The most significant complication of treatment in patients with hemophilia A is the development of alloantibodies that inhibit factor VIII activity. In the presence of inhibitory antibodies, replacement of the missing clotting factor by infusion of factor VIII becomes less effective. Once replacement therapy is ineffective, acute management of bleeding requires agents that bypass factor VIII activity. Long-term management consists of eradicating the inhibitor through immune tolerance. Despite success in the treatment of acute bleeding and inhibitor eradication, there remains an inability to predict or prevent inhibitor formation. Ideally, prediction and ultimately prevention will come with an improved understanding of how patient-specific and treatment-related factors work together to influence anti–factor VIII antibody production.

Hemophilia A (HA) is an X-linked congenital bleeding disorder resulting from a deficiency of factor VIII (fVIII). Therapy to prevent or treat bleeding is replacement of fVIII. The availability of purified plasma-derived and recombinant fVIII products has led to dramatic improvements in the health and well-being of many affected by HA. However, as a consequence of treatment patients with HA may develop inhibitory IgG antibodies to fVIII, termed inhibitors. Inhibitors bind fVIII and prevent its hemostatic action. When this occurs, treatment becomes more costly and morbidity increases. Inhibitor formation, occurring in up to 36% of patients with severe HA,1,2  is currently one of the most significant complications affecting patients with HA.

Despite understanding several well-established risk factors for inhibitor development (Table 1), why some patients develop an inhibitor and others do not remains unclear. This lack of clarity is likely a consequence of the complex interplay between host genetic factors and the circumstances that surround the delivery of fVIII. In this review we will discuss the detection of inhibitors, the current understanding of why inhibitors develop, and management of patients with inhibitors during acute bleeding and long-term inhibitor eradication. Our discussion will focus on HA.

Table 1

Hypothesized risk factors for inhibitor development

Patient-relatedTreatment-related
Race Number of fVIII exposure days 
Family history Age at first exposure to fVIII concentrates 
fVIII mutation Type of fVIII concentrate used 
MHC class Concurrent infection/inflammatory state 
Polymorphisms of immune-response genes (IL10, TNF, CTLA4) Intensive exposure to fVIII concentrates 
Patient-relatedTreatment-related
Race Number of fVIII exposure days 
Family history Age at first exposure to fVIII concentrates 
fVIII mutation Type of fVIII concentrate used 
MHC class Concurrent infection/inflammatory state 
Polymorphisms of immune-response genes (IL10, TNF, CTLA4) Intensive exposure to fVIII concentrates 

Inhibitors should be suspected when there is a lack of response to fVIII infusion as a result of poor recovery, shortened half-life, or inadequate clinical response. When an inhibitor is suspected, testing using a Bethesda inhibitor assay (BIA) should be performed. It is also generally accepted that inhibitor screening should occur before invasive procedures and at regular intervals during the initial 50 treatment days as this is the highest risk period for inhibitor development.1  After a patient has received factor for 150 treatment days, the rate of inhibitor development is substantially reduced.3  Although rates of detection are low with routine surveillance in those with greater than 150 treatment days, we recommend annual testing to facilitate postmarketing surveillance by the Centers for Disease Control and Prevention Universal Data Collection (UDC) Project and other national surveillance programs.4 

The BIA consists of incubating a 1:1 mixture of a dilution of the patient's plasma with undiluted normal plasma for 2 hours, followed by assay of residual fVIII activity.5  The control incubation consists of a 1:1 mixture of the buffer diluent and undiluted normal plasma. The inhibitor titer is the reciprocal of the dilution of inhibitor plasma that neutralizes 50% of fVIII activity in the normal plasma. Inhibitor titers that are less than 5 BU/mL are considered low titer, whereas those that are equal to or greater than 5 BU/mL are considered high titer. A low-responding inhibitor is one in which the titer remains less than 5 BU/mL despite repeated fVIII infusions and, once it is equal to or greater than 5 BU/mL at any time, it is considered high responding.6  In the absence of fVIII exposure, high-responding inhibitors may decrease and may even become undetectable. Classically, when these patients are re-exposed to fVIII, their titer will increase over 4 to 7 days. This response is called anamnesis and is a hallmark of a high-responding inhibitor. However, we have rarely observed patients with historically high-responding inhibitors but who have had an undetectable titer for years who do not have anamnesis upon re-exposure to fVIII. Low-titer inhibitors comprise 25% to 50% of observed inhibitors and approximately 10% of these are considered transient, disappearing over weeks to months despite continued treatment with fVIII.7-10 

Why some patients develop inhibitors is poorly understood. The inability to predict inhibitor development may reflect the complexity of interactions involved in an immunologic response to a foreign protein. Tolerance allows the differentiation of self from nonself, and in the absence of sufficient protein to produce tolerance, patients with hemophilia will recognize infused clotting factor as nonself. The propensity to develop an inhibitor is likely influenced by congenital or acquired variances at multiple steps in this cellular immune cascade that begins with antigen uptake by antigen presenting cells and ends with antibody production (Figure 1).

Figure 1

MHC class II cellular immune cascade. Exogenous peptide antigens such as fVIII are processed through MHC class II mechanisms. Antigen-presenting cells take fVIII into endocytic vesicles where it is bound to an MHC class II molecule. Bound peptides are then presented on the surface of the cell to specific T-cell receptors (TCR) on CD4+ T lymphocytes. In response to antigen presentation, T lymphocytes elaborate cytokines and up-regulate several surface molecules. These surface molecules interact with corresponding proteins on B lymphocytes, leading to maturation of B cells and antibody formation. APC indicates antigen-presenting cell; CD4, CD4+ T lymphocyte; B cell, B lymphocyte; MHC, major histocompatibility complex; TCR, T-cell receptor; CD, cluster of differentiation; IL, interleukin. Copyright G.C. White II.

Figure 1

MHC class II cellular immune cascade. Exogenous peptide antigens such as fVIII are processed through MHC class II mechanisms. Antigen-presenting cells take fVIII into endocytic vesicles where it is bound to an MHC class II molecule. Bound peptides are then presented on the surface of the cell to specific T-cell receptors (TCR) on CD4+ T lymphocytes. In response to antigen presentation, T lymphocytes elaborate cytokines and up-regulate several surface molecules. These surface molecules interact with corresponding proteins on B lymphocytes, leading to maturation of B cells and antibody formation. APC indicates antigen-presenting cell; CD4, CD4+ T lymphocyte; B cell, B lymphocyte; MHC, major histocompatibility complex; TCR, T-cell receptor; CD, cluster of differentiation; IL, interleukin. Copyright G.C. White II.

