Download Document 8916867

Document related concepts

Lymphopoiesis wikipedia , lookup

T cell wikipedia , lookup

Infection control wikipedia , lookup

DNA vaccination wikipedia , lookup

Phagocyte wikipedia , lookup

Globalization and disease wikipedia , lookup

Social immunity wikipedia , lookup

Duffy antigen system wikipedia , lookup

Monoclonal antibody wikipedia , lookup

Immune system wikipedia , lookup

Hygiene hypothesis wikipedia , lookup

Molecular mimicry wikipedia , lookup

Neonatal infection wikipedia , lookup

Polyclonal B cell response wikipedia , lookup

Adoptive cell transfer wikipedia , lookup

Cancer immunotherapy wikipedia , lookup

Adaptive immune system wikipedia , lookup

Malaria wikipedia , lookup

Mass drug administration wikipedia , lookup

Psychoneuroimmunology wikipedia , lookup

Innate immune system wikipedia , lookup

Immunosuppressive drug wikipedia , lookup

Immunomics wikipedia , lookup

Plasmodium falciparum wikipedia , lookup

Transcript
Doctoral thesis from the Department of Immunology,
The Wenner-Gren Institute, Stockholm University, Sweden
Human genetic factors involved in immunity to
Plasmodium falciparum infection
Manijeh Vafa Homann
Stockholm 2008
Human genetic factors involved in immunity to Plasmodium falciparum infection
All previously published papers were reproduced with permission from the publishers.
© Manijeh Vafa Homann, Stockholm 2008
ISBN 978-91-7155-614-1
Cover illustration: Palle Ryde
Printed in Sweden by Universitetsservice AB, Stockholm 2008
Distributer: Stockholm University Library
ii
To my parents, Mohammad Esmael Vafa and Sara Madadi
And to my husband, Peter Homann
Manijeh Vafa Homann
iii
Human genetic factors involved in immunity to Plasmodium falciparum infection
To be conscious that you are ignorant is a great step of knowledge.
Aristotle (384-322 BC)
iv
SUMMARY
Development of clinical immunity can be influenced by host genetic factors.
Some genetic disorders in erythrocytes have been shown to confer protection against malaria.
Inter-ethnic differences in malaria susceptibility offer an approach to identify genes involved
in immunity to malaria. The Fulani ethnic group, living in several African countries, has been
shown to be more immunologically responsive to, and less frequently infected by
Plasmodium falciparum, than their sympatric ethnic groups. Fewer malaria attacks and higher
spleen rates have also been seen in the Fulani compared to the non-Fulani. IL-4 and IL-10 are
two cytokines with crucial roles in the regulation of both inflammatory- and antibody
responses. IL-4 -590 C/T and IL-10 -1087 A/G single nucleotide polymorphisms (SNP) have
been shown to result in variable production of the respective cytokine.
This study investigates; the associations between the above mentioned
polymorphisms and malariometric indexes in the Fulani and the Dogon in Mali; the
correlations between malaria-specific IgG, IgG1-4, IgE, as well as total IgE and parasite
prevalence, multiplicity of malaria infection as well as splenomegaly; and the impact of IL-4 590 variants on the levels of the studied antibodies.
The allele and genotype frequencies of both studied SNPs differed significantly
between the two groups. The Fulani IL-4 T allele carriers had a significantly higher infection
prevalence compared with those carrying the CC genotype. No correlation between antimalarial antibody levels and parasite prevalence was seen in any of the communities. In the
Fulani, the increase in total IgE levels was related to the presence of infection. Malariaspecific IgG4 levels were negatively correlated to the number of clones within the Fulani. The
Fulani IL-4 T allele carriers had higher total and malaria-specific IgE levels, compared to the
CC genotype carriers.
Taken together, these results suggest that the amount of antibodies may not be
the key element in the protection against malaria. Rather, specificity, function or affinity
might be of more importance in this context. IgG4 might be involved in protection against
malaria. The impact of IL-4 -590 variants on the antibody levels may be affected by other
genetic/epigenetic/epistatic or environmental factors. The role of IgE in malaria remains to be
clarified.
Multiplicity of infection (MOI) is the number of various P. falciparum strains
co-infecting a single individual. MOI has been shown to be influenced by age, seasonal
variation and transmission intensity. Both negative and positive effects of MOI on the risk of
clinical malaria have been reported. We investigated the impact of erythrocyte variants on the
MOI, the correlation between MOI and malaria morbidity, and association between MOI and
age, parasite density and aslo seasonal variation. MOI increased after the transmission season
in all subjects, except in α-thalassaemic and in G6PD-mutated children. MOI was positively
correlated to parasitemia, and did not vary over age (in the range of 2 to 10 years). No relation
between MOI and clinical attack was noted.
These data suggest that α-thalassaemia may protect against infection by certain
parasite strains. G6PD-mutated individuals may resist against increase in MOI after the
transmission season due to rapid clearance of infection at an early stage. Furthermore HbAs
and the ABO system may have an impact on protection against severe disease, but do not
seem to affect MOI in asymptomatic individuals. Further studies are required to confirm the
age-independency of MOI and to better clarify the relations between MOI and clinical attacks.
Manijeh Vafa Homann
v
Human genetic factors involved in immunity to Plasmodium falciparum infection
ORIGINAL ARTICLES
This doctoral thesis is based on the following original papers, which in the text, are referred to
by their roman numerals.
vi
I.
Manijeh Vafa, Bakary Maiga, Klavs Berzins, Masashi Hayano, Sandor Bereczky,
Amagana Dolo, Modibo Daou, Charles Arama, Bourema Kouriba, Anna Färnert,
Ogobara K. Doumbo and Marita Troye-Blomberg, Associations between the IL-4 590 T allele and Plasmodium falciparum infection prevalence in asymptomatic
Fulani of Mali. Microbes Infect. 2007; 9: 1043-8
II.
Manijeh Vafa, Elisabeth Israelsson, Bakary Maiga, Amagana Dolo, Ogobara K.
Doumbo and Marita Troye-Blomberg. Relationship between immunoglobulin
isotype response to Plasmodium falciparum blood stage antigens and
parasitological indexes as well as splenomegaly in sympatric ethnic groups living
in Mali. Acta Tropica, under revision
III.
Manijeh Vafa, Elisabeth Israelsson, Bakary Maiga, Amagana Dolo, Ogobara K.
Doumbo and Marita Troye-Blomberg. Impact of the IL-4 -590 C/T transition on
the levels of Plasmodium falciparum specific IgE, IgG, IgG subclasses and total
IgE in two sympatric ethnic groups living in Mali. Manuscript
IV.
Manijeh Vafa, Marita Troye-Blomberg, Judith Anchang, Andre Garcia and
Florence Migot-Nobias, Multiplicity of Plasmodium falciparum infection in
asymptomatic children in Senegal: relation to transmission, age and erythrocyte
variants. Malaria Journal 2008; 7:17
TABLE OF CONTENTS
SUMMARY ................................................................................................................................................................................... V
ORIGINAL ARTICLES ............................................................................................................................................................ VI
ABBREVIATIONS .......................................................................................................................................................................1
INTRODUCTION .........................................................................................................................................................................3
THE IMMUNE SYSTEM ............................................................................................................................................. 3
Innate immunity ................................................................................................................................................ 3
Cells of the innate immune system ................................................................................................................................ 4
Acquired immunity ............................................................................................................................................ 5
T lymphocytes ................................................................................................................................................................ 5
B lymphocytes ................................................................................................................................................................ 8
Antibodies and Ig class switching .................................................................................................................................. 9
FcRs .............................................................................................................................................................................. 11
Cytokines ......................................................................................................................................................... 11
IFN-γ ............................................................................................................................................................................. 12
IL-4 ............................................................................................................................................................................... 12
IL-10 ............................................................................................................................................................................. 13
MALARIA .............................................................................................................................................................. 13
Life cycle of malaria parasite ......................................................................................................................... 14
Clinical features of malaria infection ............................................................................................................ 15
IMMUNE RESPONSE TO MALARIA .......................................................................................................................... 16
Clinical immunity ............................................................................................................................................ 16
Innate immunity to blood stages ..................................................................................................................... 16
Acquired immunity to blood stages ................................................................................................................ 18
Role of antibodies in P. falciparum immunity ............................................................................................................ 18
CD4+ T lymphocytes .................................................................................................................................................... 19
Cytokines ......................................................................................................................................................... 20
RELATED BACKGROUND .....................................................................................................................................................22
HUMAN GENETIC FACTORS AND MALARIA IMMUNITY ......................................................................................... 22
RBC variants ................................................................................................................................................... 22
The ABO blood group system ..................................................................................................................................... 22
The haemoglobinopathies ............................................................................................................................................ 23
G6PD deficiency .......................................................................................................................................................... 25
Others............................................................................................................................................................................ 25
Human leukocyte antigen (HLA) .................................................................................................................... 25
Ethnicity .......................................................................................................................................................... 26
FcγR gene polymorphisms .............................................................................................................................. 27
