitors in haemophilia A patients--a review of recent studies of recombinant and plasma-derived factor VIII concentrates.
Haemophilia
1999
, vol. 
5
 
3
(pg. 
145
-
154
)
62
Scandella
 
D
Human anti-factor VIII antibodies: epitope localization and inhibitory function.
Vox Sang
1996
, vol. 
70
 
suppl 1
(pg. 
9
-
14
)
63
Kahle
 
J
Orlowski
 
A
Stichel
 
D
, et al. 
Epitope mapping via selection of anti-FVIII antibody-specific phage-presented peptide ligands that mimic the antibody binding sites.
Thromb Haemost
2015
, vol. 
113
 
2
(pg. 
396
-
405
)
64
Markovitz
 
RC
Healey
 
JF
Parker
 
ET
Meeks
 
SL
Lollar
 
P
The diversity of the immune response to the A2 domain of human factor VIII.
Blood
2013
, vol. 
121
 
14
(pg. 
2785
-
2795
)
65
Nguyen
 
PC
Lewis
 
KB
Ettinger
 
RA
, et al. 
High-resolution mapping of epitopes on the C2 domain of factor VIII by analysis of point mutants using surface plasmon resonance.
Blood
2014
, vol. 
123
 
17
(pg. 
2732
-
2739
)
66
Pratt
 
KP
Inhibitory antibodies in hemophilia A.
Curr Opin Hematol
2012
, vol. 
19
 
5
(pg. 
399
-
405
)
67
Lin
 
JC
Ettinger
 
RA
Schuman
 
JT
, et al. 
Six amino acid residues in a 1200 Å2 interface mediate binding of factor VIII to an IgG4κ inhibitory antibody.
PLoS One
2015
, vol. 
10
 
1
pg. 
e0116577
 
68
Astermark
 
J
Inhibitor development: patient-determined risk factors.
Haemophilia
2010
, vol. 
16
 
102
(pg. 
66
-
70
)
69
Ghosh
 
K
Shetty
 
S
Immune response to FVIII in hemophilia A: an overview of risk factors.
Clin Rev Allergy Immunol
2009
, vol. 
37
 
2
(pg. 
58
-
66
)
70
Gouw
 
SC
van der Bom
 
JG
Ljung
 
R
, et al. 
PedNet and RODIN Study Group
Factor VIII products and inhibitor development in severe hemophilia A.
N Engl J Med
2013
, vol. 
368
 
3
(pg. 
231
-
239
)
71
Franchini
 
M
Coppola
 
A
Rocino
 
A
, et al. 
Italian Association of Hemophilia Centers (AICE) Working Group
Systematic review of the role of FVIII concentrates in inhibitor development in previously untreated patients with severe hemophilia a: a 2013 update.
Semin Thromb Hemost
2013
, vol. 
39
 
7
(pg. 
752
-
766
)
72
Ettingshausen
 
CE
Kreuz
 
W
Recombinant vs. plasma-derived products, especially those with intact VWF, regarding inhibitor development.
Haemophilia
2006
, vol. 
12
 
suppl 6
(pg. 
102
-
106
)
73
Klukowska
 
A
Szczepański
 
T
Vdovin
 
V
Knaub
 
S
Jansen
 
M
Liesner
 
R
Novel, human cell line-derived recombinant factor VIII (human-cl rhFVIII, Nuwiq(®)) in children with severe haemophilia A: efficacy, safety and pharmacokinetics [published online ahead of print September 14, 2015].
Haemophilia
74
Lissitchkov
 
T
Hampton
 
K
von Depka
 
M
, et al. 
Novel, human cell line-derived recombinant factor VIII (human-cl rhFVIII; Nuwiq(®)) in adults with severe haemophilia A: efficacy and safety [published online ahead of print August 28, 2015].
Haemophilia
75
Dasgupta
 
S
Repessé
 
Y
Bayry
 
J
, et al. 
VWF protects FVIII from endocytosis by dendritic cells and subsequent presentation to immune effectors.
Blood
2007
, vol. 
109
 