Close modal

Genetic risk factors

The best characterized risk factor is the type of fVIII mutation that underlies hemophilia. Mutations that are associated with a high prevalence of inhibitors include null mutations, large deletions, nonsense mutations, and intron 22 inversions (Table 2).11 

Table 2

FVIII mutations and inhibitor prevalence in all severities of hemophilia A69,70 

MutationRelative incidence, %Inhibitor prevalence, %
Large deletions 3.0 41 
    Multidomain  88 
    Single domain  25 
Nonsense mutations 9.3 31 
    Light chain  40 
    Heavy chain  17 
Intron-22 inversion 35.7 21 
Small deletions 10.2 16 
Missense 38.2 
    C1/C2 domain  10 
    Non-C1/C2 domain  
Splice site 2.4 17 
MutationRelative incidence, %Inhibitor prevalence, %
Large deletions 3.0 41 
    Multidomain  88 
    Single domain  25 
Nonsense mutations 9.3 31 
    Light chain  40 
    Heavy chain  17 
Intron-22 inversion 35.7 21 
Small deletions 10.2 16 
Missense 38.2 
    C1/C2 domain  10 
    Non-C1/C2 domain  
Splice site 2.4 17 

Relative incidence is given for each mutation type and subtype, along with the prevalence of inhibitor development in patients with that mutation type or subtype.

The type of fVIII mutation may also influence the titer of inhibitor. Oldenburg et al12  found that 68.8% of those with large deletions had high-titer inhibitors compared with only 21.2% with missense mutations, and 30% to 40% with all other mutation types.

Despite the strong influence of fVIII genotype on inhibitor development, it is not adequately predictive for clinical purposes.13-15  Studies investigating the role of major histocompatibility complex (MHC) class II alleles in inhibitor development have suggested a weak association (Table 3).16-20  Astermark et al21-23  evaluated the effect of polymorphisms in immune response genes on inhibitor development (Table 3). The association of these polymorphisms with inhibitor formation strongly suggests that, in addition to a lack of self-tolerance to fVIII, individual variation in the immune response to foreign antigen influences the risk of developing an inhibitor in patients with severe HA. The interplay between molecular defect and immune response genes has been discussed.24 

Table 3

Non-fVIII genes associated with inhibitor formation

GeneReferenceSevere hemophilia A OR (95% CI)All hemophilia A OR (95% CI)
DQA0102 20  2.7 (1.2-5.9)  
IL10 allele 134 22  5.4 (2.1-9.5) 4.4 (2.1-13.7) 
TNF-a −308 A/A genotype 21  19.2 (2.4-156.5) 4.0 (2.1-13.7) 
CTLA4 −318 T allele 23  0.3 (0.1-0.8)  
GeneReferenceSevere hemophilia A OR (95% CI)All hemophilia A OR (95% CI)
DQA0102 20  2.7 (1.2-5.9)  
IL10 allele 134 22  5.4 (2.1-9.5) 4.4 (2.1-13.7) 
TNF-a −308 A/A genotype 21  19.2 (2.4-156.5) 4.0 (2.1-13.7) 
CTLA4 −318 T allele 23  0.3 (0.1-0.8)  

Treatment-related risk factors

Although the influence of both fVIII and immune response genes is compelling, treatment-related variables may also play a role, as evidenced by the discordance in inhibitor development between monozygotic twins.14 

In 3 small cohort studies, the rate of inhibitor development was greater in those who received their first fVIII infusion before 6 months of age.25-27  However, in a larger cohort, the effect of age disappeared after adjustment for confounding variables.28  Thus, age at first infusion is likely a surrogate for severity of disease leading to the requirement for early intensive therapy. Accordingly, necessary treatment should not be altered to avoid fVIII infusions at a young age.

It has been proposed that inhibitor development can be influenced by the circumstances in which fVIII is used (Table 4). In a cohort of previously untreated patients (PUPs), 65% of those in which surgery was the first indication for fVIII developed an inhibitor compared with approximately 23% in those with other indications for first treatment.15  In those who received 5 or more consecutive days of fVIII at the time of their first exposure, 56% developed an inhibitor, compared with 19% in the group that received fewer than 3 consecutive days of fVIII. However, this was not adjusted for the indication for treatment. In contrast, those who received regular prophylaxis (at least once weekly) had a reduced risk of inhibitor development.

Table 4

Treatment-related risk factors for inhibitor development

Risk factorRelative risk95% CI
Age at first infusion < 6 months of age vs > 12 or 18 months 1.828  0.7-4.7 
 1.727  1.3-1.9 
Surgery at first infusion vs treatment of a bleed at first infusion > 5 days of treatment at first infusion vs < 2 days 2.628  1.3-5.1 
 3.328  2.1-5.3 
Prophylaxis vs no prophylaxis 0.428  0.2-0.8 
 0.2*71  0.06-0.5 
Plasma-derived product vs recombinant product 2.427  1.0-5.8 
 0.832  0.5-1.3 
Risk factorRelative risk95% CI
Age at first infusion < 6 months of age vs > 12 or 18 months 1.828  0.7-4.7 
 1.727  1.3-1.9 
Surgery at first infusion vs treatment of a bleed at first infusion > 5 days of treatment at first infusion vs < 2 days 2.628  1.3-5.1 
 3.328  2.1-5.3 
Prophylaxis vs no prophylaxis 0.428  0.2-0.8 
 0.2*71  0.06-0.5 
Plasma-derived product vs recombinant product 2.427  1.0-5.8 
 0.832  0.5-1.3 
*

Odds ratio.

The differential influence of the indication for fVIII concentrates (prophylaxis vs surgery) may reflect how the environment can influence antigen presentation to T cells. Injury or inflammation at the time of fVIII exposure has been hypothesized to send “danger” signals. In the danger model, distressed cells send alarm signals that activate antigen-presenting cells (APCs), further amplifying immunologic responses.29  Although the danger model may apply to the overall result of the CANAL study,15  approximately 20% of subjects still developed an inhibitor in the absence of circumstances that could be associated with these danger signals, whereas still others did not develop an inhibitor despite the presence of danger signals.

