• No results found

Molecular determinants of FVIII immunogenicity in hemophilia A - Chapter 1: General introduction

N/A
N/A
Protected

Academic year: 2021

Share "Molecular determinants of FVIII immunogenicity in hemophilia A - Chapter 1: General introduction"

Copied!
20
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

UvA-DARE is a service provided by the library of the University of Amsterdam (https://dare.uva.nl)

UvA-DARE (Digital Academic Repository)

Molecular determinants of FVIII immunogenicity in hemophilia A

Wróblewska, A.

Publication date

2013

Link to publication

Citation for published version (APA):

Wróblewska, A. (2013). Molecular determinants of FVIII immunogenicity in hemophilia A.

General rights

It is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s)

and/or copyright holder(s), other than for strictly personal, individual use, unless the work is under an open

content license (like Creative Commons).

Disclaimer/Complaints regulations

If you believe that digital publication of certain material infringes any of your rights or (privacy) interests, please

let the Library know, stating your reasons. In case of a legitimate complaint, the Library will make the material

inaccessible and/or remove it from the website. Please Ask the Library: https://uba.uva.nl/en/contact, or a letter

to: Library of the University of Amsterdam, Secretariat, Singel 425, 1012 WP Amsterdam, The Netherlands. You

will be contacted as soon as possible.

(2)

1

(3)
(4)

Gener

al intr

oduc

tion

1

Hemophilia A

Hemophilia A is hereditary X-linked disorder caused by dysfunction or absence

of blood coagulation factor VIII. The bleeding syndrome has been recognized

as early as in the fifth century, but only since 1937 was linked to a deficiency

of a plasma component, named in 1962 “factor VIII”.

1

Depending on the

severity of the disease, it can be divided into three categories: mild (5-25% of

normal

plasma

levels

of FVIII),

moderate (1-5%)

and severe

(<1%)

hemophilia A.

In its severe form it can lead to spontaneous joint and muscle bleeds, which

consequently can cause deformation of joints requiring surgical intervention.

Trauma can cause life-threatening internal bleedings and hemorrhages. Treating

hemophilia became possible in 1964, when Judith Graham Pool, working at

Stanford University, described how to obtain cryoprecipitate from human

plasma.

2

Since then, so-called replacement therapy comprising regular injections

of FVIII of either recombinant or plasma-derived origin became commonly

used in hemophilia care. However, such treatment is hampered by high costs

of the FVIII products – it is estimated that more than 75% of the hemophilia

community

worldwide

receive

either

inadequate

or

no

treatment

whatsoever.

3

In response to frequent FVIII infusions, a subset of patients develops anti-FVIII

antibodies. In severe hemophilia A, inhibitors develop after a median of 10

to 15 days of treatment with FVIII.

4

After 50 exposure days, the cumulative

incidence of inhibitors reaches a plateau, after which the occurrence of

inhibitors is exceptional.

5

Anti-FVIII antibodies can rapidly inhibit FVIII function,

rendering FVIII infusion therapy ineffective. Inhibitory antibodies interfere with

binding of FVIII to other coagulation factors such as factor IIa, IXa and X or to

phospholipids.

6,7

Non-inhibitory antibodies can also compromise hemophilia

treatment by influencing FVIII stability and/or its pharmacokinetics by interfering

with binding to von Willebrand factor (VWF).

8,9

The majority of inhibitory

antibodies directed towards FVIII are of subclass IgG1 and IgG4.

10-12

Factor VIII: structure and function

FVIII is synthesized as a polypeptide chain comprising a signal peptide of 19

amino acids and a mature protein of 2332 residues. FVIII consists of three A, two

C domains and one unique B domain that are arranged in the following order:

A1-a1-A2-a2-B-a3-A3-C1-C2 (Figure 1).

13,14

The a1 (residues 337-372), a2 (residues

711-740) and a3 (residues 1649-1689) regions bordering the A domains are

enriched for aspartic and glutamic acid residues.

13

In plasma, FVIII circulates as a

hetero-dimer consisting of a 90-220 kDa heavy chain that is non-covalently linked

to a 80 kDa light chain (Figure 1).

15

Cleavage at Arg1648 by as yet unidentified

protease releases the light chain (a3-A3-C1-C2); additional processing at various

positions (including Arg740 and Arg1312) within the B domain yields a heavy

chain (A1-a1-A2-a2-B) that is heterogeneous in size. In the circulation, FVIII is

tightly associated with its carrier protein – von Willebrand factor (VWF), which

protects it from proteolytic degradation and premature clearance.

16-20

Sulfation

of Tyr1680 within the acidic a3 region is required for high affinity binding of

FVIII to VWF.

21-23

Also residues in the C1 and C2 domain have been implicated

(5)

1

Gener

al intr

oduc

tion

in binding of FVIII to VWF.

16,21-25

FVIII circulates as an inactive precursor that

can only act as a cofactor for the FIXa-dependent conversion of FX following its

activation by thrombin or FXa.

26

Cleavage by thrombin at Arg1689 releases the

acidic a3 region which promotes rapid disassembly of the FVIII-VWF complex

27,28

;

cleavage at Arg372 and Arg740 results in a hetero-trimeric molecule comprising

the A3-C1-C2, A1-a1 and A2-a2 domains.

27

The resulting hetero-trimer can

efficiently catalyze the conversion of factor X to Xa by factor IXa on phospholipid

surfaces.

14

Dissociation of the A2 domain from hetero-trimeric FVIII results in

a rapid decline of FVIII cofactor activity.

29-31

Also cleavage of FVIII by activated

protein C at positions Arg336 and Arg562 abolishes cofactor activity of activated

FVIII.

32,33

Residues 1811-1818 in the A3 domain and residues 558-565 in the

A2 domain have been shown to contribute to the binding of factor IXa.

34-37

Limited information is available with respect to the binding site for factor X on

activated FVIII; binding sites for factor X in the C2 domain and the a1 region have

been proposed.

38,39

It is now well-established that both the C1 and C2 domain

contribute to binding of FVIII to negatively charged phospholipids.

40-44

Analysis of

the three-dimensional structure of the C2 domain and site-directed mutagenesis

have firmly implicated a role for Met2199, Phe2200, Leu2251 and Leu2252 in

the binding of FVIII to negatively charged phospholipids.

45-48

More recent

studies have identified an exposed surface loop in the C1 domain harbouring

Arg2090, Gln2091, Lys2092 and Phe2093 that modulates binding of FVIII to

surfaces containing a low percentage of negatively charged,

phosphatidylserine-containing phospholipids.

42,43

Figure 1 provides an overview of interactive sites

on FVIII for its major binding partners.

FVIII precursor form

FVIII circulating in plasma

Heavy chain Light chain

Me2+ A1 A2 A3 C1 C2

Figure 1. Structure of blood coagulation factor VIII (FVIII). Upper left panel shows

schematic domain organization of FVIII; lower left panel – schematic representation of circulating FVIII, where the heavy (A1-A2-B) and the light (A3-C1-C2) chain are non-covalently linked via a metal ion-dependent interaction between the A1 and A3 domain. In activated FVIII, the A1 and A3 domains remain non-covalently bound, while the A2 domain is weakly associated with the A1/A3-C1-C2 dimer. Right panel provides a crystal structure of B domain-deleted FVIII (pdb code 3cdz). Interactive sites for phospholipids, VWF, FIXa and FX are indicated in the model.