2
(pg. 
610
-
612
)
76
Delignat
 
S
Repessé
 
Y
Navarrete
 
AM
, et al. 
Immunoprotective effect of von Willebrand factor towards therapeutic factor VIII in experimental haemophilia A.
Haemophilia
2012
, vol. 
18
 
2
(pg. 
248
-
254
)
77
Sorvillo
 
N
Hartholt
 
RB
Bloem
 
E
, et al. 
von Willebrand factor binds to the surface of dendritic cells and modulates peptide presentation of factor VIII.
Haematologica
2016
, vol. 
101
 
3
(pg. 
309
-
318
)
78
Gharagozlou
 
S
Sharifian
 
RA
Khoshnoodi
 
J
, et al. 
Epitope specificity of anti-factor VIII antibodies from inhibitor positive acquired and congenital haemophilia A patients using synthetic peptides spanning A and C domains.
Thromb Haemost
2009
, vol. 
101
 
5
(pg. 
834
-
839
)
79
Meeks
 
SL
Cox
 
CL
Healey
 
JF
, et al. 
A major determinant of the immunogenicity of factor VIII in a murine model is independent of its procoagulant function.
Blood
2012
, vol. 
120
 
12
(pg. 
2512
-
2520
)
80
Lin
 
Y
Yang
 
X
Chevrier
 
MC
, et al. 
Relationships between factor VIII:Ag and factor VIII in recombinant and plasma-derived factor VIII concentrates.
Haemophilia
2004
, vol. 
10
 
5
(pg. 
459
-
469
)
81
Behrmann
 
M
Pasi
 
J
Saint-Remy
 
JM
Kotitschke
 
R
Kloft
 
M
Von Willebrand factor modulates factor VIII immunogenicity: comparative study of different factor VIII concentrates in a haemophilia A mouse model.
Thromb Haemost
2002
, vol. 
88
 
2
(pg. 
221
-
229
)
82
Goudemand
 
J
Laurian
 
Y
Calvez
 
T
Risk of inhibitors in haemophilia and the type of factor replacement.
Curr Opin Hematol
2006
, vol. 
13
 
5
(pg. 
316
-
322
)
83
Iorio
 
A
Halimeh
 
S
Holzhauer
 
S
, et al. 
Rate of inhibitor development in previously untreated hemophilia A patients treated with plasma-derived or recombinant factor VIII concentrates: a systematic review.
J Thromb Haemost
2010
, vol. 
8
 
6
(pg. 
1256
-
1265
)
84
Goudemand
 
J
Rothschild
 
C
Demiguel
 
V
, et al. 
FVIII-LFB and Recombinant FVIII study groups
Influence of the type of factor VIII concentrate on the incidence of factor VIII inhibitors in previously untreated patients with severe hemophilia A.
Blood
2006
, vol. 
107
 
1
(pg. 
46
-
51
)
85
Gouw
 
SC
van der Bom
 
JG
Auerswald
 
G
Ettinghausen
 
CE
Tedgård
 
U
van den Berg
 
HM
Recombinant versus plasma-derived factor VIII products and the development of inhibitors in previously untreated patients with severe hemophilia A: the CANAL cohort study.
Blood
2007
, vol. 
109
 
11
(pg. 
4693
-
4697
)
86
Peyvandi
 
F
Mannucci
 
PM
Garagiola
 
I
, et al. 
A randomized trial of factor VIII and neutralizing antibodies in hemophilia A.
N Engl J Med
2016
, vol. 
374
 
21
(pg. 
2054
-
2064
)
87
d’Oiron
 
R
Pipe
 
SW
Jacquemin
 
M
Mild/moderate haemophilia A: new insights into molecular mechanisms and inhibitor development.
Haemophilia
2008
, vol. 
14
 
suppl 3
(pg. 
138
-
146
)
88
James
 
EA
Kwok
 
WW
Ettinger
 
RA
Thompson
 
AR
Pratt
 
KP
T-cell responses over time in a mild hemophilia A inhibitor subject: epitope identification and transient immunogenicity of the corresponding self-peptide.
J Thromb Haemost
2007
, vol. 
5
 