Cytokine gene polymorphisms ........................................................................................................................ 28
GENETIC DIVERSITY OF MALARIA PARASITE ........................................................................................................ 28
Multiplicity of malaria infection..................................................................................................................... 29
MSP2 ............................................................................................................................................................... 30
PRESENT STUDY ......................................................................................................................................................................31
AIM OF STUDY ...................................................................................................................................................... 31
METHODOLOGY .................................................................................................................................................... 32
RESULTS AND DISCUSSIONS.................................................................................................................................. 33
Paper I............................................................................................................................................................. 33
Paper II ........................................................................................................................................................... 34
Paper III .......................................................................................................................................................... 36
Paper IV .......................................................................................................................................................... 38
CONCLUDING REMARKS AND FUTURE PERSPECTIVES ..........................................................................................40
ACKNOWLEDGMENT .............................................................................................................................................................42
REFERENCES ............................................................................................................................................................................44
Manijeh Vafa Homann
vii
Human genetic factors involved in immunity to Plasmodium falciparum infection
viii
ABBREVIATIONS
ADCC
ADCI
APC
BCR
CM
CR1
CSR
CTLA4
DC
DP
ELISA
FcR
Foxp3
G6PD
GPI
Hb
HC
HLA
IFN
Ig
IL
iRBC
IRF-1
LC
MHC
MOI
MSA
MSP
NK
NKR
NO
PAMPs
PCR
PfEMP1
PfRBCs
PRRs
RBCs
SNP
STAT
Tc
TCR
TGF
Th
TLR
TNF
Treg
Antibody-dependent cellular cytotoxicity
Antibody-dependent cellular inhibition
Antigen-presenting cell
B-cell receptor
Cerebral malaria
Complement receptor 1
Class switching recombination
Cytotoxic T lymphocyte-associated antigen-4
Dendritic cell
Double positive
Enzyme-linked immunosorbent assay
Fc receptor
Forkhead box P3
Glucose-6-phosphate dehydrogenase
Glycosylphosphatidylinositol
Haemoglobin
Heavy chain
Human leukocyte antigen
Interferon
Immunoglobulin
Interleukin
Infected-Red-Blood Cell
Interferon-regulatory factor-1
Light chain
Major-histocompatibility complex
Multiplicity of infection
Merozoite-surface antigen
Merozoite-surface protein
Natural killer
NK receptors
Nitric oxide
Pathogen-associated molecular pattern
Polymerase-chain reaction
P. falciparum-erythrocyte membrane protein 1
P. falciparum-infected red-blood cells
Pattern-recognition receptors
Red-blood cells
Single-nucleotide polymorphism
Signal transduction and activator of transcription
T cytotoxic
T-cell receptor
Transforming-growth factor
T helper
Toll-like receptor
Tumor-necrosis factor
Regulatory T-cell
Manijeh Vafa Homann
1
Human genetic factors involved in immunity to Plasmodium falciparum infection
2
INTRODUCTION
The immune system
The immune system consists of a substantial number of components and
mechanisms by which humans are protected from invaders. It detects a wide variety of
foreign agents and discriminates them from self components. The immune system is branched
into two arms, the innate and the acquired immune system.
Innate immunity
The innate immune system provides an immediate defence against infections
and has been found in almost all multicellular plants and animals. The innate immune system
consists of physical, chemical and cellular barriers. The invaders that manage to pass the nonspecific host anatomical barriers, encounter a number of cells, with variable capacities in
defence mechanisms, such as neutrophils, macrophages, dendritic cells (DCs) and natural
killer cells (NK).
The innate immune system detects infections through an array of soluble and
membrane bound sensors called pattern-recognition receptors (PRRs). PRRs recognize
conserved microbial components, which are usually crucial for microbial survival, known as
pathogen-associated molecular patterns (PAMPs). The scavenger receptors (SRs) are
members of the PRRs family, present on macrophages, DCs and certain epithelial cells. SRs
not only recognize different microbial ligands, but also detect endogenous and modified hostderived molecules [1]. Scavenger receptors are structurally diverse and accordingly
categorized into eight classes (A to H) [2]. Pathogens evading from extra-cellular recognition,
face another line of detection in the host cytosol. Nucleotide binding and oligomerization
domains (NOD)-like receptors (NLRs) and retionid acid-inducable gene I (RIG-I)-like
receptors are two recently identified cytosolic PRR that recognize PAMPs in the intracellular
compartments [3]. Toll-like receptors (TLRs) are the most studied PRRs. They recognize a
wide range of microbial motifs at the cell surface or within the endosomes. TLRs are highly
expressed by macrophages, DCs and B cells. Hence they also play an important role in the
induction of the adaptive immune response. Eleven TLRs have been identified in humans
Manijeh Vafa Homann
3
Human genetic factors involved in immunity to Plasmodium falciparum infection
until now. Based on the cellular localization, TLRs are classified into two groups. TLR 3, 7, 8
and 9 are expressed in endosomes, while the second group, including TLR1, 2, 4, 5 and 6, are
present on the surface of many different cells [4, 5]. Complement proteins, mannose-binding
lectin (MBL) and C-reactive protein (CRP) are other members of the PRR family. Once PRRs
detect PAMPs, multiple pro-inflammatory signalling pathways are activated and an effective
response against the intruder is mounted [3].
Cells of the innate immune system
Macrophages and DCs take up, degrade and present antigens, associated with
major histocompatibility (MHC) molecules, on their surface to activate the cells of the
acquired immune system. Thus these cells are also called antigen presenting cells (APCs).
MHC molecules are glycoproteins found on the surface of cells and are categorized into class
I and II. Class I MHC is expressed by almost all nucleated cells, while MHC class II is
expressed by professional antigen-presenting cells (i.e. macrophages, DCs and B cells).
DCs are leukocytes specialized in presenting antigens. They can activate naïve T
and B cells, and thus link the innate and the adaptive immune responses. Immature DCs are
widely scattered in peripheral tissues, at the entrance sites of pathogens, and are very efficient
in engulfing antigens. Following the ingestion of antigen, DCs travel to lymph nodes, where
they present antigens to T cells. Meanwhile, they mature, and their ability to take up antigens
is reduced. Instead they get more efficient in presentation of antigens through the upregulation of MHC class II and co-stimulatory CD80 and CD86 molecule expression [6].
Macrophages also express more MHC class II molecules, after activation, but in contrast to
DCs, exhibit a greater phagocytic ability.
NK cells are large, granular lymphocytes, widely distributed in the body, and
can be recruited to the inflammation sites in different tissues. They are the first line of defence
against various viral infections. NK cells are capable to induce the death of self cells, which
have been modified upon intracellular infections and malignant transformation. NK cells
discriminate between target and non-target cells using a set of activating (e.g. NKG2D,
NKp30, NKp 44, NKp 46 and CD16) or inhibitory receptors, such as KIR-inhibitory receptors
(KIR stands for killer-cell immunoglobulin-like receptor) [7]. Inhibitory receptors recognize
4
MHC class I molecules, as a marker of self cells, and inhibit the activation of NK cells. The
expression of MHC class I molecule is usually down-regulated in infected or transformed
cells. Thereby, interaction of activating receptors with the ligands on infected or transformed
cells, in the absence of MHC class I, leads to NK cells activation. Activated NK cells release
perforin and granzymes containing cytotoxic granules, which induce programmed cell death
in the target cells [8]. NK cells are identified as CD3-CD56+ lymphocytes. Human NK cells
consist of two subsets, CD56bright and CD56dim NK cells. The latter subset is mainly found in
the blood and spleen and display stronger cytotoxic effector functions than CD56bright cells,
which are more abundant in the lymph nodes [7]. NK cells can rapidly get activated and
produce pro-inflammatory cytokines, in particular interferon (IFN)-γ. Therefore, they can
influence the adaptive immune response by inducing T helper (Th) 1 type of responses [9].
Acquired immunity
While the innate immune system recognizes patterns shared by groups of nonself molecules, T- and B lymphocytes more specifically identify foreign molecules through
the specificity of their antigen receptors, T- and B-cell receptors (TCR and BCR). The TCR is
a heterodimer of either αβ or γδ chains, which recognizes antigen and is responsible for the
specificity of the T cell. This heterodimer is associated with CD3, a multicomponent signaltransducing complex. CD3 triggers intracellular signalling, induced by binding of the antigen
to the TCR. In contrast to the BCR, which can identify various cell-surface or soluble
antigens without prior modifications, the TCR recognizes antigen-derived peptides in
association with MHC molecules, presented on the surface of other cells [10].
T lymphocytes
T lymphocytes are derived from bone-marrow-resident stem cells, and develop
into self-tolerant, functional T cells through a multistage differentiation process in the thymus.
These stages are distinguished by the expression of the key surface molecules, CD4 and CD8.
The very first immature cells express none of these molecules and therefore are called double
negative (DN). At this stage, αβ and γδ lineages are generated based on the expression of the
αβ or γδ subunits of the TCR complex. The αβ lineage differentiates into a CD4CD8 double
positive (DP) stage. Further positive and negative selections occur that lead to deletion of non
Manijeh Vafa Homann
5
Human genetic factors involved in immunity to Plasmodium falciparum infection
functional and autoreactive TCRs. Eventually, MHC-restricted and self-tolerant CD4 and
CD8 single positive (SP) T cells are generated and leave the thymus. Thus, mature αβ T
lymphocytes comprise of CD8+ T cytotoxic (Tc), CD4+ helper- and regulatory T cell lineages
[11].
Despite the expression of antigen-receptors, γδ T cells display innate
characteristics and form an independent population of lymphocytes [12]. γδ T cells, in most
instances, do not detect peptide antigens presented on the surface of APCs, as αβ T cells do.
Instead they recognize alkalamines, nonpeptideprenyl pyrophosphates and phosphorylated
uridine and thymidine compounds [13].
CD8+ T lymphocytes participate in responses to infections, through their
cytotoxicity and cytokine production. They can produce IFN-γ, tumor necrosis factor (TNF)α, interleukin (IL)-2, IL-4, IL-5, IL-10 and transforming growth factor (TGF)-β. Based on
their migration capacity and cytokine production, CD8+ T cells are named Tc1 and Tc2. CD8+
T lymphocytes exert their cytotoxicity through two mechanisms. They can induce apoptosis
in target cells by the expression of FasL/CD95L, that makes the target cell to terminate the
expression of the receptor Fas/CD95, and thereby initiating apoptosis. They also can kill the
target cell via exocytosis of lytic granules that forms pores in the membrane of the target cell,
which then allows the lytic granules to enter the cell and to induce apoptosis. CD8+ T cells
recognize antigens presented by MHC class I, and need help from DCs and CD4+ T cells to be
fully activated. However, the exact mechanism behind this has not yet been fully understood.
It has been proposed that DCs present antigen to CD4+ T cells in associations with MHC class
II. CD4+ T cells differentiate, proliferate and up-regulate the expression of the CD40L which
interacts with CD40 molecules on DCs. This enables DC to activate and transform CD8+ T