(6)

Gener

al intr

oduc

tion

1

Epitope mapping studies have revealed that inhibitory antibodies that develop

in patients with hemophilia A bind to the A2 and C2 domain of FVIII.

7,49-52

More

detailed binding studies have shown that residues 484-508 provide a major binding

site for anti-A2 domain antibodies.

52

The mode of action of anti-A2 inhibitors is not

yet entirely clear; an early study by Lollar and co-workers suggested that anti-A2

antibodies act as non-competitive inhibitors of intrinsic Factor X activation complex

by blocking the conversion of FXase/FX complex to the transition state.

53

Follow-up

studies using isolated A2 domain suggested that anti-A2 antibodies directly inhibit

the interaction of A2 subunit with factor IXa, thus abrogating the contribution of this

subunit to cofactor activity.

54

Anti-C2 domain antibodies have been shown to bind to

exposed residues overlapping with the phospholipid binding site.

55

Co-crystalization

of a human monoclonal anti-C2 domain antibody, BO2C11, revealed that Arg2215,

Arg2220, Met2199, Phe2200, Val2223, Leu2251 and Leu2252 comprise contact

residues for this inhibitor.

56

Inhibitory antibodies directed towards residues

1811-1818 in the A3 domain have also been detected in patients with hemophilia A.

34,57

Antibodies binding to these sites have been shown to limit the binding of FIXa to

FVIII.

34,57

Moreover, in a small number of patients antibodies directed towards the

acidic a1-region have been identified.

24,41,58

Replacement therapy and inhibitor development in hemophilia A

The development of inhibitors occurs in approximately 5% of mild or moderate

hemophilia A patients, and in 25% of severe hemophilia A patients.

59

The low prevalence of inhibitor development in patients with mild and moderate

hemophilia A is most likely caused by the presence of endogenous levels of

circulating FVIII which render this group of patients tolerant to subsequent

replacement therapy.

60

The lack of endogenous levels of circulating FVIII most

likely underlies the increased frequency of FVIII inhibitors in patients with

severe hemophilia A. A large number of studies have addressed the correlation

between FVIII genotype and inhibitor development.

5,61,62

Patients carrying intron

22 inversions, nonsense mutations or large deletions are more prone to develop

inhibitors than those with small deletions and missense mutations.

63,64

Nonsense

mutations that affect the light chain of FVIII are more frequently associated with

inhibitor development than those present in the FVIII heavy chain.

65,66

Formation

of inhibitors in mild hemophilia associates with a limited number of high-risk

FVIII mutations within the A2, C1 or C2 domain, namely Arg593Cys, Arg2150His

or Trp2229Cys.

67,68

Why only a fraction of patients generate antibodies against

FVIII is still poorly understood. Both treatment-related and genetic risk factors

have been shown to contribute to inhibitor development in hemophilia A.

Intensity of FVIII treatment is an acknowledged risk factor, as FVIII administered

to treat bleeding episodes or to support surgery delivers high concentrations

of FVIII that could promote antibody development.

69-72

Apart from FVIII gene

mutations

73,74

, polymorphisms within the IL-10

75

and TNFA gene

76

have been

associated with inhibitor development. Interestingly, a C/T polymorphism in the

promoter region of the CTLA-4 gene

77

was overrepresented in hemophilia A

(7)

1

Gener

al intr

oduc

tion

formation. Several studies have explored whether inhibitor development is linked

to type of FVIII product. There are conflicting findings about the immunogenicity

of recombinant versus plasma derived FVIII.

78-81

Results from a large

multi-center cohort study suggest that the risk of inhibitor formation is similar for

recombinant and plasma derived FVIII products.

78-81

Also the von Willebrand

factor content of therapeutic FVIII products seems not to be associated with the

risk of inhibitor development.

81

FVIII and the immune system

It is now well-established that the formation of high-affinity IgG molecules

requires FVIII-specific CD4

+

T-cell help.

10,82

Initial evidence for involvement of T

cells in the development of inhibitors in hemophilia A patients came from studies

involving HIV-positive individuals, that due to diminished number of CD4

+

T cells

showed also decline in anti-FVIII antibody responses.

83

The first step leading

to activation of specific T- and B-cell responses is recognition of antigen by

specialized antigen-presenting cells (APCs). Subsequently, antigen is processed

into small peptides which are loaded on MHC class II. Presentation of

derived peptides on the surface of APCs triggers the activation of

antigen-specific CD4

+

T cells (Figure 2). In the past years, several studies have addressed

how FVIII is processed by APCs.

84-87

Although a role for several receptors has

been suggested (for a review see Chapter 2)

84,85

, the exact mechanism of FVIII

endocytosis by APCs remains unclear. APCs, equipped with receptors recognizing

foreign, pathogen-derived molecules (so-called pathogen-recognition receptors,

PRRs), are able to sense “danger”, which prompts maturation of these cells.

Upon maturation, they upregulate a number of co-stimulatory molecules such

as CD40, CTLA-4, CD80 and CD86 and due to their simultaneous ability to release

cytokines APCs can activate and modulate antigen-specific T cell responses

(Figure 2). The importance of co-stimulatory mechanism for FVIII-specific T-cell

activation has been illustrated by several studies on the blockade of CD40/CD40L

interactions in vivo.

88,89

CD40/CD40L ligation provides a key event to induce

humoral responses against antigens

90

; furthermore, blockage of CD40/CD40L

interaction leads to long-lasting tolerance in mice.

91,92

However, even though

the disruption of CD40/CD40L interaction by pre-administration of a monoclonal

antibody targeting CD40L resulted in deficient immune responses against FVIII

in vivo, it failed to induce long-lasting tolerance.

89

Clinical trials suggested that

administration of a humanized anti-CD40 ligand antibody (hu5c8) can block

anamnestic responses to factor VIII; however it remained unclear whether that

effect would persist and result in long-lasting tolerance.

93

Moreover, due to

tromboembolic complications in patients treated with hu5c8, this approach for

treatment has been discontinued.

During TCR activation in a particular cytokine milieu, naive CD4

+

T cells may

differentiate into one of several lineages of T helper (Th) cells including Th1, Th2,

Th17 and regulatory T cells (Tregs), as defined by their pattern of cytokine production

and function. Generally, Th1 cells promote cellular immunity; Th2 cells mediate

humoral immunity; Th17 cells play an important role in clearing pathogens during

(8)

Gener

al intr

oduc

tion

1

host defense reactions and in inducing tissue inflammation in autoimmune disease

94

; Tregs are capable of inducing tolerance by suppressing T- and B-cell responses.

T-cell reponses against FVIII are of a polyclonal origin and directed against multiple

epitopes present in different domains of FVIII.

95

Th1

10

, Th2

10

as well as Th17

96

cells

have been shown to contribute to FVIII-specific immune responses in hemophilia

subjects. Eventually, FVIII-specific T cells are able to activate FVIII-specific B cells

and subsequently induce affinity maturation and class-switching of immunoglobulin

genes. As a consequence, FVIII-specific long-living plasma cells and memory B cells

are generated, which are able to rapidly respond to re-exposure to FVIII.