Some have questioned whether the method of delivery of fVIII concentrates influences inhibitor formation. In a retrospective study of patients with mild HA who had received 6 or more consecutive days of factor, inhibitors developed more frequently in patients receiving continuous infusions compared with bolus injections (57% vs 0%).30  In another retrospective study, after approximately 250 cumulative episodes of treatment with fVIII by continuous infusion, 10 patients developed an inhibitor, 5 of which had nonsevere disease.31  Although these studies are interesting, given their study designs it is difficult to base practice on their findings. At this time, we continue to prefer continuous infusion for major surgery because it avoids trough fVIII levels that may place the patient at immediate risk of bleeding. Furthermore, desmopressin should be used when appropriate in those patients with mild HA who have a confirmed response after desmopressin challenge to limit exogenous fVIII exposure.

It has also been suggested that the type of product may influence inhibitor development (Table 4). In a retrospective cohort of 148 PUPs with severe HA, after adjustment for mutation, ethnicity, family history, and age at first infusion, recombinant fVIII concentrates appeared to convey a small risk of inhibitor formation.27  However, there was no adjustment for the impact of prophylactic therapy or surgery at the time of first infusion, which may further influence the strength of the association. In a cohort of 316 subjects, Gouw et al found that there was no association between the type of product used and inhibitor development.32  In addition, since the rate of nontransient inhibitor development in prospective trials of new recombinant products was similar to studies using plasma-derived products33,34 ; and inhibitor formation is rare in previously treated patients who are switched to recombinant products33,35,36 ; we believe the overall weight of the data suggests that there is no association between the type of fVIII product and inhibitor formation. Thus, we do not consider the risk of inhibitor formation when selecting a fVIII product to use for an individual patient.

These studies designed to associate treatment conditions with inhibitor development are interesting, but until the genetic factors that underlie inhibitor development are better understood and can be used to properly stratify patients, the association between how fVIII is delivered and inhibitor formation will be hard to define.

Treatment of acute bleeding

When an inhibitor is first detected, if it is low titer, patients may continue to respond to fVIII replacement with minimal change in the fVIII dose. Such low-titer inhibitors can be observed, as some will resolve spontaneously. Above approximately 5 BU/mL, an inhibitor renders fVIII replacement ineffective and treatment of bleeding episodes requires “bypassing” the deficient clotting factor. Currently available agents include recombinant activated factor VII (rfVIIa; Novoseven; NovoNordisk, Bagsvaerd, Denmark), and FEIBA VH (Baxter, Deerfield, IL). rfVIIa is produced using baby hamster kidney (BHK) cells expressing the cloned human factor VII gene. A new formulation of rfVIIa, Novoseven RT, contains sucrose and L-methionine to allow extended storage at room temperature before reconstitution. rfVIIa facilitates hemostasis by activating factor X directly on the platelet surface thereby bypassing the tenase complex.37  The half-life is 2.3 hours in adults but potentially shorter in children.38  Attempts to protein engineer rfVIIa to have a longer circulation time are currently underway.39,40  FEIBA VH is a vapor-heated concentrate of plasma-derived vitamin K–dependent clotting factors (factors II, VII, IX, and X and others) in both zymogen and active forms. The mechanism of action is multifactorial, though prothrombin and factor X are thought to be critical components.41  rfVIIa and FEIBA VH have similar efficacy and rates of thrombosis; however, in a prospective randomized comparison of rfVIIa and FEIBA VH, approximately 30% of subjects responded more favorably to one product or the other 6 and 12 hours after treatment.42  To better tailor treatment in individual patients some have looked to thromboelastography and endogenous thrombin potential. Unfortunately, clinical studies linking these tests with clinical outcomes are lacking7,43 ; therefore, treatment with rfVIIa and FEIBA VH must be adjusted according to clinical outcomes rather than laboratory testing results.

Since rfVIIa is a recombinant product and has no potential for anamnesis to fVIII (small amounts of fVIII can be found in FEIBA VH),44  we favor the initial use of rfVIIa as a bypassing agent for acute management of bleeding episodes in patients with inhibitors that no longer respond to fVIII. For typical joint bleeds we begin treatment with standard doses, 90 to 120 mcg/kg rounded up to the nearest vial size, given every 2 to 3 hours. This approach in a prospective clinical investigation was effective in 92% of treated bleeds after a mean of 2.2 injections.45  Target joints are more difficult to treat; therefore, based on a randomized trial that demonstrated equivalent efficacy and safety of a single dose of 270 mcg/kg and 3 doses of 90 mcg/kg,46  it is reasonable to use 270 mcg/kg for treatment of target joint bleeds. In addition, a single high dose may be preferred to standard dosing in patients with poor venous access. But since a single treatment may be effective in up to 40% of patients when initiated early and for bleeding in nontarget joints, we do not use the high dose in all patients.47  In the setting of limb or life-threatening bleeding, we begin treatment using at least 120 mcg/kg every 2 hours. Higher initial doses (up to 300 mcg/kg rfVIIa) have also been used in this setting with no untoward effect and should be considered if an adequate clinical response is not achieved using 120 mcg/kg. Once hemostasis is achieved, treatment can be tapered by extending the interval between doses or reducing the dose if high doses are used.

When patients fail to respond to rfVIIa, FEIBA VH should be tried as a single agent.42  FEIBA VH can be used at doses of 50 to 100 U/kg given every 8 to 12 hours, but should not exceed 200 U/kg per day. Lower doses (50-75 U/kg) are used for routine joint bleeds, whereas higher doses (100 U/kg) are given for severe limb or life-threatening bleeding. The use of alternating rfVIIa and FEIBA VH in patients refractory to either alone has been reported to be effective,48  but should be done under careful supervision with attention paid to dose, frequency, and thrombin activation because of the risk for thrombosis.

Alternative approaches for acute bleeding management include porcine fVIII, high-dose human fVIII, and antibody removal by immunoadsorption or plasmapheresis followed by fVIII infusion. Since inhibitors have variable and limited cross reactivity with porcine fVIII, it can be used as a replacement clotting factor. Currently, porcine fVIII is not available in the United States; however recombinant porcine fVIII is in development. High-dose fVIII can also overcome inhibitors in those with a titer less than 5 BU/mL, but may lead to anamnesis in those with a high responding inhibitor. However, under life-threatening circumstances, the benefit of a therapeutic fVIII level, although short-lived, outweighs the subsequent risk of anamnesis. Dosing algorithms in this setting have little scientific basis and have not been validated. In the absence of a rational and validated approach, we use the following formula to estimate the amount of fVIII needed as a loading dose to neutralize the inhibitor [body weight (kg) × 80 × [(1-hematocrit) × antibody titer (BU/mL)] and add an additional 50 IU/kg above the calculated loading dose to achieve a measurable fVIII activity. fVIII levels should be measured 15 minutes after completion of the bolus, and adjustment to reach target levels is necessary because there is substantial individual variation.