cells to effector cells that identify antigens in association with MHC class I molecules. A
suppressive CD8+ T cell subset, with regulatory function has also been reported [14].
Upon activation, after interaction of TCR with antigen-MHC, naïve CD4+ T
cells can differentiate into Th1, Th2, Th17 or regulatory T cell (Treg), according to cytokine
milieu (Fig 1). Once differentiated, each lineage is characterized by its own cytokine profile.
In the presence of IL-12 and through the activation of signal transduction and activator of
transcription (STAT)-4 and T-bet, CD4+ T cells differentiate into a Th1 lineage that produces
6
IFN-γ, a key cytokine in response to intracellular infections. Naïve CD4+ T cells become Th2,
when IL-4 is present and this process is regulated by STAT-6 and GATA-3. Th2 cells
produce IL-4, a cytokine with an important role in the host defence against helminth
infections. IL-23 synergizes with IL-1 and IL-6, inducing Th17 differentiation, a process
regulated by STAT-3 and retinoic acid-related orphan receptor (RORγt) transcription factors.
Th17 cells secrete the pro-inflammatory cytokine IL-17, and have been proposed to
participate in responses against extra-cellular bacteria and fungi, and also to mediate
autoimmune diseases. TGF-β1 upregulates forkhead box P3 (Foxp3) expression and induces
the differentiation of naïve CD4+ T cells to become Treg, which produce TGF-β1 [15]. This
process is regulated by STAT-5. CD4+ T cell subsets can regulate each other via changes in
the cytokine milieu.
Figure 1: Modified from Chen and O’Shea [15]
Treg cells are named after their functional ability to regulate/suppress immune
responses, and can develop in the thymus (natural Treg or nTreg) or be induced in the
periphery (inducible or iTreg). Foxp3, IL-2 production and CD25 (IL-2 receptor α chain)
Manijeh Vafa Homann
7
Human genetic factors involved in immunity to Plasmodium falciparum infection
expression are essential for Treg development, and their function and survival. Naturally
occurring CD4+CD25+ Treg cells are key elements of self-tolerance maintenance. In the
presence of TGF-β, uncommitted CD4+ T cells in periphery can convert into iTreg and release
the regulatory cytokine IL-10 (Tr1 subset) or TGF-β (Th3 subset). In vitro studies have shown
that suppression of CD4+CD25- effector T cells, by Treg cells, depends on cell-cell contact
rather than soluble mediators. Membrane bound TGF-β, the cytotoxic T lymphocyteassociated antigen-4 (CTLA-4) and the glucocorticoid-induced tumour necrosis factor
(GITR), are suggested to be important in this contact. Tregs have been shown to directly kill
activated CD4+ and CD8+ T cells in a perforin- or granzyme-dependent manner. It has been
proposed that upon cell-cell contact, Tregs modulate APCs function and make them unable to
activate effector T cells. Moreover, Tregs mediate immune-response suppression via the
induction of naïve T-cell differentiation into regulatory cells rather than pathogen-specific
effector cells. This phenomenon is called “infectious tolerance” and was suggested to be
mediated by IL-10 or TGF-β [16].
Another subset of T lymphocytes, which does not fit well in acquired immunity
is the NKT cell subset. These cells are derived from DP thymocytes, and are mainly present in
the bone marrow and liver, but can also be found in other lymphoid compartments. NKT cells
share characteristics of both NK and T cells [17]. NKT cells get activated upon interaction of
their TCR with glycolipids associated with the MHC class I-like molecule CD1d, which is
expressed by APCs. NKT cells are known to be major producers of IFN-γ, a crucial cytokine
for early activation of macrophages [18].
B lymphocytes
B-cell development from hematopoietic cells occurs before birth in the foetal
liver and subsequently in the bone marrow. In the bone marrow, B cells develop from pro-B
to pre-B cells and eventually to membrane-bound IgM-expressing immature B cells. Naïve
mature B cells leave the bone marrow and travel to secondary lymphoid tissues, where they
become activated. Following activation, B cells become either memory or plasma cells.
Memory B cells are more abundant than naïve resting B cells and can be rapidly re-stimulated
by antigens. They often harbour membrane-bound IgG and IgA. Plasma cells are antibody-
8
producing cells. They are large cells, enriched in endoplasmic reticulum, ribosome and golgi.
Plasma cells do not express membrane Ig and do not divide [19].
Antibodies and Ig class switching
The BCR is present on the surface of B cells, and is a complex of
immunoglobulin (Ig) associated with two accessory proteins, called Igα and Igβ. The BCR
identifies and binds the antigen, and Igα and Igβ trigger intra-cellular signalling induced by
binding of the antigen [10].
Antibodies are expressed and secreted by activated B lymphocytes. Antibodies
are glycoproteins with a common structure, but differing in size, charge and amino acid
sequence. The basic structure consists of two identical light chains (LC) which are bound to
two identical heavy chains (HC) via disulfide bonds and non-covalent interactions. The
amino-terminal portion of the chains, called variable (V) region, vary amongst different Igs
and determine the antibody specificities. The carboxyl-terminal portions of the chains
constitute the constant (C) region and are rather conserved. According to the sequence
patterns of the constant region, the heavy chain is grouped into five classes: μ, δ, γ, ε and α,
each named an isotype, and the light chain is found in two isoforms, κ and λ. In the process of
Ig synthesis, rearrangements in gene segments (VHDJH in HC and VLJL in LC) and somatic
mutations produce variations in amino acid sequences in the amino-terminal portions of the
chains. This generates 106 to 107
6
different variable regions, which reflects the generation of
7
a repertoire of Igs with 10 to 10 various specificities. Based on the heavy chain, antibodies
are also classified into five classes, IgM, IgD, IgG, IgE and IgA. In humans, minor variations
in amino acid sequences of the α and the γ heavy chains result in further classification of IgA
and IgG into two (IgA1 and IgA2) and four (IgG1-4) subclasses, respectively.
Rearrangement in the Ig heavy chain occurs during B cell development from
pro-B to pre-B cell. In the pre-B cell, the membrane bound μ chain and the surrogate light
chain (a complex of two proteins which are non-covalently bound to form a light chain-like
structure) associate with the Igα/Igβ heterodimer to build the pre-BCR. The light-chain gene
is then rearranged, and intact IgM is displayed on the immature B cells. In an antigenindependent manner, naïve mature B cells begin to splice the variable region genes with the δ
Manijeh Vafa Homann
9
Human genetic factors involved in immunity to Plasmodium falciparum infection
chain constant region and express IgD (in addition to IgM) on their surface to form fully
mature naïve B cells.
All Igs consist of two regions, the Fab and the Fc. The Fab contains the variable
region of the Ig, differs from one antibody to the other, and is the part of the antibody that
identifies and binds antigenic epitopes. The Fc fragment is conserved in each class and
subclass of the antibodies and mediates effector functions specific to that class.
A set of antibodies of different classes can be produced in response to a
particular antigen as a result of class switching recombination (CSR). In the IgH locus, the
genes for the constant regions are positioned in the following order, Cμ, Cδ, Cγ, Cα and Cε.
In CSR, one of these C regions recombines to the rearranged variable (VDJ) region to
generate the respective isotype of the antibody. The first antibody produced after exposure to
antigen is mainly of the IgM class. By substituting the CH regions of IgM with those of IgG,
IgA and IgE, CSR modulates the effector functions of antibodies without changing their
specificity. Antibody-class switching is mainly regulated by extra-cellular signals, in
particular cytokines. [19-23].
Antibody-mediated effector functions
Antibodies recognize and bind to antigens, but they can not remove or kill the
pathogen by themselves. Antibodies interact with other proteins or cells to perform the
effector function of a humoral response. Binding of antibody-antigen complexes to the Fc
receptors (FcR) on the surface of phagocytic cells, promotes the phagocytosis of the antigen
(opsonization).
In humans, IgM, IgG1, IgG2 and IgG3 activate the classical pathway of
complement. IgA1, IgA2 and IgE can activate the alternative pathway of complement.
Complement refers to a set of serum glycoproteins which interact and cooperate with other
immune molecules or cells to remove the antigen or kill the pathogen.
IgG also mediates antibody-dependent cellular cytotoxicity (ADCC). ADCC is
the initiation of cytotoxic activities of the effector cells against the pathogen, upon binding of
antibody in complex with antigen, to the FcR on a number of immune cells, particularly NK
cells. No effector function has been identified for IgD [19, 20, 24].
10
FcRs
Antibody-mediated responses are regulated by different FcRs. Unlike, CH
regions of IgM and IgD, CH regions of IgG, IgE and IgA can bind to their respective receptors
(Fcγ, Fcε and Fcα) on innate immune cells and promote their pro-inflammatory and killing
functions. This links the specificity of adaptive responses to potent effector functions of the
innate immune response. Co-expression of activating and inhibitory FcRs on innate and
adaptive immune cells regulates the immune response and thus avoids tissue damage induced
by exaggerated immune responses. In addition to immune cells, FcRs are expressed on
endothelial cells, mesangial cells and osteoclasts [22, 25]. B cells only express inhibitory
FcRs (FcγRIIB) [26].
Cytokines
Cytokines are low molecular weight polypeptides produced by different cells,
whereby cells communicate with each other. Cytokines are released in response to a variety of
inflammatory stimuli such as infectious agents and their products, or in response to other
cytokines. In general, a cascade of cytokines is produced by different cell types, eventually
ending with the final immune-effector function. The majority of cytokines are transiently
secreted and generally have a short half-life. Accordingly, cytokines are mostly produced and
act locally, either on the same cell that produced them (autocrine effect) or on the adjacent
cells (paracrine effect). A minority of cytokines enters the circulation and acts on distant
tissues (endocrine or hormonal effect).
According to their role in inflammation, cytokines are classified into pro- and
anti-inflammatory cytokines. Cells at the site of inflammation such as DCs, macrophages and
mast cells release a panel of pro-inflammatory cytokines, including IL-12, IL-18, IL-8, TNFα and IFN-γ, which causes the inflammation. Anti-inflammatory cytokines, like IL-4, IL-10
and IL-13, are able to down-regulate the production of pro-inflammatory cytokines. The
balance between pro- and anti-inflammatory cytokines during the course of an inflammation
determines the outcome of the inflammatory response. Importantly, the early inflammatory
cytokine production also influences the subsequent development of the adaptive immune
response [27, 28].
Cytokines are also involved in the regulation of the adaptive immunity. They
play an important role in activating and regulating lymphocytes [27]. Cytokines are important
Manijeh Vafa Homann
11
Human genetic factors involved in immunity to Plasmodium falciparum infection
in the regulation of humoral response and antibody-class switching. For example in humans,
IgG4 and IgE production is induced by IL-4 and IL-13. IL-4-induced IgE production is
inhibited by IL-10, IL-12, IFN-α, IFN-γ and TGF-β. Production of IgM, IgG1, IgG3 and IgA
is enhanced by IL-10. TGF-β induces isotype switching to IgA, and IgG2 production is
induced by IFN-γ [19, 21].
In addition to the regulation of the innate and the adaptive immune responses,
cytokines also regulate cell growth and differentiation, and influence embryo implantation
and foetal development [27, 28].
Below, cytokines of interest in recent studies are described more in detail.
IFN-γ
IFN-γ is crucial in both innate and adaptive immune responses. It plays a critical
role in immunity against intracellular pathogens and in tumor control. IFN-γ is mainly
produced by NK, NKT, Th1 and CD8+ T lymphocytes. The constitutive expression of IFN-γ
mRNA by NK- and NKT cells allows them to mount a rapid IFN-γ response against
infections. CD4+- and CD8+ T cells produce little IFN-γ shortly after initial activation. IFN-γ
induces expression of MHC class I and II molecules on APCs, and consequently increases
their antigen-presentation capacity. IFN-γ enhances the phagocytic capacity of macrophages.
IFN-γ also induces the production of pro-inflammatory cytokines and antimicrobial products
such as nitric oxide (NO) in macrophages. In neutrophils and monocytes, IFN-γ stimulates the