12
(pg. 
2399
-
2407
)
89
James
 
EA
van Haren
 
SD
Ettinger
 
RA
, et al. 
T-cell responses in two unrelated hemophilia A inhibitor subjects include an epitope at the factor VIII R593C missense site.
J Thromb Haemost
2011
, vol. 
9
 
4
(pg. 
689
-
699
)
90
Kim
 
YC
Zhang
 
AH
Su
 
Y
, et al. 
Engineered antigen-specific human regulatory T cells: immunosuppression of FVIII-specific T- and B-cell responses.
Blood
2015
, vol. 
125
 
7
(pg. 
1107
-
1115
)
91
Scott
 
DW
Pratt
 
KP
Miao
 
CH
Progress toward inducing immunologic tolerance to factor VIII.
Blood
2013
, vol. 
121
 
22
(pg. 
4449
-
4456
)
92
Sabatino
 
DE
Nichols
 
TC
Merricks
 
E
Bellinger
 
DA
Herzog
 
RW
Monahan
 
PE
Animal models of hemophilia.
Prog Mol Biol Transl Sci
2012
, vol. 
105
 (pg. 
151
-
209
)
93
Lei
 
TC
Scott
 
DW
Induction of tolerance to factor VIII inhibitors by gene therapy with immunodominant A2 and C2 domains presented by B cells as Ig fusion proteins.
Blood
2005
, vol. 
105
 
12
(pg. 
4865
-
4870
)
94
Jiang
 
H
Lillicrap
 
D
Patarroyo-White
 
S
, et al. 
Multiyear therapeutic benefit of AAV serotypes 2, 6, and 8 delivering factor VIII to hemophilia A mice and dogs.
Blood
2006
, vol. 
108
 
1
(pg. 
107
-
115
)
95
Matsui
 
H
Shibata
 
M
Brown
 
B
, et al. 
A murine model for induction of long-term immunologic tolerance to factor VIII does not require persistent detectable levels of plasma factor VIII and involves contributions from Foxp3+ T regulatory cells.
Blood
2009
, vol. 
114
 
3
(pg. 
677
-
685
)
96
Di Paola
 
J
Lethagen
 
S
Gill
 
J
, et al. 
Presurgical pharmacokinetic analysis of a von Willebrand factor/factor VIII (VWF/FVIII) concentrate in patients with von Willebrand’s disease (VWD) has limited value in dosing for surgery.
Haemophilia
2011
, vol. 
17
 
5
(pg. 
752
-
758
)
97
Canis
 
K
McKinnon
 
TA
Nowak
 
A
, et al. 
Mapping the N-glycome of human von Willebrand factor.
Biochem J
2012
, vol. 
447
 
2
(pg. 
217
-
228
)
98
Canis
 
K
McKinnon
 
TA
Nowak
 
A
, et al. 
The plasma von Willebrand factor O-glycome comprises a surprising variety of structures including ABH antigens and disialosyl motifs.
J Thromb Haemost
2010
, vol. 
8
 
1
(pg. 
137
-
145
)
99
Gornik
 
O
Wagner
 
J
Pucić
 
M
Knezević
 
A
Redzic
 
I
Lauc
 
G
Stability of N-glycan profiles in human plasma.
Glycobiology
2009
, vol. 
19
 
12
(pg. 
1547
-
1553
)
100
Knezevic
 
A
Gornik
 
O
Polasek
 
O
, et al. 
Effects of aging, body mass index, plasma lipid profiles, and smoking on human plasma N-glycans.
Glycobiology
2010
, vol. 
20
 
8
(pg. 
959
-
969
)
101
Grewal
 
PK
Uchiyama
 
S
Ditto
 
D
, et al. 
The Ashwell receptor mitigates the lethal coagulopathy of sepsis.
Nat Med
2008
, vol. 
14
 
6
(pg. 
648
-
655
)
102
Ellies
 
LG
Ditto
 
D
Levy
 
GG
, et al. 
Sialyltransferase ST3Gal-IV operates as a dominant modifier of hemostasis by concealing asialoglycoprotein receptor ligands.
Proc Natl Acad Sci USA
2002
, vol. 
99
 