As with those with HA without an inhibitor, management of bleeding requires a multidisciplinary approach. After joint or muscle bleeding, physical therapy to facilitate maintenance of range of motion and strength is imperative.

Perioperative management of hemophiliacs with inhibitors

Because of an inability to reliably achieve and monitor hemostasis, surgery in patients with hemophilia A complicated by a high-responding inhibitor should be undertaken with caution. Although there are no comparative clinical studies, both rfVIIa and FEIBA VH can be used for the management of hemostasis in the surgical setting. When choosing which product to use for an individual patient, the product that leads to the best treatment response for acute bleeding is the product to use at the time of surgery. rFVIIa can be used either as continuous infusion or bolus injection. In clinical studies, no differences have been found between these 2 approaches, though the studies were small and underpowered.49  In the absence of clinical data to guide decision making, we prefer to use bolus dosing as it is our opinion that the burst of thrombin generation achieved with bolus dosing is important for hemostasis. Similarly to treatment of severe bleeding, 90 to 120 mcg/kg rfVIIa is given every 2 hours for the first 48 hours. Treatment is then tapered by increasing the interval between doses to complete a course of treatment. As with management of HA patients without an inhibitor, longer durations are needed for major surgery, whereas short durations (1-3 days) are adequate for invasive procedures and minor surgery. If FEIBA VH is used, major surgery requires higher doses (200 U/kg per day) for the first 2 to 4 days which are then tapered over the subsequent days to complete a course of treatment. For minor surgical procedures, such as placement of central venous access device, less intensive therapy (150 U/kg per day) can be used and treatment limited to 3 days.50  Disseminated intravascular coagulation (DIC) has been reported to occur at doses greater than 200 U/kg per day and with more extended treatment courses. In these instances, fibrinogen, and d-dimer should be monitored to detect the onset of DIC.

Prevention of bleeding in hemophiliacs with inhibitors

The benefit of prophylactic therapy in hemophiliacs without an inhibitor has led many to consider prophylactic infusions of bypassing agents in hemophiliacs with an inhibitor.51  A clinical trial compared the frequency of joint bleeds during a pretreatment observation period and treatment period of 3 months during which patients were randomized to rfVIIa 90 mcg/kg per day or 270 mcg/kg per day.52  Both regimens led to a reduction in bleeding frequency and improvements in health-related quality of life.52,53  The ongoing ProFEIBA study is a randomized crossover evaluation of the use of FEIBA VH, 85 U/kg 3 times per week, over a 6-month period. Although bypassing agents may reduce bleeding frequency leading to fewer missed days from work and improved quality of life, whether prophylaxis can improve joint health or reduce the rate of joint deterioration in inhibitor patients is unknown and will likely require additional clinical trials. Factors to consider when deciding whether to use rfVIIa or FEIBA VH prophylaxis includes: frequency of infusions, volume of infusion, cost, and anamnestic response. We prefer to use rfVIIa for prophylaxis in those that are planning to undergo immune tolerance induction (ITI) to avoid the small risk of anamnesis and FEIBA prophylaxis in those that are currently on ITI, are not planning on starting ITI, or have failed ITI to limit the number of infusions. Regardless of the product used, the frequency and dose should be adjusted to find a regimen that is practical, financially feasible, and effective. When to start prophylaxis is subjective since what is considered frequent bleeding will vary significantly among patients. In general, when bleeding is perceived to be interfering with the patient's activities and quality of life, prophylaxis should be considered.

Radionucleotide synovectomy (RNS) is an alternative or adjunctive approach to prophylaxis in the setting of recurrent joint hemorrhage. RNS should be considered in patients with recurrent hemorrhage in a target joint that has evidence of synovial hypertrophy, ideally before significant bone or cartilage damage has been done.

Immune tolerance can be achieved in approximately 70% of patients who receive regular and prolonged infusions of fVIII with or without immune modulation.54,55  Despite the development of multiple ITI protocols since its inception in the 1970s (Table 5), the mechanism of tolerance induction and the best means to achieve tolerance remains unknown. Proposed mechanisms of tolerance development include clonal deletion, anergy or ignorance, induction of suppressor T cells and synthesis of anti-idiotype antibodies.56,57 

Table 5

Immune-tolerance induction protocols

Bonn protocol72 Malmo protocol63 Van Creveld73 
fVIII 100 U/kg BID Immunoadsorption using protein A column if inhibitor titer >10 BU/mL Factor VIII 25-50 IU/kg 
FEIBA 100 U/kg BID  BID for 1-2 weeks, then 25 IU/kg every other day 
 Cyclophosphamide 12-15 mg/kg IV daily × 2 days then 2-3 mg/kg PO daily × 8-10 days  
 FVIII is given to achieve a 40%-100% fVIII level followed by fVIII infusion every 8-12 hours to achieve 30%-80% level  
 IVIG 2.5-5 g IV immediately after the first fVIII infusion followed by 0.4 g/kg daily days 4-8  
Bonn protocol72 Malmo protocol63 Van Creveld73 
fVIII 100 U/kg BID Immunoadsorption using protein A column if inhibitor titer >10 BU/mL Factor VIII 25-50 IU/kg 
FEIBA 100 U/kg BID  BID for 1-2 weeks, then 25 IU/kg every other day 
 Cyclophosphamide 12-15 mg/kg IV daily × 2 days then 2-3 mg/kg PO daily × 8-10 days  
 FVIII is given to achieve a 40%-100% fVIII level followed by fVIII infusion every 8-12 hours to achieve 30%-80% level  
 IVIG 2.5-5 g IV immediately after the first fVIII infusion followed by 0.4 g/kg daily days 4-8  

Data on parameters influencing the success of ITI has been gained from single institutions using a standard approach or from registries. In both the North American Immune Tolerance Registry (NAITR) and the International Immune Tolerance Registry (IITR), the pretreatment inhibitor titer (< 10 BU/mL) and the maximum historical titer (< 200 BU/mL) predicted successful ITI.54,55  In contrast to inhibitor titer, the 2 registries have found conflicting importance of daily dose. In the IITR, patients receiving more than 200 IU/kg per day had the most favorable outcome whereas the NAITR found an inverse correlation between fVIII dose and success rate. However, lower doses required a longer duration to achieve tolerance. Accordingly, the optimal dosing scheme of fVIII for ITI is unclear. Currently, there is an ongoing International Immune Tolerance Trial comparing 200 IU/kg per day with 50 IU/kg 3 times per week in children with HA and high titer inhibitor (> 5 BU/mL).58  Outside of a clinical trial, we favor using higher doses of fVIII (100 IU/kg per day) to achieve tolerance in the shortest possible time. However, in patients who are reluctant to do daily venipuncture and wish to avoid placement of a central venous access device, 3 times per week treatment can be considered and may be successful.