expression of high-affinity FcγRs. IFN-γ is a growth-promoting factor for T lymphocytes,
whereas, IFN-γ inhibits the IL-4-induced B-cell growth. IFN-γ amplifies the recruitment of
lymphocytes into tissues and prolongs their activation. The excessive production of IFN-γ has
been associated with a number of autoimmune and inflammatory diseases. IFN-γ production
by NK cells is inhibited by TGF-β [8, 29].
IL-4
IL-4 is mainly produced by activated Th2 cells [28]. Mast cells also secrete IL4, upon engagement of their FcεR or FcγR [30]. IL-4 induces proliferation and differentiation
12
of activated B cells, and enhances the expression of MHC class II and the IgE low affinity
receptor (CD23) on resting B cells [31]. In activated B cells, IL-4 induces IgG4 production.
IL-4 also induces isotype switching from IgM to IgE. IL-4-induced antibody-class switching
is antagonized by IFN-γ [32, 33]. IL-4 stimulates NK cells, CD4+ T- and CD8+ T-cell
proliferation [34]. IL-4 suppresses the production of IL-1β, TNF-α and IL-8 [28].
IL-10
A wide variety of cells, such as activated monocytes, macrophages, mast cells,
epithelial cells, B cells, CD4+- and CD8+ T cells, produce IL-10. Also a variety of blood cells
are affected by the inhibitory or stimulatory functions of IL-10. IL-10 promotes B-cell
proliferation, survival and differentiation into Ig-secreting plasma cells. In addition, IL-10
displays inhibitory effects on various blood cells. The inhibitory effect of IL-10 is believed to
be due to the induction of regulatory T cells. IL-10 inhibits IL-1, IL-6, TNF-α, IL-12 and
subsequently IFN-γ production. IL-10 also down-regulates macrophages, NK cells and Th1
cells. IL-4- and IFN-γ-induced expression of MHC class II-antigen is suppressed by IL-10.
Thus, IL-10 is an important regulatory cytokine [16, 19, 35-38].
Malaria
The term malaria originates from Italian ‘mal aria’, which means ‘bad air’, for
the incidence of the disease in marshy areas [39]. The malaria parasite has infected humans
for 50,000-100,000 years [40] and remains a major cause of morbidity and mortality in the
developing word, particularly in sub-Saharan Africa. About half a billion episodes of clinical
malaria is estimated to occur each year, which results in deaths of more than one million
children in sub-Saharan Africa alone [41].
Malaria is a vector-borne infectious disease caused by an intracellular protozoan
parasite of the genus Plasmodium. More than 100 species of Plasmodium parasites are
known, of which only four types can infect humans: P. falciparum, P. vivax, P. ovale and P.
malariae. Among these four species, P. vivax and P. falciparum are the most prevalent
Manijeh Vafa Homann
13
Human genetic factors involved in immunity to Plasmodium falciparum infection
species and P. falciparum is the main cause of severe and fatal malaria [42]. Plasmodium
parasites are transmitted to the vertebrate host by the female Anopheles mosquitoes.
Life cycle of malaria parasite
The life cycle of the malaria parasite is complex, includes sexual and asexual
developmental stages and requires both mosquito and vertebrate hosts to be completed.
The infected Anopheles mosquitoes carry Plasmodium sporozoites in their salivary glands.
When an infected mosquito takes a blood meal, it injects 5 to 20 infective sporozoites into the
bloodstream of its victim (Fig. 2).
http://en.wikipedia.org/wiki/Image:MalariacycleBig.jpg
Figure 2: Plasmodium falciparum life cycle
Sporozoites migrate to the liver within 30 minutes and invade hepatocytes, where they
multiply asexually for a period of 9-16 days. This division is called schizogony and generates
merozoite forms of the parasite. Each sporozoite gives rise to thousands of merozoites which
14
are released into the bloodstream when the infected hepatocyte ruptures. Released merozoites
invade erythrocytes. Once inside the erythrocyte, the merozoite enlarges to a uni-nucleated
cell named trophozoite. Trophozoites undergo asexual divisions and produce schizonts with
multiple nuclei. The schizonts then segment to generate 15 to 30 merozoites which are
released into the blood, ready to invade new erythrocytes. Several such cycles occur and
periodic, synchronized release of merozoites and toxins into the blood, results in malaria
associated waves of fever (within 48–72 hours, depending on the species of Plasmodium).
A
small
proportion
of
merozoites
differentiate
into
micro-
and
macrogametocytes (male and female, respectively). A mosquito taking a blood meal on an
infected individual may ingest these gametocytes. In the mosquito gut, the gametocytes form
male and female gametes, which fertilize to form zygotes. These zygotes further differentiate,
migrate through the mosquito gut wall, and undergo repeated divisions to produce large
number of sporozoites. The infective sporozoites travel to the salivary gland and the life cycle
is completed.
Clinical features of malaria infection
Infections with malaria parasites have a wide variety of symptoms, ranging from
asymptomatic carriage of infection to severe malaria or even death. A large number of
individuals living in malaria endemic areas carries parasites without presenting signs of
clinical malaria. Clinical malaria can be categorized into mild (uncomplicated) or severe
(complicated) malaria. In general, malaria is a curable disease if diagnosed and treated
quickly and properly.
A combination of fever, chills, sweats, headaches, nausea, malaise, cough,
diarrhoea and muscular pain is presented in uncomplicated forms of malaria disease.
Uncomplicated malaria can be caused by all four Plasmodium species, whereas P. falciparum
is responsible for severe pathology and almost all deaths.
Sequestration, a characteristic feature of P. falciparum infection, happens when
P. falciparum infected red blood cells (iRBC) adhere to uninfected RBCs, endothelial cells
and in some cases to placental cells (Fig. 1). Sequestration of iRBCs in various organs such as
brain, kidney, placenta and lung may lead to serious organ failures or abnormalities in the
patients’ blood and consequently to severe pathology [43]. Cerebral malaria (CM), severe
Manijeh Vafa Homann
15
Human genetic factors involved in immunity to Plasmodium falciparum infection
anaemia, hemoglobinuria, hypoglycaemia, renal failure, pulmonary oedema, abnormalities in
blood coagulation and thrombocytopenia, convulsion and shock are some complications of
severe malaria.
Immune response to malaria
The immune response to the malaria parasite, as the parasite’s life cycle, is
complex and depends on the parasite species and on the stage of infection [44]. The infection
starts when the sporozoites enter the human body. At this level protective immune responses
to the parasite rely on antibodies against antigens on the surface of sporozoite, like the
circumsporozoite protein (CSP). During the liver stage, when the parasite is inside the
hepatocyte, a predominant cellular immune response is required. Later on, when the infection
involves erythrocytes, protective immunity depends mainly on antibody responses.
Merozoite-derived antigens expressed on the surface of the infected RBCs, merozoite surface
proteins/antigens (MSP/MSA) and parasite antigens released into circulation, following
erythrocyte rupture, are the major targets of antibody response in this stage [45].
Clinical immunity
Individuals residing in malaria endemic areas experience repeated infections and
acquire immunity to clinical malaria, termed “clinical immunity”, which prevents the disease
despite the persistence of infection [46].
Clinical symptoms of malaria are related to the blood stage of infection. Thus, it
is important to understand the immune response during this stage. The present study has
mainly focused on the blood stage of the malaria infection. Hence, immune responses to
asexual blood-stage of malaria parasite, specifically P. falciparum, are overviewed in this
thesis.
Innate immunity to blood stages
Following the initial rise in parasitemia in blood, innate type of immune
responses control parasite expansion, but this is usually not enough to eliminate the infection.
16
Macrophages, DCs, NK cells and γδ T cells are effector cells involved in the innate immune
responses to the malaria parasite [47].
PRRs mediate direct interaction between innate immune cells and the parasite.
To date, TLR2, 4 and 9 have been shown to recognize malarial antigens in humans. The P.
falciparum glycosyl-phosphatidylinositol (GPI) interacts with the TLR2/TLR1 complex, with
a minor contribution of TLR4, and induces cytokine production in macrophages. Hemozoin, a
product of haemoglobin digestion by P. falciparum in RBC, was surprisingly identified as a
non-DNA ligand for TLR9. TLR9 was earlier described as a receptor for CpG-containing
DNA. More recent studies have suggested that hemozoin presents DNA to TLR9, while
hemozoin by itself, is not able to induce innate immune responses [5]. Plasmodium profilinlike proteins have been shown to activate the innate immune system in mice through TLR11
[48]. Of other PRRs involved in recognition of P. falciparum-derived antigens, the scavenger
receptor CD36 can be mentioned. The P. falciparum erythrocyte membrane protein-1
(PfEMP-1) expressed on the surface of iRBC is the ligand for CD36 [4]. Binding of PfEMP-1
to CD36 enables macrophages to attach to, and engulf iRBCs in the absence of cytophilic or
opsonizing malaria-specific antibodies. This suggests an important role for macrophages in
innate immunity to malaria [49].
Besides ingesting iRBCs, macrophages can take the parasites out of iRBCs and
let the parasite-free RBCs back to circulation [50, 51]. Macrophages can also kill/inhibit in
vitro growth of the asexual blood stage through the release of NO and H2O2 and also mediate
antibody-dependent cellular inhibition (ADCI) [52, 53].
Activation of DCs is one of the first events in the innate response to malaria.
Malaria parasites interact with DCs to modulate their functions. Although studies have
revealed a role of DCs in immune responses to many intracellular pathogens, not much is
known about the role of DCs in malaria immunity. The first study on the role of DCs in
malaria, showed that P. falciparum infected red blood cells (PfRBC) adhere to DCs, inhibit
the maturation of DCs [54], and subsequently reduce their capacity to activate T cells [55].
PfRBCs were later shown to interact with the CD36 and CD51 molecules on DCs [54, 56].
Upon this interaction, DCs upregulate the expression of the co-stimulatory molecules CD40,
CD80 and CD86, and instead of initiating a pro-inflammatory response, produce IL-10 and
Manijeh Vafa Homann
17
Human genetic factors involved in immunity to Plasmodium falciparum infection
elicit an anti-inflammatory response [57]. However, studies using experimental animals have
shown that DCs from infected mice are fully functional APCs [58].
γδT-cells have been reported to expand during the first few days of P.
falciparum infection [59]. P. falciparum–derived phosphorylated non-peptide antigens have
been identified as ligands for human γδT-cells [60]. In vitro studies have shown that Pfactivated γδT-cells produce large amounts of IFN-γ [61, 62] and inhibit parasite growth [63].
Extracellular merozoites or late schizonts have been suggested to be targets for γδT-cells,
since γδT-cells have been shown to inhibit growth of the asexual blood stages in vitro at the
time of reinvasion [64]. γδT-cells kill the parasite by exocytosis of granulysin and perforin,
following direct contact with iRBCs [65].
NK cells can lyse PfRBC without prior sensitization. Studies by ArtavanisTsakonas et al. have suggested that NK cells are the earliest source of pro-inflammatory
cytokine IFN-γ [66], which is essential to control malaria parasitemia [67, 68], and that NK
cells produce IFN-γ following a direct interaction with PfRBC [69]. However, recent studies
have revealed that the innate IFN-γ production was mainly (in 93.3% of the analysed
individuals) derived from γδT-cells. The majority of these γδT-cells expressed NK receptors
(NKR). The overall IFN-γ levels produced by NK-like γδT-cells were positively associated
with the expression of CD94 on these cells [70].
Acquired immunity to blood stages
Role of antibodies in P. falciparum immunity
Erythrocytes do not express the classical MHC class I and II molecules.
Therefore, antibodies and B-cells are crucial in clinical immunity. Reduction of parasitaemia
and clinical disease after passive transfer of gamma-globulins from immune donors to malaria
patients first suggested a protective role for IgG antibodies [71]. Antibodies are generally
accepted as indicators of infection.
Malaria protective antibodies are mainly restricted to certain classes and
subclasses. Analyses of anti-malarial IgG subclass profiles, in humans, have suggested that
18
IgG1 and IgG3 (called cytophilic antibodies) provide protection, whereas IgG2 and IgG4 do
not. However, IgG2 has recently been demonstrated to be associated with malaria protection
in individuals carrying a certain polymorphism in the FcγRII gene [72]. IgG4 is thought to