15
(pg. 
10042
-
10047
)
103
Pegon
 
JN
Kurdi
 
M
Casari
 
C
, et al. 
Factor VIII and von Willebrand factor are ligands for the carbohydrate-receptor Siglec-5.
Haematologica
2012
, vol. 
97
 
12
(pg. 
1855
-
1863
)
104
Smith
 
NL
Chen
 
MH
Dehghan
 
A
, et al. 
Wellcome Trust Case Control Consortium
Novel associations of multiple genetic loci with plasma levels of factor VII, factor VIII, and von Willebrand factor: The CHARGE (Cohorts for Heart and Aging Research in Genome Epidemiology) Consortium.
Circulation
2010
, vol. 
121
 
12
(pg. 
1382
-
1392
)
105
Rydz
 
N
Swystun
 
LL
Notley
 
C
, et al. 
The C-type lectin receptor CLEC4M binds, internalizes, and clears von Willebrand factor and contributes to the variation in plasma von Willebrand factor levels.
Blood
2013
, vol. 
121
 
26
(pg. 
5228
-
5237
)
106
Lenting
 
PJ
Neels
 
JG
van den Berg
 
BM
, et al. 
The light chain of factor VIII comprises a binding site for low density lipoprotein receptor-related protein.
J Biol Chem
1999
, vol. 
274
 
34
(pg. 
23734
-
23739
)
107
Saenko
 
EL
Yakhyaev
 
AV
Mikhailenko
 
I
Strickland
 
DK
Sarafanov
 
AG
Role of the low density lipoprotein-related protein receptor in mediation of factor VIII catabolism.
J Biol Chem
1999
, vol. 
274
 
53
(pg. 
37685
-
37692
)
108
Rastegarlari
 
G
Pegon
 
JN
Casari
 
C
, et al. 
Macrophage LRP1 contributes to the clearance of von Willebrand factor.
Blood
2012
, vol. 
119
 
9
(pg. 
2126
-
2134
)
109
Morange
 
PE
Tregouet
 
DA
Frere
 
C
, et al. 
Biological and genetic factors influencing plasma factor VIII levels in a healthy family population: results from the Stanislas cohort.
Br J Haematol
2005
, vol. 
128
 
1
(pg. 
91
-
99
)
110
Bovenschen
 
N
Rijken
 
DC
Havekes
 
LM
van Vlijmen
 
BJ
Mertens
 
K
The B domain of coagulation factor VIII interacts with the asialoglycoprotein receptor.
J Thromb Haemost
2005
, vol. 
3
 
6
(pg. 
1257
-
1265
)
111
Yee
 
A
Gildersleeve
 
RD
Gu
 
S
, et al. 
A von Willebrand factor fragment containing the D’D3 domains is sufficient to stabilize coagulation factor VIII in mice.
Blood
2014
, vol. 
124
 
3
(pg. 
445
-
452
)
112
Tang
 
L
Leong
 
L
Sim
 
D
, et al. 
von Willebrand factor contributes to longer half-life of PEGylated factor VIII in vivo.
Haemophilia
2013
, vol. 
19
 
4
(pg. 
539
-
545
)
113
Roopenian
 
DC
Akilesh
 
S
FcRn: the neonatal Fc receptor comes of age.
Nat Rev Immunol
2007
, vol. 
7
 
9
(pg. 
715
-
725
)
114
Mahlangu
 
J
Powell
 
JS
Ragni
 
MV
, et al. 
A-LONG Investigators
Phase 3 study of recombinant factor VIII Fc fusion protein in severe hemophilia A.
Blood
2014
, vol. 
123
 
3
(pg. 
317
-
325
)
115
Shapiro
 
AD
Ragni
 
MV
Kulkarni
 
R
, et al. 
Recombinant factor VIII Fc fusion protein: extended-interval dosing maintains low bleeding rates and correlates with von Willebrand factor levels.
J Thromb Haemost
2014
, vol. 
12
 