The type of fVIII product to use during ITI is also debatable. Some have observed a higher success rate when fVIII products containing VWF are used.59,60  Some patients who did not respond to ITI with high-purity or recombinant fVIII products have subsequently responded to intermediate purity fVIII products containing VWF.61  However, this observation has not been studied in a prospective, controlled fashion. Given the lack of compelling evidence for one product type over another, we use the product that the patient was using at the time of inhibitor development.

It is our practice to consider ITI in all patients with a newly diagnosed inhibitor and in adults with a long-standing inhibitor that have not previously received ITI. The latter is particularly important in the adult patient if a surgical procedure is necessary, has developed a serious bleed necessitating the use of fVIII, or has frequent bleeding with a marginal response to bypass therapy. Although tolerance is less likely to be achieved in patients with a longstanding inhibitor or with a historical inhibitor titer more than 200 BU/mL, we do not consider these exclusion criteria. In patients with a newly diagnosed inhibitor in which the inhibitor titer is greater than 10 BU/mL before starting ITI, we use bypassing agents (preferably rfVIIa) until the inhibitor is less than 10 BU/mL.44  Ideally, once the inhibitor titer is less than 10 BU/mL, ITI is initiated without delay. However, because of the high degree of commitment required, not all patients and families are suitable to begin ITI at the time a nontransient inhibitor is diagnosed. In these cases, education accompanied by planned initiation for a later date, preferably within 5 years, is an appropriate alternative approach. Successful tolerance is defined as a titer of less than 0.6 BU/mL, a recovery greater than 66% of normal, and a half-life of fVIII of more than 6 hours.62  When to consider someone an ITI failure is difficult and needs to be assessed individually. Some physicians suggest that the likelihood of success is not clinically meaningful if tolerance has not been achieved after 2 years of ITI. In the Immune Tolerance Trial, failure is defined as a lack of a 20% decrease in the inhibitor titer over a 6-month period or a lack of tolerance by 33 months. Although both of these definitions can be helpful in understanding when failure may occur on a population basis, individual patient failure should be determined within the context of that patient's clinical course. We favor continuing ITI if a patient is continuing to make progress and tolerating therapy, even if there is less than a 20% decrease over a 6-month period. In addition, we favor continuing ITI in patients who achieve a detectable fVIII level and/or a favorable clinical response (decreased bleeding frequency) despite a persistently positive inhibitor titer or abnormal recovery.

Since the development of alloantibodies depends on the immune system, it has been postulated that modulation of the immune system can improve response rates to ITI. Immune modulation, although part of the Malmo protocol,63  is not routinely used in other ITI protocols. Early approaches include intravenous immunoglobulin, cyclophosphamide, corticosteroids, and immunoadsorption. In the NAITR, 40% of patients received at least one of these methods of immune modulation without impact on outcome.54  More recently, the combination of rituximab (Rituxan) and ITI has been reported to be successful in several patients who had previously failed ITI alone.64-68  A prospective investigation of rituximab in HA inhibitor patients is ongoing. We do not use immune modulation in ITI.

The development of inhibitory antibodies in patients with HA remains a major complication of therapy. Important areas for ongoing research include (1) improving our understanding of why some develop inhibitors to facilitate better risk assessment; (2) developing alternative factor products with reduced immunogenicity or other therapeutic modalities to prevent inhibitor formation that can be used in patients at high risk for inhibitor development; and (3) improving our understanding of factors necessary for successful immune tolerance therapy, specifically which patients will and will not benefit from ITI and which treatment regimen will provide the highest success rate. With better understanding of the factors involved in the immune response to fVIII, inhibitor development in HA will become predictable and avoidable and more targeted approaches to inhibitor treatment will be feasible.

We thank Drs Thomas Abshire, Donald Harvey, and Pete Lollar for their thoughtful review of this manuscript.

Contribution: C.L.K. and G.C.W. wrote the paper and are responsible for its content, style, and composition.

Conflict-of-interest disclosure: C.L.K. has consulted for Biomea-sure Inc. G.C.W. is a member of the Data Safety Monitoring Board (Baxter), Global Advisory Board (Wyeth, Bayer), Scientific Advisory Board (Entegrion, Stem Cell Products), and Board of Directors (Blood Center of Wisconsin, GTI Milwaukee, WI).

Correspondence: Christine L. Kempton, 2015 Uppergate Drive, Suite 400, Atlanta, GA 30322; e-mail: ckempto@emory.edu.

1
Darby
 
SC
Keeling
 
DM
Spooner
 
RJ
, et al. 
The incidence of factor VIII and factor IX inhibitors in the hemophilia population of the UK and their effect on subsequent mortality, 1977-99.
J Thromb Haemost
2004
, vol. 
2
 (pg. 
1047
-
1054
)
2
Ehrenforth
 
S
Kreuz
 
W
Scharrer
 
I
, et al. 
Incidence of development of factor VIII and factor IX inhibitors in haemophiliacs.
Lancet
1992
, vol. 
339
 (pg. 
594
-
598
)
3
Kempton
 
CL
Soucie
 
JM
Abshire
 
TC
Incidence of inhibitors in a cohort of 838 males with hemophilia A previously treated with factor VIII concentrates.
J Thromb Haemost
2006
, vol. 
4
 (pg. 
2576
-
2581
)
4
Soucie
 
JM
Cianfrini
 
C
Janco
 
RL
, et al. 
Joint range-of-motion limitations among young males with hemophilia: prevalence and risk factors.
Blood
2004
, vol. 
103
 (pg. 
2467
-
2473
)
5
Kasper
 
CK
Aledort
 
L
Aronson
 
D
, et al. 
Proceedings: a more uniform measurement of factor VIII inhibitors.
Thromb Diath Haemorrh
1975
, vol. 
34
 pg. 
612
 