counteract with anti-parasitic functions of IgG1 and IgG3 through inhibiting their binding to
the antigens [73]. While IgG has been correlated to protection against malaria disease,
antibodies of the IgE isotype have been linked to severe pathology. Higher levels of total and
anti-malarial IgE were detected in patients suffering from cerebral malaria than individuals
with uncomplicated malaria [74]. However, recent studies in Tanzania and Senegal have
suggested a protective role for IgE against malaria, where elevated levels of P. falciparumspecific IgE were shown to reduce the risk of subsequent malaria attack [75, 76]. Thus, the
role of IgE (total and malaria specific) in the pathogenesis/immunity to malaria is still
unclear.
Malaria-specific antibodies exert their protective roles through mediating a
number of anti-parasitic effector functions including: inhibition of cytoadherence and thus
preventing sequestration of iRBCs and instead letting them be removed by spleen [77];
inhibition of RBC invasion [78] and antibody-dependent cytotoxicity [79]. Anti-malarial
antibodies are involved in the ADCI mechanism where the parasite growth is inhibited by
antibodies bound to Fc receptors on phagocytes [80]. Antibodies may also facilitate parasite
clearance by opsonization and phagocytosis by macrophages [81].
CD4+ T lymphocytes
CD4+ T cells are important in the development and regulation of the immune
response to malaria infection. Murine experimental systems have shown that Th1-produced
cytokines, such as IFN-γ, initiate inflammatory responses needed to control infection at early
stage. Cytokines produced by Th2 lymphocytes, e.g. IL-4 and IL-10, not only help B cells to
produce antibodies, but also suppress Th1 responses, which if left uncontrolled, can lead to
immunopathology [82, 83]. While Th1/Th2 type of responses seem to be of importance in
certain murine experimental systems, the situation is less clear in humans.
Another subset of CD4+ T lymphocytes involved in malaria blood stage immune
response is Treg cells. A direct evidence for the role of Tregs in the outcome of malaria
infection comes from studies in mice. The significant delay in parasite growth in CD4+ CD25+
depleted naïve mice, infected with P. berghei NK65, revealed that effector responses are
Manijeh Vafa Homann
19
Human genetic factors involved in immunity to Plasmodium falciparum infection
enhanced in the absence of Tregs. Depletion of CD4+ CD25+ T cells in P. yoelii 17XL
infected mice, was also shown to result in amplified T-cell responses to parasite antigens and
control the normally lethal infection [84]. Moreover, studies in P. yoelii 17XL susceptible
(BALB/c) and resistant (DBA/2) mice have indicated an increase in the Treg population early
after infection, in BALB/c but not DBA/2 mice, suggesting that the suppression of the Th1
response during early infection probably results in enhanced parasitemias and fatal disease in
BALB/c mice [85]. Similarly, in vivo studies in humans have shown that blood-stage
infection rapidly induces Tregs which in turn results in a burst of TGF-β production. This
leads to reduced pro-inflammatory cytokine production and antigen-specific immune
responses and consequently gives rise to higher rates of parasite growth [86]. Optimally, it
looks as if Treg activation and/or Treg-derived cytokine (IL-10 and TGF-β) production are
needed to be triggered at the time that pro-inflammatory cytokines have brought parasite
replication under control.
Cytokines
Besides the ability of individuals to control parasite multiplication, the term
“clinical immunity” also refers to the ability of individuals to control the inflammatorycytokine response induced against the parasite [87]. In mice it seems as if an early pulse of
pro-inflammatory cytokines is essential to limit the initial parasite replication [88]. An early
IFN-γ response is crucial for the resolution of the primary infection [89]. Elevated plasma
levels of IFN-γ have also been linked to resistance to re-infection by P. falciparum [90].
TNF-α synergises with IFN-γ to induce production of NO and thus parasite killing [91].
A pro-inflammatory response is essential for parasite clearance. However,
excessive production of these cytokines may lead to malaria pathogenesis. Malaria patients
have been shown to have higher plasma levels of IFN-γ as compared to asymptomatic
individuals, suggesting a correlation between IFN-γ levels and malaria disease. Elevated IFNγ levels have also been associated with the onset of CM [92]. Circulating levels of TNF-α
have been correlated with the risk of death from CM [93].
Anti-inflammatory cytokines have also been shown to play a critical role in
malaria pathogenesis. An infection by P. chabaudi chabaudi AS has been shown to be more
lethal in IL-10-deficient mice than in their wild type counterparts [94]. Similarly, in humans,
lower IL-10 levels were found in fatal cases than survivors [95]. African children suffering
20
from anaemia have been reported to have low serum levels of IL-10 [96]. In anaemic children,
the ratio of IL-10/TNF-α concentrations was linked to the severity of anaemia [97].
Thus, a delicate balance in levels and timing of pro- and anti-inflammatory
responses may determine the degree of parasitemia, severity and outcome of the disease.
Manijeh Vafa Homann
21
Human genetic factors involved in immunity to Plasmodium falciparum infection
RELATED BACKGROUND
Human genetic factors and malaria immunity
Development of clinical immunity may be influenced by age, pregnancy,
nutritional status, co-infections with other pathogens and host genetic factors [98]. The
importance of human genetic factors in immunity to malaria has long been suggested by the
high frequency of some genetic disorders in malaria endemic areas, which despite general
disadvantages, confer protection against malaria.
RBC variants
RBCs are crucial in the life cycle of the malaria parasite. The parasite uses the
RBC as shelter and food resources. Moreover, the interaction between infected RBCs and
uninfected RBCs and other tissues results in pathogenesis of the disease. Therefore, changes
in RBC structure or function may interfere with RBC invasion by the parasite and/or parasite
growth and multiplication inside RBCs, and consequently affect the outcome of the disease.
Following, some RBC variants important in malaria immunity and included in this study are
listed.
The ABO blood group system
Polymorphisms in carbohydrate structure of glycoproteins and glycolipids on
the surface of erythrocytes generate the ABO system, which is the most medically important
RBC polymorphism. Studies on associations between the ABO system and malaria, have
reported contradictory results. Some have shown a correlation between the ABO system and
severity of the disease, while others reported no correlations between malaria disease and the
ABO blood groups [99, 100]. However, blood groups A, B and AB have been shown to form
rosettes (binding of iRBCs to uninfected ones and to endothelial cells, which is linked to
severe pathology caused by P. falciparum), more readily than blood group O [101]. Thus,
individuals with blood group A, B and AB are more likely to develop CM compared to
22
individuals with blood group O [102]. In malaria endemic areas, the ratio of blood group O to
non-O is greater than 1, whereas this ratio is smaller than 1 in other populations. This has
been suggested to be the result from the selection of blood group O by P. falciparum
infections. However, no study has linked the ABO system to prevalence of infection,
frequency of repeated malaria episodes or anti-malaria antibody levels [99, 100].
The haemoglobinopathies
Haemoglobinopathies, disorders of haemoglobin (Hb) structure and production,
are the most common human genetic disorders. Hb is the iron-containing protein found in
RBCs that transports oxygen from the lungs to the rest of the body. Hb molecules consist of
four polypeptide chains named globins (usually 2 alpha and 2 beta chains) and an ironcontaining molecule, termed haem, attached to each chain [103].
The haemoglobinopathies are classified into two categories. One caused by
variations in Hb structure such as HbS, C and E, and the other results from variable
production of normal α- or β-globin, causing α- and β- thalassemia, respectively [104]. The
high prevalence of these disorders in malaria-endemic areas, are considered as being the result
of selection under malaria pressure [105].
Sickle cell disease
Sickle cell disease results from a point mutation in the β-globin subunit of
haemoglobin. Haemoglobin S is formed when two wild-type α-globin subunits are associated
with two mutant β-globins. Normal RBCs contain HbA and are quite elastic, which enables
the cells to deform in order to pass through capillaries. HbS carrying RBCs are distorted, tend
to lose their elasticity, and have shorter life time as compared to the normal ones. The
homozygous form of HbS (HbSS) causes a severe disease and is lethal at early childhood, in
the absence of adequate treatment. The heterozygous state (HbAS), also called sickle-cell
trait, has long been linked to protection against severe malaria [106, 107].
A number of mechanisms, by which sickle-cell disease may mediate resistance
to malaria, has been proposed in several studies [108]. In vitro studies have indicated:
enhanced phagocytosis of ring-stage parasitized HbAS erythrocytes [109], reduced
Manijeh Vafa Homann
23
Human genetic factors involved in immunity to Plasmodium falciparum infection
erythrocyte invasion and impaired intra-erythrocytic parasite growth in AS-RBCs, as
compared to infected normal erythrocytes [110, 111]. HbAS parasitized erythrocytes also tend
to sickle [110], which may enhance removal of premature infected erythrocytes in the spleen
[110, 112]. Enhanced immune responses to parasitized HbAS erythrocytes have also been
reported [113, 114]. A recent study has shown significantly reduced binding of P. falciparum
infected AS erythrocytes to micro-vascular-endothelial cells and blood monocytes as
compared to infected normal RBCs [115]. This may take part in protection to severe malaria
conferred by the sickle cell trait. HbAS has not been shown to influence asymptomatic
parasitaemia, however, it protects against severe and fatal malaria [107]. Thus, the precise
mechanism whereby sickle cell disease protects against malaria is yet unclear.
Alpha-thalassemia
The haemoglobin-alpha-chain gene is located on chromosome 16. Normally,
there are 2 alpha-chain genes on each chromosome, 4 in total. Alpha-chain-protein production
is evenly divided among the four genes. Alpha-thalassemia occurs when one or more genes
are mutated or fail to work. The severity and type of disease depend on the number of affected
genes, ranging from 4 genes deleted, to one deleted gene. The first defect is fatal prior to or
shortly after birth, whereas the latter form is clinically silent (α+) [103]. Depending on the
size of the deletion, α+ is categorized into different types [116]. In Africa, -α3.7 is the most
prevalent α+–thalassemia [105].
Several studies have indicated that α-thalassemia confers significant protection
against severe and fatal malaria [117, 118]. It has also been related to protection against mild
malaria [118, 119]. However, α-thalassemia does not protect against asymptomatic
parasiteamia [104], and does not influence parasite densities in individuals that go on to
develop the disease [117-119]. Studies by Carlson et al. [120] proposed that α-thalassemic
individuals are less likely to develop severe malaria, probably due to impaired ability of αthalassemic RBCs to form rosettes. It has been shown that α-thalassemia gives rise to reduced
expression of the red-cell complement receptor 1 (CR1), an essential receptor for rosetting
[121]. Therefore, impaired CR1-mediated rosetting may confer resistance to severe malaria
seen in α-thalassemic individuals.
24
G6PD deficiency
Glucose-6-phosphate dehydrogenase (G6PD) is a housekeeping enzyme that
functions as an antioxidant, and thus protects the cell from oxidative damage. Although the
lack of this enzyme in RBCs is fatal, mutant forms with defects in function are common. The
G6PD gene is located on the long arm of the X chromosome. Four hundred and forty two
G6PD variants have been identified. More than 130 variants were found at the DNA level that
lead to impaired enzyme activity [122].
Malaria parasites break down haemoglobin inside the RBC. This gives rise to
products such as oxidized iron, which are toxic to the parasite. Any defects in G6PD function,
which is important for cells to overcome the oxidative stress, is lethal for the parasite and can
thus confer resistance to malaria infections [122]. The geographic distribution of G6PD
variants matches with the historical distribution of malaria disease [123, 124]. The mutant
G6PD A- with 10-50 percent of normal enzyme activity, is the most common G6PD
deficiency allele in sub-Saharan Africa [125-127]. In two large case-control studies, Ruwende