11
(pg. 
1788
-
1800
)
116
Powell
 
JS
Josephson
 
NC
Quon
 
D
, et al. 
Safety and prolonged activity of recombinant factor VIII Fc fusion protein in hemophilia A patients.
Blood
2012
, vol. 
119
 
13
(pg. 
3031
-
3037
)
117
Peters
 
RT
Toby
 
G
Lu
 
Q
, et al. 
Biochemical and functional characterization of a recombinant monomeric factor VIII-Fc fusion protein.
J Thromb Haemost
2013
, vol. 
11
 
1
(pg. 
132
-
141
)
118
Zollner
 
SB
Raquet
 
E
Müller-Cohrs
 
J
, et al. 
Preclinical efficacy and safety of rVIII-SingleChain (CSL627), a novel recombinant single-chain factor VIII.
Thromb Res
2013
, vol. 
132
 
2
(pg. 
280
-
287
)
119
Zollner
 
S
Raquet
 
E
Claar
 
P
, et al. 
Non-clinical pharmacokinetics and pharmacodynamics of rVIII-SingleChain, a novel recombinant single-chain factor VIII.
Thromb Res
2014
, vol. 
134
 
1
(pg. 
125
-
131
)
120
Schulte
 
S
Innovative coagulation factors: albumin fusion technology and recombinant single-chain factor VIII.
Thromb Res
2013
, vol. 
131
 
suppl 2
(pg. 
S2
-
S6
)
121
Mahlangu
 
J
Kuliczkowski
 
K
Karim
 
FA
, et al. 
AFFINITY Investigators
Efficacy and safety of rVIII-SingleChain: results of a phase 1/3 multicenter clinical trial in severe hemophilia A.
Blood
2016
, vol. 
128
 
5
(pg. 
630
-
637
)
122
Liu
 
T
Chhabra
 
ES
Kulman
 
J
, et al. 
Prolonged efficacy in hemophilia A mouse bleeding models of a recombinant FVIII-XTEND/D’D3 heterodimer with four-fold extended half-life in circulation.
Haemophilia
2014
, vol. 
20
 
s3
pg. 
76
 
123
Shima
 
M
Hanabusa
 
H
Taki
 
M
, et al. 
Factor VIII-mimetic function of humanized bispecific antibody in hemophilia A.
N Engl J Med
2016
, vol. 
374
 
21
(pg. 
2044
-
2053
)
124
Schellekens
 
H
Hennink
 
WE
Brinks
 
V
The immunogenicity of polyethylene glycol: facts and fiction.
Pharm Res
2013
, vol. 
30
 
7
(pg. 
1729
-
1734
)
125
Tiede
 
A
Brand
 
B
Fischer
 
R
, et al. 
Enhancing the pharmacokinetic properties of recombinant factor VIII: first-in-human trial of glycoPEGylated recombinant factor VIII in patients with hemophilia A.
J Thromb Haemost
2013
, vol. 
11
 
4
(pg. 
670
-
678
)
126
Konkle
 
BA
Stasyshyn
 
O
Chowdary
 
P
, et al. 
Pegylated, full-length, recombinant factor VIII for prophylactic and on-demand treatment of severe hemophilia A.
Blood
2015
, vol. 
126
 
9
(pg. 
1078
-
1085
)
127
Coyle
 
TE
Reding
 
MT
Lin
 
JC
Michaels
 
LA
Shah
 
A
Powell
 
J
Phase I study of BAY 94-9027, a PEGylated B-domain-deleted recombinant factor VIII with an extended half-life, in subjects with hemophilia A.
J Thromb Haemost
2014
, vol. 
12
 
4
(pg. 
488
-
496
)
128
Lenting
 
PJ
van Schooten
 
CJ
Denis
 
CV
Clearance mechanisms of von Willebrand factor and factor VIII.
J Thromb Haemost
2007
, vol. 
5
 
7
(pg. 
1353
-
1360
)
129
Dimitrov
 
JD
Christophe
 
OD
Kang
 
J
, et al. 
Thermodynamic analysis of the interaction of factor VIII with von Willebrand factor.
Biochemistry
2012
, vol. 
51
 
20
(pg. 
4108
-
4116
)
Sign in via your Institution