6
White
 
GC
Rosendaal
 
F
Aledort
 
LM
Lusher
 
JM
Rothschild
 
C
Ingerslev
 
J
Definitions in hemophilia. Recommendation of the scientific subcommittee on factor VIII and factor IX of the scientific and standardization committee of the International Society on Thrombosis and Haemostasis.
Thromb Haemost
2001
, vol. 
85
 pg. 
560
 
7
Negrier
 
C
Gomperts
 
ED
Oldenburg
 
J
The history of FEIBA: a lifetime of success in the treatment of haemophilia complicated by an inhibitor.
Haemophilia
2006
, vol. 
12
 (pg. 
4
-
13
)
8
The Recombinate Study Group
Bray
 
GL
Gomperts
 
ED
Courter
 
S
, et al. 
A multicenter study of recombinant factor VIII (Recombinate): safety, efficacy, and inhibitor risk in previously untreated patients with hemophilia A.
Blood
1994
, vol. 
83
 (pg. 
2428
-
2435
)
9
Kogenate Previously Untreated Patient Study Group
Lusher
 
JM
Arkin
 
S
Abildgaard
 
CF
Schwartz
 
RS
Recombinant factor VIII for the treatment of previously untreated patients with hemophilia A. Safety, efficacy, and development of inhibitors.
N Engl J Med
1993
, vol. 
328
 (pg. 
453
-
459
)
10
Recombinant Factor VIII Study Group
Schwartz
 
RS
Abildgaard
 
CF
Aledort
 
LM
, et al. 
Human recombinant DNA-derived antihemophilic factor (factor VIII) in the treatment of hemophilia A.
N Engl J Med
1990
, vol. 
323
 (pg. 
1800
-
1805
)
11
Oldenburg
 
J
Schroder
 
J
Brackmann
 
HH
Muller-Reible
 
C
Schwaab
 
R
Tuddenham
 
E
Environmental and genetic factors influencing inhibitor development.
Semin Hematol
2004
, vol. 
41
 (pg. 
82
-
88
)
12
Oldenburg
 
J
Pavlova
 
A
Genetic risk factors for inhibitors to factors VIII and IX.
Haemophilia
2006
, vol. 
12
 
suppl 6
(pg. 
15
-
22
)
13
Gill
 
JC
The role of genetics in inhibitor formation.
Thromb Haemost
1999
, vol. 
82
 (pg. 
500
-
504
)
14
Astermark
 
J
Berntorp
 
E
White
 
GC
Kroner
 
BL
The Malmo International Brother Study (MIBS): further support for genetic predisposition to inhibitor development in hemophilia patients.
Haemophilia
2001
, vol. 
7
 (pg. 
267
-
272
)
15
Gouw
 
SC
van der Bom
 
JG
Marijke van den Berg
 
H
Treatment-related risk factors of inhibitor development in previously untreated patients with hemophilia A: the CANAL cohort study.
Blood
2007
, vol. 
109
 (pg. 
4648
-
4654
)
16
Frommel
 
D
Muller
 
JY
Prou-Wartelle
 
O
Allain
 
JP
Possible linkage between the major histocompatibility complex and the immune response to factor VIII in classic haemophilia.
Vox Sang
1977
, vol. 
33
 (pg. 
270
-
272
)
17
Frommel
 
D
Allain
 
JP
Saint-Paul
 
E
, et al. 
HLA antigens and factor VIII antibody in classic hemophilia. European study group of factor VIII antibody.
Thromb Haemost
1981
, vol. 
46
 (pg. 
687
-
689
)
18
Lippert
 
LE
Fisher
 
LM
Schook
 
LB
Relationship of major histocompatibility complex class II genes to inhibitor antibody formation in hemophilia A.
Thromb Haemost
1990
, vol. 
64
 (pg. 
564
-
568
)
19
Oldenburg
 
J
Picard
 
JK
Schwaab
 
R
Brackmann
 
HH
Tuddenham
 
EG
Simpson
 
E
HLA genotype of patients with severe haemophilia A due to intron 22 inversion with and without inhibitors of factor VIII.
Thromb Haemost
1997
, vol. 
77
 (pg. 
238
-
242
)
20
Hay
 
CR
Ollier
 
W
Pepper
 
L
, et al. 
HLA class II profile: a weak determinant of factor VIII inhibitor development in severe haemophilia A. UKHCDO Inhibitor Working Party.
Thromb Haemost
1997
, vol. 
77
 (pg. 
234
-
237
)
21
Astermark
 
J
Oldenburg
 
J
Carlson
 
J
, et al. 
Polymorphisms in the TNFA gene and the risk of inhibitor development in patients with hemophilia A.
Blood
2006
, vol. 
108
 (pg. 
3739
-
3745
)
22
Astermark
 
J
Oldenburg
 
J
Pavlova
 
A
Berntorp
 
E
Lefvert
 
AK
Polymorphisms in the IL10 but not in the IL1beta and IL4 genes are associated with inhibitor development in patients with hemophilia A.
Blood
2006
, vol. 
107
 (pg. 
3167
-
3172
)
23
Astermark
 
J
Wang
 
X
Oldenburg
 
J
Berntorp
 
E
Lefvert
 
AK
Polymorphisms in the CTLA-4 gene and inhibitor development in patients with severe hemophilia A.
J Thromb Haemost
2007
, vol. 
5
 (pg. 
263
-
265
)
24
White
 
GC
Kempton
 
CL
Grimsley
 
A
Nielsen
 
B
Roberts
 
HR
Cellular immune responses in hemophilia: why do inhibitors develop in some, but not all hemophiliacs?
J Thromb Haemost
2005
, vol. 
3
 (pg. 
1676
-
1681
)
25
van der Bom
 
JG
Mauser-Bunschoten
 
EP
Fischer
 
K
van den Berg
 
HM
Age at first treatment and immune tolerance to factor VIII in severe hemophilia.
Thromb Haemost
2003
, vol. 
89
 (pg. 
475
-
479
)
26
Lorenzo
 
JI
Lopez
 
A
Altisent
 
C
Aznar
 
JA
Incidence of factor VIII inhibitors in severe haemophilia: the importance of patient age.
Br J Haematol
2001
, vol. 
113
 (pg. 
600
-
603
)
27
Goudemand
 