97,98

FVIII-specific B cell responses are also, similar to T-cell responses, of a polyclonal origin

and directed against various epitopes. Determination of classes and subclasses of

anti-FVIII antibodies revealed dominant contribution of IgGs and their subclasses

– IgG1, IgG2 and IgG4.

8,10,12

Generated antibodies can inhibit FVIII function by

interfering with its interaction with other coagulation factors (mainly A2 and A3-C1

antibodies) and/or phospholipids (C1/C2-directed antibodies), as described earlier.

It has been reported, that a subset of anti-factor VIII IgG hydrolyzes FVIII. These

so-called catalytic antibodies were found in over 50% of inhibitor-positive patients with

severe hemophilia A, but not in inhibitor-negative individuals.

100

FVIII-hydrolyzing IgG

from each patient tested exhibit multiple cleavage sites on FVIII and the specificity

of cleavage varied from one patient to another.

100

Circulating antibodies that do not

inhibit FVIII function can influence FVIII half-life, either by interfering with binding

of FVIII to VWF or by formation of immune complexes that can be efficiently cleared

via Fc receptors.

Antigen Dendritic cell T helper cell B cell Plasma cells Cytokines MHC II TCR CD28 B7.2 Antibodies

Figure 2. Simplified overview of the development of humoral immune responses. Antigen is endocytosed by

antigen-presenting cell (such as dendritic cell), processed and presented on MHC class II. For efficient activation, co-stimulatory molecules present on APCs and T cell receptors need to interact, while cytokines released by APCs determine the future direction of ongoing T-cell responses. Once primed, T cells can activate B cells in an antigen-specific manner, which leads to formation of long-living plasma cells producing high-affinity antibodies.

(9)

1

Gener

al intr

oduc

tion

Immune tolerance induction (ITI) therapy – eradication of inhibitors

Development of inhibitors is a serious complication in hemophilia care. The

magnitude of antibody responses is quantified by a functional assay and

expressed in Bethesda units (BU). Patients with low titer of inhibitors (< 5 BU/

ml) can be treated with higher and/or more frequent doses of FVIII, which leads

to saturation of pre-existing inhibitors and still provides enough FVIII to restore

hemostasis and normal coagulation. For patients with higher inhibitor titers

(>5 BU/ml), simple infusion therapy becomes ineffective, therefore bypassing

agents such as activated factor VII or activated prothrombin-complex

concentrates need to be used.

101,102

Simultaneously, eradication of inhibitors and

immune tolerance induction (ITI) therapy is introduced. ITI comprises regular

injections of FVIII for a period varying from several weeks up to two years and

since its first description in 1977

103

remains the only strategy that proved to both

eradicate FVIII inhibitors as well as lead to induction of FVIII-specific immune

tolerance. The most commonly used protocols, known as the Bonn

104

, the Van

Creveld

105

and the Malmö

106-108

protocols, although considerably different

65

,

result in comparable success rates (up to 87%). A recent study showed that

high-dose ITI leads to faster recovery and tolerance induction, accompanied by fewer

bleeding episodes as compared to low-dose ITI.

109

Van Helden and co-workers

showed a correlation between distribution of IgG-subclasses of anti-FVIII

antibodies and outcome of ITI therapy. A predominance of IgG4 antibodies was

observed in patients who needed prolonged ITI treatment.

11

Also complications,

such as venous catheter infection, can prolong the course of ITI or even lead to its

failure.

110

The immunological mechanisms underlying success of ITI therapy

remain unclear. In a naïve, non-primed environment, chronic exposure to high

doses of antigen would activate regulatory T cells able to suppress

antigen-specific (in this case FVIII-antigen-specific) effector T cells, resulting in tolerance

induction.

111,112

Consequently, no T cell help would be provided to FVIII-specific

B cells that could not differentiate into long-living antibody producing plasma

cells and, as a result, would be depleted. This scenario is however unlikely to

happen in patients with pre-existing antibodies that have enriched populations

of FVIII-specific memory T and B cells. In this case, memory B cells serve as highly

efficient antigen-presenting cells, able to effectively re-stimulate FVIII-specific

memory T cells. Moreover, upon encounter with antigen, memory B cells rapidly

differentiate into specific antibody producing plasma cells, enriching already

pre-existing pool of such cells. However, the amount of antigen seems to be

crucial factor for an optimal stimulation of memory B cells.

113

Studies using

FVIII-deficient mice showed that concentrations of FVIII below the physiologic plasma

concentration of 0.1 μg/ml (1 U/mL) restimulate FVIII-specific memory B cells

and induce their differentiation into antibody-secreting cells (ASCs).

113

However,

high concentrations (above 0.1 μg/ml), inhibit memory B cell restimulation and

prevent the formation of ASCs.

113

The inhibition of FVIII-specific memory B-cell

responses seems to be irreversible and not mediated by FVIII-specific T cells. Such

depletion of memory B-cell can be an early event in the inhibitors eradication

in patients undergoing ITI therapy, who receive high doses of FVIII. Gilles and

(10)

Gener

al intr

oduc

tion

1

colleagues

114

followed two patients with FVIII inhibitors during their course of

ITI and suggested that induction of anti-idiotypic antibodies, neutralizing the

existing anti-FVIII antibodies, might be one of the reasons underlying success of

ITI.

Animal models of hemophilia A

In

1995

Bi

and

co-workers

115,116

introduced

two

mouse

models

for

hemophilia A,

which over the years proved to be very useful for broadening our knowledge

on induction and modulation of immune responses against FVIII. Targeted

disruption of exon 16 (E16-KO mice) or exon 17 (E17-KO) resulted in the absence

of functional FVIII in the circulation (<1% normal FVIII activity). Both strains

displayed a bleeding tendency

117

, which could be corrected by infusions with

human FVIII.

118

Similar to hemophilia A patients, FVIII-deficient mice develop

an immune response following repetitive intravenous injections of therapeutic

doses of FVIII.

113,119,120

Generated antibodies are directed both to the light as

well

as

to

the

heavy

chain

of

FVIII

120

,

persist

in

the

circulation

for

a

long

time

98

and

their development is strictly dependent on CD4

+

T helper cells.

121,122

Cytokine

profiles of factor VIII-specific T cells indicate that the regulation of the

anti-FVIII antibody response in hemophilic mice involves both Th1- and Th2-type

cells.

92,122

Over the years, studies involving FVIII-deficient mice enabled us to

gain more knowledge regarding modulation of immune response to FVIII and

tolerance induction. Furthermore, it allowed for evaluation of a number of

novel therapeutic approaches to prevent or eradicate inhibitor development in

hemophilia A. Administration of complexes of FVIII and phosphatidylserine (PS)

liposomes resulted in reduced antibody formation against FVIII in hemophilia A

mice.

123

PS associated with apoptotic cells is known to induce anti-inflammatory

responses in APCs

124

, moreover, PS liposomes reduce the maturation,

pro-inflammatory cytokine production and T cell priming of APCs.

124

Formation of

complex between FVIII and phospholipid molecules is mediated by residues

located in the C1 and C2 domains of FVIII.