et al. have shown that G6PD A- causes a 46% reduction in the risk of developing severe
malaria in heterozygous females, and 58% in hemizygous males [125]. Studies by Roth et al.
revealed that parasitemia levels in cultures with G6PD-deficient RBCs (from hemizygous
males and heterozygous females) were three times lower than those cultured in normal blood
(blood from individuals with normal RBC) [128]. It has also been shown that ring-stage
infected G6PD deficient RBCs are phagocytosed 2.3 times more intensely than infected
normal RBCs [129]. However, the actual mechanism(s) whereby G6PD deficiencies protect
against malaria is still not well understood.
Others
Duffy and Gerbich blood-group antigen deficiencies, HbC, HbE, β-thalassemia,
CR1 polymorphism and Southeast Asian Ovalocytosis (SAO) are other examples of RBC
variants, involved in malaria resistance [104].
Human leukocyte antigen (HLA)
The MHC gene loci in humans are referred to the HLA complex and are highly
polymorphic. There are evidences that specific HLA alleles might be correlated to
Manijeh Vafa Homann
25
Human genetic factors involved in immunity to Plasmodium falciparum infection
susceptibility/resistance to malaria disease. For instance, studies in The Gambia have
correlated an HLA class I antigen, HLA-B*5301, and an HLA class II haplotype, HLADRB1*1302-DQB1*0501, with resistance to severe malaria [130]. HLA-DQB1*0501 has
been linked to the reduced risk of re-infection and anaemia in Gabonese children [131], where
DRB1*04 and DPB1*1701 were related to severe disease [132]. A case-control study in
Ghana has revealed an association between HLA-DRB1*04 and severe malaria [133].
Ethnicity
A recent study in Kenya has estimated that almost 25% of the variation in
malaria incidence is a consequence of host genetic factors. This ratio was far beyond the
effects expected for the haemoglobinopathies alone. The HbS gene, which is the strongest
known resistance genetic factor, constitutes only 2% of the total variation. This suggests that a
large number of protective genes exits in the population, each of which taking a small part in
the protection [134].
Ethnic differences in susceptibility to malaria offer an important approach to
identify immune genes behind the resistance or susceptibility to malaria. Studies in Mali
[135], Burkina Faso [136]and Sudan [137] have revealed a significant inter-ethnic variation in
susceptibility to malaria. In spite of exposure to the same parasite and living under the same
transmission intensity, the Fulani ethnic group less frequently develop clinical malaria, as
compared to their sympatric ethnic groups (the Dogon in Mali, the Mossi and Rimaibe in
Burkina Faso and the Masaleit in Sudan). The Fulani have also been reported to have higher
anti-malarial-immune reactivity and lower asymptomatic infection prevalence than their
neighbours [135, 136, 138-142]. This relative resistance to malaria has been shown not to be
correlated to classical malaria-resistance genes, since higher frequencies of α-thalassaemia,
G6PD, HbC and HLA-B*5301 were found in the Mossi and Rimaibe as compared to the
Fulani. Similar frequencies of HbS were observed amongst three sympatric groups [143].
Similarly in Mali, HbC was more frequent in the Dogon group than in the Fulani, and HbS
showed a low similar frequency in both communities [135]. A recent study in Sudan has
revealed a significantly higher frequency of HbAS in non-Fulani (Masaleit) compared to the
Fulani [144]. Higher frequencies of spleen enlargement have been reported in the Fulani than
non-Fulani groups [135, 137, 145-147].
26
A recent study in Burkina Faso has revealed enhanced expression of both Th1(IL-18 and IFN-γ) and Th2 (IL-4 and GATA 3) related genes, in peripheral blood
mononuclear cells (PBMC) from the Fulani compared to the Mossi. However, the Fulani had
a reduced expression of Treg determinant genes (Foxp3 and CTLA4), suggesting that the
defect in functional Treg in the Fulani could reflect their higher resistance to malaria [148].
These data are in agreement with previous reports from Mali, showing a more efficient Th1
and Th2 response towards P. falciparum antigens in the Fulani [141]. Thus, it is obvious that
many different factors are involved in the protection against malaria, and these factors might
differ in different populations.
FcγR gene polymorphisms
Three families of human FcγRs have been identified, FcγRI (CD64), FcγRII
(CD32) and FcγRIII (CD16). Each FcγR family constitutes of several isoforms. A-, B- and C
isoforms have been described for FcγR- I and II, whereas FcγRIII includes only two isoforms,
A and B. The FcγR isoforms are expressed on a variety of cell types and vary in their binding
affinity to different IgG subclasses.
A number of studies has related some FcγRs gene polymorphisms (for instance
FcγRIIA-G494A [149-151], FcγRIIB-T695C [152] and FcγRIIIB-NA1/NA2 [149]) to
susceptibility to and/or severity of malaria disease. The G494A SNP in the FcγR IIA gene
gives rise to amino acid exchange from arginine (R) to histidine (H) at position 131. FcγR
IIA-131H is the only human FcγR which efficiently binds the IgG2 subclass [153]. The H
allele has been correlated to both protection and susceptibility to malaria in different studies
(reviwed by Braga et al. [154]). Studies in Sudan [144] and Mali [Israelsson et al.,
unpublished] have shown significantly higher frequencies of the H allele and the H/H
genotype in the Fulani as compared to the non-Fulani. However, whether or not the FcγR IIA131H allele plays any role in reducing the susceptibility to malaria disease in the Fulani, is not
clear yet.
Manijeh Vafa Homann
27
Human genetic factors involved in immunity to Plasmodium falciparum infection
Cytokine gene polymorphisms
Accumulating evidence suggests that inter-individual variations in the
production of cytokines involved in malaria immunity may influence the course of infection,
and determine the outcome of the disease. It is also evident that the ability to produce variable
amounts of cytokines can be genetically regulated. A number of polymorphisms in cytokinerelated genes that modifies cytokine expression and/or function has been identified. For
example, replacement of a C by a T at -590 position in the IL-4 gene promoter has been
described to result in enhanced IL-4 production [155]. Similarly, a G to A substitution in the
IL-10 gene promoter, position -1087, has been identified and related to elevated levels of IL10 production [156].
Quite a few studies have tried to relate polymorphisms in cytokine-related genes
to various infections/diseases, including malaria. For instance, enhanced levels of TNF-α in
cerebral malaria patients have been shown to be a consequence of a variation in the TNF-α
promoter in these individuals [157-159]. In Thai malaria patients, the IL-13 -1055 T allele has
been shown to protect against severe malaria [160]. The IL-4 -590 T allele has been correlated
to elevated IgE levels in asthmatic children [155]. This allele has also been associated with
elevated total IgE levels in children suffering from severe malaria. However, the frequency of
the T allele did not differ between severe and mild malaria cases [161, 162]. In Burkina Faso,
the IL-4 -590 T allele was associated with increased malaria-specific IgG levels within the
Fulani individuals but not in non-Fulani [163]. Polymorphisms in interferon-regulatory factor
(IRF)-1 have recently been linked to the control of P. falciparum infection in the same study
area [164]. However, none of the studied polymorphisms has yet been suggested as a
determinant of susceptibility/resistance to malaria infection/disease.
Genetic diversity of malaria parasite
The success of P. falciparum in evading host immunity, particularly in avoiding
antibody-mediated immunity against parasite-derived surface proteins, is attributed to the
antigenic diversity in the parasite population as a whole, as well as antigenic variation in the
individual parasite. The malaria parasite is able to switch the expression of surface antigens
and, thus, change the antigenicity during the course of infection [165]. The existence of
phenotypically variable strains of the parasite in endemic areas enables the P. falciparum
28
parasite to re-infect previously exposed individuals. This leads to strain and variant specific
immunity [166, 167]. Human experimental malaria studies have shown that a primary
infection with a certain strain generates protective immunity against the infecting strain but
not against infections by other strains [168]. This has been suggested to be the reason why
long term exposure (years), to a pool of variable parasite strains, is required for the
development of naturally acquired immunity to the Plasmodium parasite [169]. Thus, the
acquired immunity has been suggested to develop after infection by multiple parasite strains.
Multiplicity of malaria infection
MOI is the number of various P. falciparum strains simultaneously infecting a
single person. Individuals residing in malaria endemic areas usually carry multiple parasite
strains. MOI can be an indicator of transmission intensity and immune status. MOI has been
shown to be well related to transmission intensity [170, 171]. The presence of diverse
infections in asymptomatic inhabitants in endemic areas is suggested to reflect acquired
immunity [172], and thus may affect the outcome of the infection. Some studies have
indicated that the increase in MOI decreases the risk of developing clinical malaria [173-175],
whereas others indicated a positive link between malaria disease and MOI [176, 177]. Patients
suffering from cerebral malaria have been shown to have higher MOI than mild malaria
patients. However, a slight increase in clone numbers was found in healthy controls compared
to patients with mild malaria [178]. In contrast to this, a recent study in Nigeria has revealed
that children with single strain infection are at the higher risk to develop cerebral malaria as
compared to those with multiple infections [179]. Thus, the impact of MOI on the
development of malaria immunity is still unclear.
MOI is suggested to vary over age, and this variation is highly influenced by
transmission intensity [170, 180-182]. MOI seems to increase as immunity develops [172].
The age of acquiring natural immunity differs according to transmission intensity [183]. In
areas with intensive malaria transmission, immunity is acquired at younger ages compared to
less intense transmission areas [183]. However, the relation between infection by multiple
parasite strains and malaria disease is still unclear.
The number of msp2 variants is widely used as a marker to determine MOI
[184].
Manijeh Vafa Homann
29
Human genetic factors involved in immunity to Plasmodium falciparum infection
MSP2
MSP2, also called MSA2, is a 45 kDa polymorphic protein anchored in the
merozoite membrane. The msp2 gene is a single-copy gene and consists of five blocks,
encoding conserved, dimorphic and polymorphic sequences (Fig. 3). Blocks 1 and 5 encode
highly conserved N (43 residues) and C (74 residues) termini, respectively. Blocks 2 and 4 are
dimorphic and classified into two families, the 3D7 and FC27 (also known as type A and type
B, respectively). The central region, block 3, consists of a series of variable repeat motifs,
which vary in number, length and sequence among isolates. Individual MSP2 alleles are
defined by the central repeats [185-187]. MSP-2 is involved in RBC invasion by merozoites
[188].
FC27
3D7
Conserved regions
Family specific non-repetetive regions
Allelic specific repetetive regions
Figure 3: Modified from Damon Eisen [189]
30
PRESENT STUDY
Aim of study
The overall aim of this study was to investigate possible associations between
human genetic factors and malariometric indexes, with the following specific objectives:
•
To investigate correlations between IL-4 -590 C/T and IL-10 -1087 A/G SNPs, and
inter-ethnic differences in malariometric indexes observed between the Fulani and the
Dogon, living in Mali.
•
To correlate the levels of malaria-specific IgG, IgG1-4, IgE, as well as total IgE, to
prevalence and multiplicity of asymptomatic infection and splenomegaly in the Fulani,
and to compare the results to those observed in the Dogon.
•
To investigate the influence of IL-4 -590 C/T SNP on anti-malarial IgG, IgG1-4, IgE
and total IgE levels in the Fulani and the Dogon.
•
To study: the impact of erythrocyte variants on MOI; correlation between MOI and
the risk of clinical malaria; and associations between MOI and age, parasite density
and seasonal variation, in asymptomatic children in Senegal.
Manijeh Vafa Homann
31
Human genetic factors involved in immunity to Plasmodium falciparum infection
Methodology
Materials and methods, briefly mentioned below, are described in details in the
respective papers (manuscripts).
Paper I
Age-matched asymptomatic individuals of the Fulani (214) and the Dogon
(212), were included in this study. Finger-prick blood was collected on filter papers. Human
and parasite DNA were extracted using the Chelex- and TE buffer-based methods,