J
Rothschild
 
C
Demiguel
 
V
, et al. 
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
 (pg. 
46
-
51
)
28
Gouw
 
SC
van den Berg
 
HM
le Cessie
 
S
van der Bom
 
JG
Treatment characteristics and the risk of inhibitor development: a multicenter cohort study among previously untreated patients with severe hemophilia A.
J Thromb Haemost
2007
, vol. 
5
 (pg. 
1383
-
1390
)
29
Matzinger
 
P
The danger model: a renewed sense of self.
Science
2002
, vol. 
296
 (pg. 
301
-
305
)
30
Sharathkumar
 
A
Lillicrap
 
D
Blanchette
 
VS
, et al. 
Intensive exposure to factor VIII is a risk factor for inhibitor development in mild hemophilia A.
J Thromb Haemost
2003
, vol. 
1
 (pg. 
1228
-
1236
)
31
von Auer
 
C
Oldenburg
 
J
von Depka
 
M
, et al. 
Inhibitor development in patients with hemophilia A after continuous infusion of FVIII concentrates.
Ann N Y Acad Sci
2005
, vol. 
1051
 (pg. 
498
-
505
)
32
Gouw
 
SC
van der Bom
 
JG
Auerswald
 
G
Ettinghausen
 
CE
Tedgard
 
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
 (pg. 
4693
-
4697
)
33
Lusher
 
JM
Hemophilia treatment. Factor VIII inhibitors with recombinant products: prospective clinical trials.
Haematologica
2000
, vol. 
85
 (pg. 
2
-
5
discussion 5–6
34
Wight
 
J
Paisley
 
S
The epidemiology of inhibitors in haemophilia A: a systematic review.
Haemophilia
2003
, vol. 
9
 (pg. 
418
-
435
)
35
Gringeri
 
A
Tagliaferri
 
A
Tagariello
 
G
Morfini
 
M
Santagostino
 
E
Mannucci
 
P
Efficacy and inhibitor development in previously treated patients with haemophilia A switched to a B domain-deleted recombinant factor VIII.
Br J Haematol
2004
, vol. 
126
 (pg. 
398
-
404
)
36
Hoots
 
WK
Lusher
 
J
High-titer inhibitor development in hemophilia A: lack of product specificity.
J Thromb Haemost
2004
, vol. 
2
 (pg. 
358
-
359
)
37
Roberts
 
HR
Monroe
 
DM
White
 
GC
The use of recombinant factor VIIa in the treatment of bleeding disorders.
Blood
2004
, vol. 
104
 (pg. 
3858
-
3864
)
38
Lindley
 
CM
Sawyer
 
WT
Macik
 
BG
, et al. 
Pharmacokinetics and pharmacodynamics of recombinant factor VIIa.
Clin Pharmacol Ther
1994
, vol. 
55
 (pg. 
638
-
648
)
39
Yatuv
 
R
Dayan
 
I
Carmel-Goren
 
L
, et al. 
Enhancement of factor VIIa haemostatic efficacy by formulation with PEGylated liposomes.
Haemophilia
2008
, vol. 
14
 (pg. 
476
-
483
)
40
Weimer
 
T
Wormsbacher
 
W
Kronthaler
 
U
Lang
 
W
Liebing
 
U
Schulte
 
S
Prolonged in-vivo half-life of factor VIIa by fusion to albumin.
Thromb Haemost
2008
, vol. 
99
 (pg. 
659
-
667
)
41
Turecek
 
PL
Varadi
 
K
Gritsch
 
H
, et al. 
Factor Xa and prothrombin: mechanism of action of FEIBA.
Vox Sang
1999
, vol. 
77
 
suppl 1
(pg. 
72
-
79
)
42
Astermark
 
J
Donfield
 
SM
DiMichele
 
DM
, et al. 
A randomized comparison of bypassing agents in hemophilia complicated by an inhibitor: the FEIBA NovoSeven Comparative (FENOC) Study.
Blood
2007
, vol. 
109
 (pg. 
546
-
551
)
43
Dargaud
 
Y
Lambert
 
T
Trossaert
 
M
New advances in the therapeutic and laboratory management of patients with haemophilia and inhibitors.
Haemophilia
2008
, vol. 
14
 
suppl 4
(pg. 
20
-
27
)
44
Brackmann
 
HH
Effenberger
 
E
Hess
 
L
Schwaab
 
R
Oldenburg
 
J
NovoSeven in immune tolerance therapy.
Blood Coagul Fibrinolysis
2000
, vol. 
11
 
suppl 1
(pg. 
S39
-
S44
)
45
Key
 
NS
Aledort
 
LM
Beardsley
 
D
, et al. 
Home treatment of mild to moderate bleeding episodes using recombinant factor VIIa (Novoseven) in haemophiliacs with inhibitors.
Thromb Haemost
1998
, vol. 
80
 (pg. 
912
-
918
)
46
Santagostino
 
E
Mancuso
 
ME
Rocino
 
A
Mancuso
 
G
Scaraggi
 
F
Mannucci
 
PM
A prospective randomized trial of high and standard dosages of recombinant factor VIIa for treatment of hemarthroses in hemophiliacs with inhibitors.
J Thromb Haemost
2006
, vol. 
4
 (pg. 
367
-
371
)
47
Santagostino
 
E
Gringeri
 
A
Mannucci
 
PM
Home treatment with recombinant activated factor VII in patients with factor VIII inhibitors: the advantages of early intervention.
Br J Haematol
1999
, vol. 
104
 (pg. 
22
-
26
)
48
Schneiderman
 
J
Nugent
 
DJ
Young
 
G
Sequential therapy with activated prothrombin complex concentrate and recombinant factor VIIa in patients with severe haemophilia and inhibitors.
Haemophilia
2004
, vol. 
10
 (pg. 
347
-
351
)
49
Pruthi
 
RK
Mathew
 
P
Valentino
 
LA
Sumner
 
MJ
Seremetis
 
S
Hoots
 
WK
Haemostatic efficacy and safety of bolus and continuous infusion of recombinant factor VIIa are comparable in haemophilia patients with inhibitors undergoing major surgery. Results from an open-label, randomized, multicenter trial.
Thromb Haemost
2007
, vol. 
98
 (pg. 
726
-
732
)
50
Stine
 
KC
Shrum
 
D
Becton
 
DL
Use of FEIBA for invasive or surgical procedures in patients with severe hemophilia A or B with inhibitors.
J Pediatr Hematol Oncol
2007
, vol. 
29
 (pg. 
216
-
221
)
51
Manco-Johnson
 