42,44

Therefore, a possible alternative

explanation of the observed inhibitory effect is that PS occupies residues of FVIII

that are crucial for its endocytosis by APCs. CTLA4-IgG, blocking the co-stimulatory

interaction between B7 and CD28, has been shown to transiently inhibit

anti-FVIII antibody formation in hemophilia A mice.

88

An independent study revealed

that simultaneous blockage of CD40-CD40L and B7-CD28 pathways abolishes

development of inhibitors and promotes long-term immune tolerance specific

for FVIII.

125

CD3 antibody has been studied as a tolerance-inducing agent for

several autoimmune and inflammatory diseases.

126-128

It modulates the

CD3–T-cell receptor (TCR) complex and leads to anergy or apoptosis of effector T CD3–T-cells,

or to the expansion of regulatory CD4

+

CD25

+

T cells.

128

Consequently, anti-CD3

treatment proved to successfully prevent inhibitor formation in hemophilia A

mice with both BALB/c and C57BL/6 background.

129,130

Rituximab, a therapeutic

anti-CD20 antibody, has been used for several years to treat patients with

inhibitors, although with various success rates.

131,132

In hemophilia A mice,

(11)

1

Gener

al intr

oduc

tion

titers in FVIII-challenged animals.

133

However, hyporesponsiveness to FVIII was

sustained over time only when an anti-CD20 antibody of subclass IgG1 was used,

which, in contrast to the other isotype tested – IgG2a, did not deplete marginal

zone B cells.

133

Moreover, treatment with anti-CD20 IgG1, but not IgG2a, was

accompanied by substantial increase of splenic regulatory T cells, implicating a

possible tolerogenic role for the remaining B cell population.

133

A recent study

by Sack and colleagues, where anti-CD20 treatment was combined with

AAV-based gene therapy, showed that B cell depletion can render FVIII-deficient mice

hyporesponsive to FVIII, but results were dependent on the genetic background

of the strain used.

134

Transduction with retroviral constructs encoding FVIII

A2- or C2-domain in frame with an IgG heavy chain backbone of B cell blasts

has been shown to induce immune tolerance to FVIII in hemophilic mice.

135

Mice treated with transduced B cells showed reduced inhibitor titers, which

were persistently low after additional challenges with FVIII. Furthermore, the

lower antibody titers correlated with an increased frequency of FVIII-specific

regulatory T cells.

135,136

Factor VIII deficient mice have also been successfully

used to study the presence and persistence of memory B cells, as in detailed

described in the previous paragraph.

113

Bril and colleagues

137

generated transgenic mice expressing human FVIII with the

Arg593 to Cys mutation, which is associated with mild hemophilia phenotype.

68,138,139

Unlike E16-KO or E17-KO mice, these animals did not develop inhibitor titers upon

repetitive intravenous injections of human FVIII, however induction of an immune

response took place after subcutaneous FVIII administration in the presence of a

strong adjuvant.

137

A similar model was described by van Helden and colleagues who

constructed a transgenic mouse expressing full length human FVIII.

140

In accordance

with findings by Bril and co-workers no inhibitor formation was observed in this model

following the intravenous administration of human FVIII. Interestingly, infusion of

PEGylated FVIII evoked an immune response in transgenic mice expressing human

FVIII. These results suggest that transgenic mouse models that express human

FVIII are useful models for assessing the potential immunogenicity of genetically or

chemically modified FVIII variants. More recently, the development of a humanized

hemophilic E17 HLA-DRB1*1501 mouse model has been described.

89,141

Humanized

mice have been utilized to study the regulation of HLA class II-restricted immune

responses to various antigens and they proved to be highly suitable for in vivo

research into the mechanistic basis of human diseases associated with activation of

CD4

+

T cells.

142

HLA-DRB1*1501 was selected due to a strong connection between

this haplotype and many immunologic diseases

143

, as well as a previously noted

link

between

inhibitor

incidence

and

DRB1*1501

in

patients

with

severe

hemophilia

A.

144

Despite some obvious limitations, such new models can be used to analyze the

(12)

Gener

al intr

oduc

tion

1

Scope of the thesis

Understanding the recognition of FVIII by antigen-presenting cells and subsequent

activation of specific T and B cells is essential for development of new strategies

for treatment and/or prevention of inhibitor formation to FVIII in hemophilia

A patients. In this thesis we explored the mechanism of FVIII endocytosis by

both human and mouse dendritic cells (Chapter 3 and Chapter 6). A search for

structural determinants important for immune recognition of FVIII revealed that

C1 domain-targeting monoclonal antibody KM33 prevents the uptake of FVIII

by APCs and delays the formation of anti-FVIII antibodies in hemophilia A mice.

Chapter 4 follows up on this finding and shows that modification of C1 domain

residues crucial for KM33 binding diminishes FVIII uptake by dendritic cells.

Upon in vivo administration in FVIII-deficient mice these C1 domain variants

developed significantly lower anti-FVIII antibody titers and reduced CD4

+

T cell

responses. In Chapter 5 we address the role of immune complex formation in

FVIII endocytosis by APCs and its influence on subsequent FVIII-specific T cell

responses. Chapter 6 elaborates in more detail on the possible mechanism

of FVIII endocytosis by human dendritic cells and the potential role of the C2

domain in this process. Together our findings provide more insight into immune

recognition of FVIII that can be utilized to develop novel strategies for treatment

or prevention of inhibitor formation in hemophilia A.

(13)

1

Gener

al intr

oduc

tion

References

1. Ingram GI. The history of haemophilia. Journal of Clinical Pathology. 1976;29(6):469-479. 2. Pool JG, Gershgold EJ, Pappenhagen AR. High-Potency Antihaemophilic Factor Concentrate Prepared from Cryoglobulin Precipitate. Nature. 1964;203:312.

3. Skinner MW. Treatment for all: a vision for the future. Haemophilia. 2006;3:169-173.

4. Wight J, Paisley S. The epidemiology of inhibitors in haemophilia A: a systematic review. Haemophilia. 2003;9(4):418-435.

5. 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;109(11):4648-4654.

6. Ananyeva NM, Lacroix-Desmazes S, Hauser CA, et al. Inhibitors in hemophilia A: mechanisms of inhibition, management and perspectives. Blood Coagul Fibrinolysis. 2004;15(2):109-124.

7. Scandella D. New characteristics of anti-factor VIII inhibitor antibody epitopes and unusual immune responses to Factor VIII. Semin Thromb Hemost. 2002;28(3):291-296.

8. Gilles JG, Arnout J, Vermylen J, Saint-Remy JM. Anti-factor VIII antibodies of hemophiliac patients are frequently directed towards nonfunctional determinants and do not exhibit isotypic restriction. Blood. 1993;82(8):2452-2461.

9. Dazzi F, Tison T, Vianello F, et al. High incidence of anti-FVIII antibodies against non-coagulant epitopes in haemophilia A patients: a possible role for the half-life of transfused FVIII. Br J Haematol. 1996;93(3):688-693.