respectively. Polymerase-chain reaction (PCR) was performed for both human (IL-4 T/C and
IL-10 A/G SNPs) and parasite (msp2/3D7 and msp2/FC27) genotyping.
Paper II
Parasite DNA was extracted from blood on filter papers using Chelex-100.
Number of concurrent parasite clones was defined by msp2 genotyping using PCR. ELISA
was performed to determine malaria-specific IgG, IgG1-4, IgE as well as total IgE antibody
levels.
Paper III
Human genomic DNA was extracted from blood on filter papers using the
Chelex-100 and genotyped for the IL-4 T/C SNP by PCR. Antibody levels were determined
using ELISA.
Paper IV
Asymptomatic children (2 -10 years old), living in Senegal, were included in
this study. Parasite densities were measured in June 2002 and January 2003. An active survey
of malaria attacks was conducted during the transmission period in 2002 for a sub-group of
154 children. ABO blood groups were determined by serology, and the other red-blood-cell
polymorphisms were detected using genomic DNA. Parasite DNA was extracted from blood
on filter papers and analysed for the number of msp2 (FC27 and 3D7) clones per infection by
PCR.
32
Results and discussions
Paper I
IL-4 and IL-10 are two crucial cytokines in immune responses in general and
anti-malarial responses in particular [190-193]. IL-4-590 T/C and IL-10 -1087 A/G
polymorphisms have been shown to result in variable production of these cytokines [155,
156]. We investigated the associations between the above mentioned SNPs and malariometric
indexes amongst sympatric ethnic groups living in Mali, the Fulani and the Dogon.
We observed that the allele and genotype frequencies of the IL-4 -590 T/C and
IL-10 -1087 A/G SNPs varied significantly between the Fulani and the Dogon, indicating that
these ethnic groups had distinct genetic backgrounds. This was consistent with studies in
Burkina Faso, where HLA class I markers in the Fulani were found to be genetically very
distant from their neighbours [194]. The Fulani from Mali (present study) and Burkina Faso
[163] represented similar allele and genotype frequencies of IL-4 -590 T/C, supporting the
same origin for the Fulani in the two countries [194].
In Burkina Faso, the Fulani T allele carriers had higher anti-malarial IgG levels
compared to the CC genotype carriers. This was not seen in the non-Fulani groups living in
the same area, suggesting that the IL-4 -590 polymorphism is not functional, but may be
linked to other factor(s), which can be the real determinant of resistance to malaria [163]. We
found that the IL-4 -590 T allele was correlated to the presence of P. falciparum infection in
the Fulani group but not in the Dogon. This finding suggests that the persistence of infection
in the T allele carriers gives rise to constitutive anti-malarial antibody production.
Pronounced differences in the IL-4 -590 allele and genotype frequencies were
observed between the non-Fulani groups living in the two countries (recent study and [163]).
The Dogon carried a T/C allelic ratio of 82/18, which was very different from those reported
for Mossi and Rimaibe (18/82 and 25/75, respectively). These data suggest that unknown
factor(s) linked to this position, may be polymorphic in non-Fulani from different areas, and
that those factors have been selected differently under variable selection pressures, including
malaria.
Within the Fulani, the T allele carriers had higher infection prevalence than
those carrying the CC genotype. Although parasite densities did not vary significantly
between the two groups, the Fulani CC genotype carriers had lower parasite loads compared
Manijeh Vafa Homann
33
Human genetic factors involved in immunity to Plasmodium falciparum infection
to the Dogon carrying the same genotype. This suggests that the Fulani CC genotype carriers
may succeed better in clearing the infection, compared to the T allele carriers, and to the
Dogon CC genotype carriers. However, to draw a firm conclusion regarding the correlations
between IL-4 genotypes and parasitemia, parasite counts need to be more accurately
determined in consecutive samplings.
In line with former studies [135, 141, 145, 146], we noted lower frequencies of
infection and higher spleen rates in the Fulani group than in the Dogon. Within the Fulani,
spleen enlargement was associated with the presence of P. falciparum infection. This was not
the case in the Dogon group. Our results contradict those reported by Bereczky et al. [195],
where splenomegaly was associated with P. falciparum infections within the Dogon but not
the Fulani. These data underline that more standardized analyses are needed to evaluate hostparasite interactions.
No significant inter- and intra-ethnic differences, in the studied malariometric
indexes, between different IL-10 genotypes were seen in this study.
All together, these data suggest that unknown factor(s) is probably linked to the
IL-4 -590 C and is polymorphic in the non-Fulani. This factor(s) might be involved in the
control of infection and thus, avoiding the clinical presentation. However, further studies are
required to better clarify the contribution of IL-4 and particularly the studied SNP in malaria
immunity.
Paper II
Antibodies play important roles in clearance of blood-stage parasites [196]. In
this study, we assessed correlations between anti-malarial antibody levels and malariometric
characteristics in asymptomatic individuals of the Fulani and the Dogon.
Higher anti-malarial IgG, IgG1-3, IgE and total IgE levels were detected within
the Fulani as compared to the Dogon. However, no significant correlations between antibody
levels and parasite prevalence, number of infective clones and spleen enlargement were noted.
The higher spleen- and lower infection rates, seen in the Fulani compared to
their neighbours, have been proposed to be the result from their enhanced anti-malarial
reactivity. Interestingly, our results revealed no significant correlation between malariaspecific antibody levels and the presence of P. falciparum infection. This suggests that the
relative resistance to malaria infection seen in the Fulani group is not mainly dependent on the
34
magnitude of anti-malarial humoral response. In support of this idea, studies in Sudan have
shown that the Fulani had lower malaria-specific IgG1 and almost equal IgG and IgG3 levels
compared to their neighbours, the Masalite (Nasr et al., unpublished).
In this study, we noted that amongst the Fulani individuals, the increase in total
IgE levels was weakly related to the presence of malaria parasites. In malaria patients,
elevated total IgE levels have been associated with the development of CM [74]. Other
studies have shown that unlike malaria-specific IgE [75, 76], total IgE levels did not influence
the risk of subsequent clinical attacks [75]. Thus, the actual consequence of the increase in
total IgE levels during malaria infection within the Fulani remains to be clarified.
Correlations between the studied antibody levels and clone numbers were then
analysed. An increase in Log10 values of malaria-specific IgG4 levels within the Fulani was
linked to a decrease in clone numbers. ADCI, which is suggested to be the most efficient antimalarial mechanisms mediated by protective antibodies, is believed not to be mediated by
IgG4 antibodies [72]. However, IgG4 might still display other anti-parasitic functions, such as
inhibition of cytoadherence or inhibition of RBC invasion by the parasite. IgG4, raised
against parasite antigens involved in cytoadherence, may help parasite clearance in the spleen
through inhibiting sequestration of iRBCs. By binding to parasite antigens involved in
parasite invasion, IgG4 by itself could interrupt invasion of new RBCs. Thus, the specificity,
function and/or affinity of an antibody might be more important than the amount of the
antibody. We observed no significant differences in IgG4 levels between the two studied
groups. In the Fulani, IgG4 levels were negatively associated with the number of clones.
Taken together, our results suggest that IgG4, in the two groups, may be raised against
different parasite epitopes or have different affinities, and thus mediate various anti-parasitic
functions.
Spleen enlargement is mostly seen in individuals at the age of 15 and younger
(children). Thus, the relation between antibody levels and splenomegaly was evaluated in the
whole study population and separately in children. A weak correlation between elevated IgG
and IgG3 levels and splenomegaly was noted in the Fulani group (in all ages included and in
children only). Despite higher levels of IgG1 compared to IgG3, splenomegaly was not linked
to IgG1 levels. This may be reflected by different specificities of IgG1 and IgG3. That is,
IgG3 may identify antigens which may lead to less sequestration and thus, to more efficient
parasite clearance by the spleen. However, our observation might also be the result from
overestimation of significances due to not correcting the P values for multiple analyses.
Manijeh Vafa Homann
35
Human genetic factors involved in immunity to Plasmodium falciparum infection
Trends for correlations between IgG2 levels and splenomegaly, and multiplicity
of infection within the Dogon were noted in this study. These results could not be explained
by the frequency of the FcγRIIA-131H variant, which enables IgG2 to act as cytophilic
antibody in parallel with IgG1 and IgG3 [73], since the H allele is less frequent than the R
allele in the Dogon (Israelsson et al., unpublished). This emphasizes the need for further
studies to investigate the actual role of IgG2 in responses to malaria.
All together, our results suggest that the stronger anti-malarial humoral response
seen in the Fulani is not responsible for the differences in malariometric indexes observed
between the Fulani and the non-Fulani ethnic groups. A rapid, efficient and well-regulated
innate immune response, which controls parasite counts before the acquired immune response
has developed, may be the reason for the relative resistance to malaria seen in the Fulani.
Paper III
The impact of IL-4 -590 T/C SNP on the level of malaria-specific IgG, IgG1-4,
IgE and total IgE levels, in asymptomatic individuals of the Fulani and the Dogon, was
analysed in this study. We noted that amongst the Fulani, the increases in total and antimalarial IgE levels were associated with the IL-4 T allele after adjustment for age and sex. No
correlation between IL-4 -590 variants and other studied antibodies were seen in any of the
studied groups.
IgE has been suggested to be involved in both pathogenesis and protection
against malaria disease. Studies have shown higher levels of total and malaria-specific IgE in
patients suffering from severe malaria compared to those with uncomplicated malaria [74,
197]. However, in asymptomatic individuals, enhanced levels of anti-malarial IgE decreased
the risk of developing clinical symptoms [75, 76]. This suggests that the role of IgE in malaria
is still unclear. In our study, we noted that total and malaria-specific IgE levels were higher in
the T allele carriers compared to the CC genotype, only within the Fulani. Our former studies
have shown a higher frequency of P. falciparum infection in the Fulani T allele carriers than
those with the CC genotype (Paper I), and that the elevated levels of total IgE were correlated
to the presence of infection (Paper II). Others have shown that the T allele is correlated to
elevated levels of total IgE amongst children with severe malaria whereas the frequencies of
the T allele did not differ significantly between severe and mild malaria cases [161, 162].
Therefore, whether the increase in the levels of IgE, seen in the T allele carriers in this study,
36
is a direct consequence of the impact of IL-4 -590 variants on the antibody-class switching, or
if it is due to the higher infection frequencies observed in the T allele carriers, is not clear and
remains to be investigated.
All together, these data suggest that IL-4 and IgE might influence the outcome
of the disease, in concert with other factors (genetic, epigenetic, epistatic or environmental). A
recent study in Burkina Faso has revealed associations between a number of SNPs in the IRF1 gene and carriage of P. falciparum infection in both the Fulani and the Mossi. However,
different patterns of polymorphisms were seen in the two groups, probably reflecting their
different haplotypic architecture [164]. IRF-1 elicits the expression of IL-12p 40 [198, 199]
and is critical for production of IL-12 [200], and subsequently for initiation of the Th1 type of
responses [201, 202]. Thus, IRF-1 can modulate the balance between Th1 and Th2 type of
responses, which is important for the IgE response in malaria [141].
Studies in Burkina Faso have revealed an association between the IL-4 -590 T
allele and elevated levels of anti-malarial IgG, in the Fulani group [163]. In contrast to this,
we noted that IgG levels did not vary significantly between IL-4 genotypes in any of the
studied groups. Production of IL-4 and any other factor(s) involved in the regulation of