MJ
Abshire
 
TC
Shapiro
 
AD
, et al. 
Prophylaxis versus episodic treatment to prevent joint disease in boys with severe hemophilia.
N Engl J Med
2007
, vol. 
357
 (pg. 
535
-
544
)
52
Konkle
 
BA
Ebbesen
 
LS
Erhardtsen
 
E
, et al. 
Randomized, prospective clinical trial of recombinant factor VIIa for secondary prophylaxis in hemophilia patients with inhibitors.
J Thromb Haemost
2007
, vol. 
5
 (pg. 
1904
-
1913
)
53
Hoots
 
WK
Ebbesen
 
LS
Konkle
 
BA
, et al. 
Secondary prophylaxis with recombinant activated factor VII improves health-related quality of life of haemophilia patients with inhibitors.
Haemophilia
2008
, vol. 
14
 (pg. 
466
-
475
)
54
DiMichele
 
DM
Kroner
 
BL
The North American Immune Tolerance Registry: practices, outcomes, outcome predictors.
Thromb Haemost
2002
, vol. 
87
 (pg. 
52
-
57
)
55
Mariani
 
G
Kroner
 
B
Immune tolerance in hemophilia with factor VIII inhibitors: predictors of success.
Haematologica
2001
, vol. 
86
 (pg. 
1186
-
1193
)
56
White
 
GC
Greenwood
 
R
Escobar
 
M
Frelinger
 
JA
Hemophilia factor VIII therapy. Immunological tolerance. A clinical perspective.
Haematologica
2000
, vol. 
85
 (pg. 
113
-
116
)
57
Gilles
 
JG
Desqueper
 
B
Lenk
 
H
Vermylen
 
J
Saint-Remy
 
JM
Neutralizing antiidiotypic antibodies to factor VIII inhibitors after desensitization in patients with hemophilia A.
J Clin Invest
1996
, vol. 
97
 (pg. 
1382
-
1388
)
58
Hay
 
CR
Hemophilia treatment. Immune tolerance induction: prospective clinical trials.
Haematologica
2000
, vol. 
85
 (pg. 
52
-
55
discussion 55–56
59
Kreuz
 
W
Ettingshausen
 
CE
Zyschka
 
A
, et al. 
Inhibitor development in previously untreated patients with hemophilia A: a prospective long-term follow-up comparing plasma-derived and recombinant products.
Semin Thromb Hemost
2002
, vol. 
28
 (pg. 
285
-
290
)
60
Kruez
 
W
Lee
 
C
Immune tolerance and choice of concentrates.
Inhibitors in patients with hemophilia
2002
Oxford
Blackwell Science Ltd
(pg. 
55
-
56
)
61
Kurth
 
MA
Dimichele
 
D
Sexauer
 
C
, et al. 
Immune tolerance therapy utilizing factor VIII/von Willebrand factor concentrate in haemophilia A patients with high titre factor VIII inhibitors.
Haemophilia
2008
, vol. 
14
 (pg. 
50
-
55
)
62
Dimichele
 
D
Immune tolerance therapy for factor VIII inhibitors: moving from empiricism to an evidence-based approach.
J Thromb Haemost
2007
, vol. 
5
 
suppl 1
(pg. 
143
-
150
)
63
Nilsson
 
IM
Berntorp
 
E
Zettervall
 
O
Induction of immune tolerance in patients with hemophilia and antibodies to factor VIII by combined treatment with intravenous IgG, cyclophosphamide, and factor VIII.
N Engl J Med
1988
, vol. 
318
 (pg. 
947
-
950
)
64
Mathias
 
M
Khair
 
K
Hann
 
I
Liesner
 
R
Rituximab in the treatment of alloimmune factor VIII and IX antibodies in two children with severe haemophilia.
Br J Haematol
2004
, vol. 
125
 (pg. 
366
-
368
)
65
Mateo
 
J
Badell
 
I
Forner
 
R
Borrell
 
M
Tizzano
 
E
Fontcuberta
 
J
Successful suppression using Rituximab of a factor VIII inhibitor in a boy with severe congenital haemophilia: an example of a significant decrease of treatment costs.
Thromb Haemost
2006
, vol. 
95
 (pg. 
386
-
387
)
66
Moschovi
 
M
Aronis
 
S
Trimis
 
G
Platokouki
 
H
Salavoura
 
K
Tzortzatou-Stathopoulou
 
F
Rituximab in the treatment of high responding inhibitors in severe haemophilia A.
Haemophilia
2006
, vol. 
12
 (pg. 
95
-
99
)
67
Fox
 
RA
Neufeld
 
EJ
Bennett
 
CM
Rituximab for adolescents with haemophilia and high titre inhibitors.
Haemophilia
2006
, vol. 
12
 (pg. 
218
-
222
)
68
Collins
 
PW
Novel therapies for immune tolerance in haemophilia A.
Haemophilia
2006
, vol. 
12
 
suppl 6
(pg. 
94
-
100
discussion 100–101
69
Oldenburg
 
J
Tuddenham
 
E
Lee
 
C
Berntorp
 
E
Hoots
 
K
Inhibitors to factor VIII-molecular basis.
Textbook of Hemophilia
2005
Malden, MA
Blackwell Publishing
(pg. 
59
-
63
)
70
Goodeve
 
AC
Peake
 
IR
The molecular basis of hemophilia A: genotype-phenotype relationships and inhibitor development.
Semin Thromb Hemost
2003
, vol. 
29
 (pg. 
23
-
30
)
71
Santagostino
 
E
Mancuso
 
ME
Rocino
 
A
, et al. 
Environmental risk factors for inhibitor development in children with haemophilia A: a case-control study.
Br J Haematol
2005
, vol. 
130
 (pg. 
422
-
427
)
72
Brackmann
 
HH
Oldenburg
 
J
Schwaab
 
R
Immune tolerance for the treatment of factor VIII inhibitors–twenty years' ‘bonn protocol’.
Vox Sang
1996
, vol. 
70
 
suppl 1
(pg. 
30
-
35
)
73
Mauser-Bunschoten
 
EP
Nieuwenhuis
 
HK
Roosendaal
 
G
van den Berg
 
HM
Low-dose immune tolerance induction in hemophilia A patients with inhibitors.
Blood
1995
, vol. 
86
 (pg. 
983
-
988
)
Sign in via your Institution