10. Reding MT, Lei S, Lei H, Green D, Gill J, Conti-Fine BM. Distribution of Th1- and Th2-induced anti-factor VIII IgG subclasses in congenital and acquired hemophilia patients. Thromb Haemost. 2002;88(4):568-575.

11. van Helden PM, van den Berg HM, Gouw SC, et al. IgG subclasses of anti-FVIII antibodies during immune tolerance induction in patients with hemophilia A. Br J Haematol. 2008;142(4):644-652.

12. Fulcher CA, de Graaf MS, Zimmerman TS. FVIII inhibitor IgG subclass and FVIII polypeptide specificity determined by immunoblotting. Blood. 1987;69(5):1475-1480.

13. Vehar GA, Keyt B, Eaton D, et al. Structure of human factor VIII. Nature. 1984;312(5992): 337-342.

14. Lenting PJ, van Mourik JA, Mertens K. The life cycle of coagulation factor VIII in view of its structure and function. Blood. 1998;92(11):3983-3996.

15. Thompson AR. Structure and function of the factor VIII gene and protein. Semin Thromb Hemost. 2003;29(1):11-22.

16. Leyte A, Mertens K, Distel B, et al. Inhibition of human coagulation factor VIII by monoclonal antibodies. Mapping of functional epitopes with the use of recombinant factor VIII fragments. Biochem J. 1989;263(1):187-194.

17. Ganz PR, Atkins JS, Palmer DS, Dudani AK, Hashemi S, Luison F. Definition of the affinity of binding between human von Willebrand factor and coagulation factor VIII. Biochem Biophys Res Commun. 1991;180(1):231-237.

18. Vlot AJ, Koppelman SJ, van den Berg MH, Bouma BN, Sixma JJ. The affinity and stoichiometry of binding of human factor VIII to von Willebrand factor. Blood. 1995;85(11):3150-3157.

19. Brinkhous KM, Sandberg H, Garris JB, et al. Purified human factor VIII procoagulant protein: comparative hemostatic response after infusions into hemophilic and von Willebrand disease dogs. Proc Natl Acad Sci USA. 1985;82(24):8752-8756.

20. Andersson LO, Brown JE. Interaction of factor VIII-von Willebrand Factor with phospholipid vesicles. Biochem J. 1981;200(1):161-167.

21. Lollar P, Hill-Eubanks DC, Parker CG. Association of the factor VIII light chain with von Willebrand factor. J Biol Chem. 1988;263(21):10451-10455.

(14)

Gener

al intr

oduc

tion

1

fragments on kinetics of interaction between the light and heavy chains of human factor VIII. Thromb Res. 1999;96(5):343-354.

23. Hamer RJ, Koedam JA, Beeser-Visser NH, Bertina RM, Van Mourik JA, Sixma JJ. Factor VIII binds to von Willebrand factor via its Mr-80,000 light chain. Eur J Biochem. 1987;166(1):37-43.

24. Foster PA, Fulcher CA, Houghten RA, Zimmerman TS. An immunogenic region within residues Val1670-Glu1684 of the factor VIII light chain induces antibodies which inhibit binding of factor VIII to von Willebrand factor. J Biol Chem. 1988;263(11):5230-5234.

25. 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;266(2):740-746.

26. Regan LM, Fay PJ. Cleavage of factor VIII light chain is required for maximal generation of factor VIIIa activity. J Biol Chem. 1995;270(15):8546-8552.

27. Eaton D, Rodriguez H, Vehar GA. Proteolytic processing of human factor VIII. Correlation of specific cleavages by thrombin, factor Xa, and activated protein C with activation and inactivation of factor VIII coagulant activity. Biochemistry. 1986;25(2):505-512.

28. Donath MS, Lenting PJ, Van Mourik JA, Mertens K. The role of cleavage of the light chain at positions Arg1689 or Arg1721 in subunit interaction and activation of human blood coagulation factor VIII. J Biol Chem. 1995;270(8):3648-3655.

29. Lollar P, Parker CG. pH-dependent denaturation of thrombin-activated porcine factor VIII. J Biol Chem. 1990;265(3):1688-1692.

30. Lollar P, Parker ET. Structural basis for the decreased procoagulant activity of human factor VIII compared to the porcine homolog. J Biol Chem. 1991;266(19):12481-12486.

31. Pipe SW, Eickhorst AN, McKinley SH, Saenko EL, Kaufman RJ. Mild hemophilia A caused by increased rate of factor VIII A2 subunit dissociation: evidence for nonproteolytic inactivation of factor VIIIa in vivo. Blood. 1999;93(1):176-183.

32. Koedam JA, Meijers JC, Sixma JJ, Bouma BN. Inactivation of human factor VIII by activated protein C. Cofactor activity of protein S and protective effect of von Willebrand factor. J Clin Invest. 1988;82(4):1236-1243.

33. Fay PJ, Smudzin TM, Walker FJ. Activated protein C-catalyzed inactivation of human factor VIII and factor VIIIa. Identification of cleavage sites and correlation of proteolysis with cofactor activity. J Biol Chem. 1991;266(30):20139-20145.

34. Zhong D, Saenko EL, Shima M, Felch M, Scandella D. Some human inhibitor antibodies interfere with factor VIII binding to factor IX. Blood. 1998;92(1):136-142.

35. Fay PJ, Beattie T, Huggins CF, Regan LM. Factor VIIIa A2 subunit residues 558-565 represent a factor IXa interactive site. J Biol Chem. 1994;269(32):20522-20527.

36. Lenting PJ, van de Loo JW, Donath MJ, van Mourik JA, Mertens K. The sequence Glu1811-Lys1818 of human blood coagulation factor VIII comprises a binding site for activated factor IX. J Biol Chem. 1996;271(4):1935-1940.

37. Lenting PJ, Donath MJ, Van Mourik JA, Mertens K. Identification of a binding site for blood coagulation factor IXa on the light chain of human factor VIII. J Biol Chem. 1994;269(10): 7150-7155.

38. Lapan KA, Fay PJ. Localization of a factor X interactive site in the A1 subunit of factor VIIIa. J Biol Chem. 1997;272(4):2082-2088.

39. Nogami K, Shima M, Hosokawa K, et al. Role of Factor VIII C2 Domain in Factor VIII Binding to Factor Xa. J Biol Chem. 1999;274(43):31000-31007.

40. Shima M, Scandella D, Yoshioka A, et al. A factor VIII neutralizing monoclonal antibody and a human inhibitor alloantibody recognizing epitopes in the C2 domain inhibit factor VIII binding to von Willebrand factor and to phosphatidylserine. Thromb Haemost. 1993;69(3):240-246.

41. Foster PA, Fulcher CA, Houghten RA, Zimmerman TS. Synthetic factor VIII peptides with amino acid sequences contained within the C2 domain of factor VIII inhibit factor VIII binding to

(15)

1

Gener

al intr

oduc

tion

phosphatidylserine. Blood. 1990;75(10):1999-2004.

42. Meems H, Meijer AB, Cullinan DB, Mertens K, Gilbert GE. Factor VIII C1 domain residues Lys 2092 and Phe 2093 contribute to membrane binding and cofactor activity. Blood. 2009;114(18): 3938-3946.

43. Lu J, Pipe SW, Miao H, Jacquemin M, Gilbert GE. A membrane-interactive surface on the factor VIII C1 domain cooperates with the C2 domain for cofactor function. Blood. 2011;117(11): 3181-3189.

44. Novakovic VA, Cullinan DB, Wakabayashi H, Fay PJ, Baleja JD, Gilbert GE. Membrane-binding properties of the Factor VIII C2 domain. Biochem J. 2011;435(1):187-196.

45. 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;402(6760):439-442.

46. Gilbert GE, Kaufman RJ, Arena AA, Miao H, Pipe SW. Four hydrophobic amino acids of the factor VIII C2 domain are constituents of both the membrane-binding and von Willebrand factor-binding motifs. J Biol Chem. 2002;277(8):6374-6381.

47. Spiegel PC, Murphy P, Stoddard BL. Surface-exposed hemophilic mutations across the factor VIII C2 domain have variable effects on stability and binding activities. J Biol Chem. 2004;279(51):53691-53698.

48. Gilbert GE, Novakovic VA, Kaufman RJ, Miao H, Pipe SW. Conservative mutations in the C2 domains of factor VIII and factor V alter phospholipid binding and cofactor activity. Blood. 2012;120(9):1923-1932.

49. Scandella D, Mattingly M, Prescott R. A recombinant factor VIII A2 domain polypeptide quantitatively neutralizes human inhibitor antibodies that bind to A2. Blood. 1993;82(6): 1767-1775.

50. Scandella D, Gilbert GE, Shima M, et al. Some factor VIII inhibitor antibodies recognize a common epitope corresponding to C2 domain amino acids 2248 through 2312, which overlap a phospholipid-binding site. Blood. 1995;86(5):1811-1819.

51. Prescott R, Nakai H, Saenko EL, et al. The inhibitor antibody response is more complex in hemophilia A patients than in most nonhemophiliacs with factor VIII autoantibodies. Recombinate and Kogenate Study Groups. Blood. 1997;89(10):3663-3671.

52. Lubin IM, Healey JF, Barrow RT, Scandella D, Lollar P. Analysis of the human factor VIII A2 inhibitor epitope by alanine scanning mutagenesis. J Biol Chem. 1997;272(48):30191-30195.

53. Lollar P, Parker ET, Curtis JE, et al. Inhibition of human factor VIIIa by anti-A2 subunit antibodies. J Clin Invest. 1994;93(6):2497-2504.

54. Fay PJ, Scandella D. Human inhibitor antibodies specific for the factor VIII A2 domain disrupt the interaction between the subunit and factor IXa. J Biol Chem. 1999;274(42):29826-29830.

55. Healey JF, Barrow RT, Tamim HM, et al. Residues Glu2181-Val2243 contain a major determinant of the inhibitory epitope in the C2 domain of human factor VIII. Blood. 1998;92(10):3701-3709.

56. Spiegel PC, Jr., Jacquemin M, Saint-Remy JM, Stoddard BL, Pratt KP. Structure of a factor VIII C2 domain-immunoglobulin G4kappa Fab complex: identification of an inhibitory antibody epitope on the surface of factor VIII. Blood. 2001;98(1):13-19.

57. Fijnvandraat K, Celie PH, Turenhout EA, et al. A human alloantibody interferes with binding of factor IXa to the factor VIII light chain. Blood. 1998;91(7):2347-2352.

58. Scandella DH. Properties of anti-factor VIII inhibitor antibodies in hemophilia A patients. Semin Thromb Hemost. 2000;26(2):137-142.

59. Mannucci PM, Tuddenham EG. The hemophilias ─ from royal genes to gene therapy. N Engl J Med. 2001;344(23):1773-1779.

60. Giuffrida AC, Genesini S, Franchini M, De Gironcoli M, Aprili G, Gandini G. Inhibitors in mild/ moderate haemophilia A: two case reports and a literature review. Blood Transfus. 2008;6(3): 163-168.

(16)

Mauser-Gener

al intr

oduc

tion

1

Bunschoten EP. Influence of the type of F8 gene mutation on inhibitor development in a single centre cohort of severe haemophilia A patients. Haemophilia. 2011;17(2):275-281.

62. Gouw SC, van den Berg HM, Oldenburg J, et al. F8 gene mutation type and inhibitor development in patients with severe hemophilia A: systematic review and meta-analysis. Blood. 2012;119(12):2922-2934.

63. Oldenburg J, Pavlova A. Genetic risk factors for inhibitors to factors VIII and IX. Haemophilia. 2006;12 Suppl 6:15-22.

64. Goodeve AC, Peake IR. The molecular basis of hemophilia A: genotype-phenotype relationships and inhibitor development. Semin Thromb Hemost. 2003;29(1):23-30.

65. van Helden PM, Kaijen PH, Fijnvandraat K, van den Berg HM, Voorberg J. Factor VIII-specific memory B cells in patients with hemophilia A. J Thromb Haemost. 2007;5(11):2306-2308.

66. Oldenburg J, Schroder J, Brackmann HH, Muller-Reible C, Schwaab R, Tuddenham E. Environmental and genetic factors influencing inhibitor development. Semin Hematol. 2004;41(1 Suppl 1):82-88.

67. Hay CR. Factor VIII inhibitors in mild and moderate-severity haemophilia A. Haemophilia. 1998;4(4):558-563.

68. Bril WS, MacLean PE, Kaijen PH, et al. HLA class II genotype and factor VIII inhibitors in mild haemophilia A patients with an Arg593 to Cys mutation. Haemophilia. 2004;10(5):509-514.

69. 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;5(7):1383-1390.

70. Ragni MV, Ojeifo O, Feng J, et al. Risk factors for inhibitor formation in haemophilia: a prevalent case-control study. Haemophilia. 2009;15(5):1074-1082.

71. 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;1(6):1228-1236.

72. 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;7(3):267-272.

73. Schwaab R, Brackmann HH, Meyer C, et al. Haemophilia A: mutation type determines risk of inhibitor formation. Thromb Haemost. 1995;74(6):1402-1406.

74. Tuddenham EG, McVey JH. The genetic basis of inhibitor development in haemophilia A. Haemophilia. 1998;4(4):543-545.

75. 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;107(8):3167-3172.

76. 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;108(12):3739-3745.

77. 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;5(2):263-265.

78. 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;5(3):145-154.

79. Peerlinck K, Arnout J, Di Giambattista M, et al. Factor VIII inhibitors in previously treated haemophilia A patients with a double virus-inactivated plasma derived factor VIII concentrate. Thromb Haemost. 1997;77(1):80-86.

80. 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;109(11):4693-4697.

(17)

1

Gener

al intr

oduc

tion

81. Gouw SC, van der Bom JG, Ljung R, et al. Factor VIII products and inhibitor development in severe hemophilia A. N Engl J Med. 2013;368(3):231-239.

82. Ragni MV, Bontempo FA, Lewis JH. Disappearance of inhibitor to factor VIII in HIV-infected hemophiliacs with progression to AIDS or severe ARC. Transfusion. 1989;29(5):447-449.

83. Qian J, Collins M, Sharpe AH, Hoyer LW. Prevention and treatment of factor VIII inhibitors in murine hemophilia A. Blood. 2000;95(4):1324-1329.

84. Dasgupta S, Navarrete AM, Bayry J, et al. A role for exposed mannosylations in presentation of human therapeutic self-proteins to CD4+ T lymphocytes. Proc Natl Acad Sci USA. 2007;104(21):

8965-8970.

85. Dasgupta S, Navarrete AM, Andre S, et al. Factor VIII bypasses CD91/LRP for endocytosis by dendritic cells leading to T-cell activation. Haematologica. 2008;93(1):83-89.

86. Herczenik E, van Haren SD, Wroblewska A, et al. Uptake of blood coagulation factor VIII by dendritic cells is mediated via its C1 domain. Journal of Allergy and Clinical Immunology. 2012;129(2):501-509, 509.e1-5.

87. van Haren SD, Wroblewska A, Fischer K, Voorberg J, Herczenik E. Requirements for immune recognition and processing of factor VIII by antigen-presenting cells. Blood Rev. 2012;26(1):43-49.

88. Qian J, Burkly LC, Smith EP, et al. Role of CD154 in the secondary immune response: the reduction of pre-existing splenic germinal centers and anti-factor VIII inhibitor titer. Eur J Immunol. 2000;30(9):2548-2554.

89. Reipert BM, Steinitz KN, van Helden PM, et al. Opportunities and limitations of mouse models humanized for HLA class II antigens. J Thromb Haemost. 2009;7 Suppl 1:92-97.

90. Foy TM, Aruffo A, Bajorath J, Buhlmann JE, Noelle RJ. Immune regulation by CD40 and its ligand GP39. Annu Rev Immunol. 1996;14:591-617.

91. Honey K, Cobbold SP, Waldmann H. CD40 ligand blockade induces CD4+ T cell tolerance and

linked suppression. J Immunol. 1999;163(9):4805-4810.

92. Rossi G, Sarkar J, Scandella D. Long-term induction of immune tolerance after blockade of CD40-CD40L interaction in a mouse model of hemophilia A. Blood. 2001;97(9):2750-2757.

93. Ewenstein BM, Hoots WK, Lusher JM, et al. Inhibition of CD40 ligand (CD154) in the treatment of factor VIII inhibitors. Haematologica. 2000;85(10 Suppl):35-39.

94. Korn T, Bettelli E, Oukka M, Kuchroo VK. IL-17 and Th17 Cells. Annu Rev Immunol. 2009;27: 485-517.

95. Wroblewska A, Reipert BM, Pratt KP, Voorberg J. Dangerous liaisons: how the immune system deals with factor VIII. J Thromb Haemost. 2013;11(1):47-55.

96. Ettinger RA, James EA, Kwok WW, Thompson AR, Pratt KP. Lineages of human T-cell clones, including T helper 17/T helper 1 cells, isolated at different stages of anti-factor VIII immune responses. Blood. 2009;114(7):1423-1428.

97. Van Helden P, Kaijen PH, Mauser-Bunschoten EP, Fischer K, HM VDB, Voorberg J. Domain specificity of factor VIII inhibitors during immune tolerance induction in patients with haemophilia A. Haemophilia. 2010;16(6):892-901.

98. Hausl C, Maier E, Schwarz HP, et al. Long-term persistence of anti-factor VIII antibody-secreting cells in hemophilic mice after treatment with human factor VIII. Thromb Haemost. 2002;87(5):840-845.

99. Lacroix-Desmazes S, Moreau A, Sooryanarayana, et al. Catalytic activity of antibodies against factor VIII in patients with hemophilia A. Nat Med. 1999;5(9):1044-1047.

100. Lacroix-Desmazes S, Wootla B, Dasgupta S, et al. Catalytic IgG from patients with hemophilia A inactivate therapeutic factor VIII. J Immunol. 2006;177(2):1355-1363.

101. von Depka M. Managing acute bleeds in the patient with haemophilia and inhibitors: options, efficacy and safety. Haemophilia. 2005;11 Suppl 1:18-23.

(18)

Gener

al intr

oduc

tion

1

agents in hemophilia complicated by an inhibitor: the FEIBA NovoSeven Comparative (FENOC) Study. Blood. 2007;109(2):546-551.

103. Brackmann HH, Gormsen J. Massive factor-VIII infusion in haemophiliac with factor-VIII inhibitor, high responder. Lancet. 1977;2(8044):933.

104. Brackmann HH, Oldenburg J, Schwaab R. Immune Tolerance for the Treatment of Factor VIII Inhibitors - Twenty Years’ ‘Bonn Protocol’. Vox Sanguinis. 1996;70:30-35.

105. Mauser-Bunschoten EP, Nieuwenhuis HK, Roosendaal G, van den Berg HM. Low-dose immune tolerance induction in hemophilia A patients with inhibitors. Blood. 1995;86(3):983-988.

106. Carlborg E, Astermark J, Lethagen S, Ljung R, Berntorp E. The Malmo model for immune tolerance induction: impact of previous treatment on outcome. Haemophilia. 2000;6(6):639-642.

107. Berntorp E, Astermark J, Carlborg E. Immune tolerance induction and the treatment of hemophilia. Malmo protocol update. Haematologica. 2000;85(10 Suppl):48-50.

108. 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;318(15):947-950.

109. Hay CR, DiMichele DM. The principal results of the International Immune Tolerance Study: a randomized dose comparison. Blood. 2012;119(6):1335-1344.

110. Hay C, Recht M, Carcao M, Reipert B. Current and future approaches to inhibitor management and aversion. SeminThrombHemost. 2006;32 Suppl 2:15-21.

111. Apostolou I, von Boehmer H. In vivo instruction of suppressor commitment in naive T cells. J Exp Med. 2004;199(10):1401-1408.

112. Chen TC, Cobbold SP, Fairchild PJ, Waldmann H. Generation of anergic and regulatory T cells following prolonged exposure to a harmless antigen. J Immunol. 2004;172(10):5900-5907.

113. Hausl C, Ahmad RU, Sasgary M, et al. High-dose factor VIII inhibits factor VIII-specific memory B cells in hemophilia A with factor VIII inhibitors. Blood. 2005;106(10):3415-3422.

114. 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;97(6):1382-1388.

115. Bi L, Lawler AM, Antonarakis SE, High KA, Gearhart JD, Kazazian HH, Jr. Targeted disruption of the mouse factor VIII gene produces a model of haemophilia A. Nat Genet. 1995;10(1):119-121.

116. Bi L, Sarkar R, Naas T, et al. Further characterization of factor VIII-deficient mice created by gene targeting: RNA and protein studies. Blood. 1996;88(9):3446-3450.

117. Muchitsch EM, Turecek PL, Zimmermann K, et al. Phenotypic expression of murine hemophilia. Thromb Haemost. 1999;82(4):1371-1373.

118. Connelly S, Andrews JL, Gallo AM, et al. Sustained phenotypic correction of murine hemophilia A by in vivo gene therapy. Blood. 1998;91(9):3273-3281.

119. Qian J, Borovok M, Bi L, Kazazian HH, Jr., Hoyer LW. Inhibitor antibody development and T cell response to human factor VIII in murine hemophilia A. Thromb Haemost. 1999;81(2):240-244.

120. Reipert BM, Ahmad RU, Turecek PL, Schwarz HP. Characterization of antibodies induced by human factor VIII in a murine knockout model of hemophilia A. Thromb Haemost. 2000;84(5): 826-832.

121. Wu H, Reding M, Qian J, et al. Mechanism of the immune response to human factor VIII in murine hemophilia A. Thromb Haemost. 2001;85(1):125-133.

122. Sasgary M, Ahmad RU, Schwarz HP, Turecek PL, Reipert BM. Single cell analysis of factor VIII-specific T cells in hemophilic mice after treatment with human factor VIII. Thromb Haemost. 2002;87(2):266-272.

123. Purohit VS, Ramani K, Sarkar R, Kazazian HH, Jr., Balasubramanian SV. Lower inhibitor development in hemophilia A mice following administration of recombinant factor VIII-O-phospho-L-serine complex. J Biol Chem. 2005;280(18):17593-17600.

(19)

1

Gener

al intr

oduc

tion

124. Hoffmann PR, Kench JA, Vondracek A, et al. Interaction between phosphatidylserine and the phosphatidylserine receptor inhibits immune responses in vivo. J Immunol. 2005;174(3): 1393-1404.

125. Miao CH, Ye P, Thompson AR, Rawlings DJ, Ochs HD. Immunomodulation of transgene responses following naked DNA transfer of human factor VIII into hemophilia A mice. Blood. 2006;108(1):19-27.

126. Keymeulen B, Vandemeulebroucke E, Ziegler AG, et al. Insulin needs after CD3-antibody therapy in new-onset type 1 diabetes. N Engl J Med. 2005;352(25):2598-2608.

127. Belghith M, Bluestone JA, Barriot S, Megret J, Bach JF, Chatenoud L. TGF-beta-dependent mechanisms mediate restoration of self-tolerance induced by antibodies to CD3 in overt autoimmune diabetes. Nat Med. 2003;9(9):1202-1208.

128. Chatenoud L, Bluestone JA. CD3-specific antibodies: a portal to the treatment of autoimmunity. Nat Rev Immunol. 2007;7(8):622-632.

129. Waters B, Qadura M, Burnett E, et al. Anti-CD3 prevents factor VIII inhibitor development in hemophilia A mice by a regulatory CD4+CD25+-dependent mechanism and by shifting cytokine

production to favor a Th1 response. Blood. 2009;113(1):193-203.

130. Peng B, Ye P, Rawlings DJ, Ochs HD, Miao CH. Anti-CD3 antibodies modulate anti-factor VIII immune responses in hemophilia A mice after factor VIII plasmid-mediated gene therapy. Blood. 2009; 114(20):4373-82.

131. Aleem A, Saidu A, Abdulkarim H, et al. Rituximab as a single agent in the management of adult patients with haemophilia A and inhibitors: marked reduction in inhibitor level and clinical improvement in bleeding but failure to eradicate the inhibitor. Haemophilia. 2009;15(1):210-216.

132. Collins PW, Mathias M, Hanley J, et al. Rituximab and immune tolerance in severe hemophilia A: a consecutive national cohort. J Thromb Haemost. 2009;7(5):787-794.

133. Zhang AH, Skupsky J, Scott DW. Effect of B-cell depletion using anti-CD20 therapy on inhibitory antibody formation to human FVIII in hemophilia A mice. Blood. 2011;117(7): 2223-2226.

134. Sack BK, Merchant S, Markusic DM, et al. Transient B cell depletion or improved transgene expression by codon optimization promote tolerance to factor VIII in gene therapy. PLoS One. 2012;7(5):e37671.

135. 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;105(12):4865-4870.

136. Skupsky J, Zhang AH, Su Y, Scott DW. B-Cell-Delivered Gene Therapy Induces Functional T Regulatory Cells and Leads to a Loss of Antigen-Specific Effector Cells. Mol Ther. 2010;18(8): 1527-35.

137. Bril WS, van Helden PM, Hausl C, et al. Tolerance to factor VIII in a transgenic mouse expressing human factor VIII cDNA carrying an Arg(593) to Cys substitution. Thromb Haemost. 2006;95(2):341-347.

138. Fijnvandraat K, Turenhout EA, van den Brink EN, et al. The missense mutation Arg593 --> Cys is related to antibody formation in a patient with mild hemophilia A. Blood. 1997;89(12): 4371-4377.

139. Thompson AR, Murphy ME, Liu M, et al. Loss of tolerance to exogenous and endogenous factor VIII in a mild hemophilia A patient with an Arg593 to Cys mutation. Blood. 1997;90(5): 1902-1910.

140. van Helden PM, Unterthurner S, Hermann C, et al. Maintenance and break of immune tolerance against human factor VIII in a new transgenic hemophilic mouse model. Blood. 2011;118(13):3698-3707.

141. Steinitz KN, van Helden PM, Binder B, et al. CD4+ T-cell epitopes associated with antibody

responses after intravenously and subcutaneously applied human FVIII in humanized hemophilic E17 HLA-DRB1*1501 mice. Blood. 2012;119(17):4073-4082.

(20)

Gener

al intr

oduc

tion

1

142. Gregersen JW, Holmes S, Fugger L. Humanized animal models for autoimmune diseases. Tissue Antigens. 2004;63(5):383-394.

143. Mangalam AK, Rajagopalan G, Taneja V, David CS. HLA Class II Transgenic Mice Mimic Human Inflammatory Diseases. In: Frederick WA, ed. Advances in Immunology. Vol. Volume 97: Academic Press; 2008:65-147.

144. Pavlova A, Delev D, Lacroix-Desmazes S, et al. Impact of polymorphisms of the major histocompatibility complex class II, interleukin-10, tumor necrosis factor-alpha and cytotoxic T-lymphocyte antigen-4 genes on inhibitor development in severe hemophilia A. J Thromb Haemost. 2009;7(12):2006-2015.

Referenties

GERELATEERDE DOCUMENTEN

The theory of gradual commonality covers the relationship between coordination, communication, ambition and context and is substantiated by providing a theoretical basis for

De chaos in de stuurgroep is bedwongen door in haar context geldende integrerende inzichten: het hanteren van een protocol waar- mee de omvang van de implementatie binnen bereik

De samenwerking aan de implementatie van de BeweegKuur kenmerkt zich door een collectief interactief proces waarbij steeds door het netwerk in interactie met haar context

Disentangling approaches to framing in conflict and negotiation research: A meta- paradigma- tic perspective.. An international approach to framing in conflict

De algemene conclusie op basis van de verschillende deelonderzoeken luidt dat samenwerking een complexe, rommelige en chaotische aangelegenheid is in een context waar steeds van

The focus is on the interaction between the people in organisations with different interests, views and practices, coming from various domains of society and faced with a

Vandaaruit kwam ze terecht bij landelijke kennisinstituten waaronder de Nederlandse Federatie voor Jeugd en Jongeren werk, het Nederlands Instituut voor Natuurvrienden werk (NIVON)

De BeweegKuur evolueert van een gesloten tot een open innovatie met een lokale point of view, hetgeen cruciaal is voor de effectiviteit van de interventie (vrij naar aanleiding van