antibody responses could also be affected by co-infections with other parasitic or bacterial
pathogens or time after last exposure to malaria. This idea is supported by findings from other
studies, where reverse or no relations between IL-4 -590 variants and IgG levels were seen
[203, 204], indicating that the impact of IL-4 variants on IgG production, in turn, is
influenced/regulated by other factors.
Correlations between IL-4 -590 variants on the level of IgG subclasses have not
been studied previously. IL-4 and IL-13 have been shown to promote IgG4 production [32,
33]. In this study, a trend for an increase in IgG4 levels in the T allele carriers was observed
only within the Fulani individuals. However, analysis of IgG4 responses is difficult due to the
generally low levels of these antibodies. Therefore, the impact of IL-4 -590 variants on IgG4
responses needs to be more analysed including larger study groups.
No significant variations in levels of malaria-specific IgG1-3 between IL-4 -590
genotypes were seen in any of the studied groups.
Taken together, our data suggest that IL-4 may play other roles in malaria
immunity in addition to the regulation of antibody responses. The implication of the IL-4 -590
T allele in susceptibility and/or resistance to malaria infection/disease within the Fulani
Manijeh Vafa Homann
37
Human genetic factors involved in immunity to Plasmodium falciparum infection
remains to be addressed in follow-up and case-control studies. Furthermore, the actual role of
IgE in protection and pathogenesis of malaria remains to be uncovered.
Paper IV
This study examined the impact of erythrocyte variants on MOI. Relations
between MOI and age, seasonal variations as well as malaria morbidity were also analysed.
MOI was measured before and after transmission season in asymptomatic children, in the age
range of 2 to 10 years, living in a malaria mesoendemic area in Senegal.
As expected, parasite density and MOI increased after the transmission season.
MOI was positively correlated to parasite density, which is in concordance with data reported
by others showing that with higher parasite counts, it is more probable to detect concurrent
clones [205, 206].
The impact of different RBC variants on MOI has been limitedly studied. We
noted that the general increase in MOI after the transmission season was not seen in αthalassaemic and G6PD mutated children. Alpha-thalassaemia alters the membrane properties
of RBCs [207, 208], which may affect their invasion by merozoites. MSP2 is a polymorphic
antigen involved in invasion [188]. Thus, the altered membrane of α-thalassaemic RBCs may
impair the invasion by certain MSP2 variants and consequently protect against infection by
certain parasite strains. However, other studies have shown no impact of α-thalassaemia on
parasite prevalence, density and diversity in symptomatic individuals [209]. A reason for this
contradiction might be due to that we studied asymptomatic subjects, whereas the former
study included symptomatic cases. It is known that α -thalassaemia does not affect parasite
loads in individuals who develop clinical disease [118, 119, 209]. The differences in sample
sizes in the two studies could also be considered as a reason for this discrepancy. These data
suggest that α-thalassaemia may protect against infection by certain parasite strains, but does
not protect against clinical malaria in ongoing infections.
Infected G6PD-deficient RBCs have been shown to be more efficiently
phagocytosed than infected normal RBCs [129]. In this study, MOI did not increase after the
transmission season in the G6PD-mutated children, which might be due to rapid parasite
clearance in these children as compared to those carrying the normal G6PD gene. This effect
was stronger in G6PD-mutated girls carrying the HbAS gene. However HbAS alone did not
38
affect MOI. G6PD deficiency did not influence MOI in male subjects, which might be due to
the small size of this group.
In the present study, the ABO blood system did not show any influence on MOI
at any time points. The ABO system has been implicated in the outcome of the disease, where
blood group O is less likely to develop CM as compared to A, B and AB carriers [101].
However, the ABO system did not seem to affect parasite density and diversity in
asymptomatic individuals.
Sickle cell trait did not influence MOI in our study. Former studies have
reported none [210] or reverse [211] relations between HbAS and MOI. The reason for these
discrepancies is not known, but they might be due to differences in the ages of the studied
populations and/or variations in transmission intensities between the study areas.
MOI has been shown to be age-dependent in areas with intense transmission,
and to be independent of age in a malaria mesoendemic area [210, 212]. In line with the latter
report, MOI was not influenced by age in the present study. However, these findings need to
be confirmed in a study population with a broader age range.
A clinical survey was performed including 154 children to evaluate correlations
between MOI and clinical attacks. Forty five percent of these children had at least one clinical
attack during the survey. No significant association between clinical malaria and MOI was
noted, at any time points. Both reverse [173-175] and positive [176, 177] correlations between
MOI and malaria morbidity have been reported. Our finding does not agree with any of those
reports, which may partly be due to different genotyping methods used in these studies.
Differences in sensitivity and accuracy of genotyping protocols have been shown to give rise
to variations in results from different laboratories [206, 210, 213, 214]. Moreover, in the
present study, sampling was not done at the time of clinical presentation. This may also be a
reason why MOI was not related to morbidity. In any case, these results emphasize that
malaria morbidity is not only affected by parasitological indexes, but could also be influenced
by host genetic factors and/or transmission levels.
Taken together, methodology should be considered when comparing results
from different studies on MOI. The results from PCR may vary if different protocols and
reagents are used or even may differ from hand to hand [206]. PCR may underestimate the
MOI. Importantly, PCR is unable to differenciate between virulent and avirulent strains. Thus,
to draw a firm conclusion from studies on MOI in different areas, a more standardized
methodology is required.
Manijeh Vafa Homann
39
Human genetic factors involved in immunity to Plasmodium falciparum infection
CONCLUDING REMARKS AND FUTURE PERSPECTIVES
We conclude from our studies that the Fulani and the Dogon are genetically
markedly different, using IL-4 -590 and IL-10 -1087 variants, as markers. The IL-4 genotype
pattern was similar in the Fulani from different areas, whereas it varied significantly between
non-Fulani groups living in different countries. The IL-4 T allele was previously suggested to
be linked to unknown factor(s), which could be the real determinant of resistance to malaria
[163]. Our results suggest that the unknown factor is probably linked to the C allele and is
polymorphic in non-Fulani. This factor might be involved in a better parasite clearance.
This study concludes that the magnitude of anti-malarial IgG response may not
be as important as it is thought to be. Instead, other features of antibodies such as specificity,
function and/or affinity may be more crucial in protective response to malaria. The
implications of IgE antibodies in malaria protection and pathogenesis remain to be better
clarified. IgG4 may not mediate ADCI, but might still have some roles in malaria protection,
through mediating anti-parasitic functions different from those mediated by IgG1 and IgG3.
Our results also indicate that the relative resistance to malaria infection seen in the Fulani
does not rely mainly on their augmented anti-malarial antibody response, as it has been
believed to be. Rather, other factors such as a rapid and well-regulated innate response, which
can control the infection and consequently avoid the clinical presentations, might also take
part in this relative resistance to malaria.
Studies on the associations between IL-4 -590 variants and IgE have indicated
that relationship between IL-4 and IgE might be affected by other genetic/epigenetic/epistatic
or environmental factors. Therefore, the outcome of the disease could be reflected by the
action of these factors together with IL-4 and IgE. The production of IL-4 and other
cytokines/factors, in turn, could be affected by the time after last exposure to malaria or by
co-infection with other pathogens which might explain the contradictory results of the impact
of IL-4 -590 variants on IgG levels observed by a number of studies. All together, studies on
IL-4 (genotype/levels/-producing cells) in Mali and Burkina Faso suggest that the
contribution of IL-4 in inter-ethnic differences in malaria susceptibility is probably not
through its role in antibody regulation, rather owing to its implication in other aspects of
immune responses, e.g. induction of NK cell proliferation which help in parasite killing
and/or suppression of anti-inflammatory cytokine productions which avoids immunopathogenesis.
40
The last study concludes that in asymptomatic children; α-thalassaemic RBCs
may resist against infection by certain parasite strains; protection conferred by G6PD
deficiency might be due to rapid parasite clearance at early stage of infection; the ABO blood
system and HbAS do not affect the MOI. The multiplicity of infection was correlated to
parasite density in a malaria mesoendemic area, but did not vary over the age, in the range of
2 to 10 years. The risk of clinical attack was not affected by MOI. However, these data need
to be confirmed using different genotyping methods and including larger size of the study
group with broader age range.
To address the actual implications of IL-4 -590 variants in malaria (especially
inter-ethnic variations in malaria susceptibility), a follow-up study of a population with
known IL-4 -590 genotypes of the Fulani and the Dogon, joined with a case-control study, is
required. Consecutive determination of antibody levels together with parasitological data,
during these studies, might help to clarify the impact of IL-4 -590 variants on antibody
responses.
Most importantly, studying other genetic factors adjacent to IL-4 gene on
chromosome 5q may help in identifying the actual factor(s), involved in resistance/protection
to malaria infection. Sequencing of this region in individuals of the two groups would be of a
great help in this regard.
Manijeh Vafa Homann
41
Human genetic factors involved in immunity to Plasmodium falciparum infection
ACKNOWLEDGMENT
I left my family and home with the wish to continue my studies and become a researcher. It
was a long, tough journey to bring the wish close to reality; a long way that I would never
have been able to make on my own. I would, therefore, like to express my sincere gratitude to
all those who have helped and supported me during these years. In particular, I would like to
thank: Professor Marita Troye-Blomberg; for accepting me into your group, for your
support and guidance, for being patient with my endless questions, and for always being calm
and caring. I learnt so much from you. Present and former seniors at the Immunology
Department; Professor Klavs Berzins, Professor Carmen Fernandez, Professor Eva
Severinson, Docent Eva Sverremark Ekström and Docent Manuchehr Abedi-Valugerdi;
for being excellent teachers, for your critique and advice, and above all for providing a
family-like environment at the department. Margareta Hagstedt (Maggan) and Ann
Sjölund; for your excellent technical support and for your incredible speed in responding to
any call for help. Gelana Yadeta and Anna-Leena Jarva; for excellent administration,
always having a smile, and for your readiness to assist. Present and former students at the
Immunology Department, for making the department a nice and friendly place to work, coauthors, for your contributions, friends, in and out of university, for the very beautiful time
we shared and the Nadjafi family in Stockholm, for giving me the warmth and feelings of
home and family whilst in cold Stockholm.
I would also like to thank my dear siblings, Aboutaleb, Mahmood, Kobra, Azra, Mehri and
Monireh, and their families; your love and prayers have been my best company all these
years. What would my life mean without you guys?! Mehri, you are the best of friends and
sister, ever. You were available anytime I needed you, and filled every moment of my
loneliness after I left home; despite miles of physical distance our hearts and minds have been
tightly tied together moment by moment, I love you and miss you very much.
My love, my friend, my husband, Peter Homann; where would I have been if not for you
going through the whole emotional rollercoaster with me during the past months; for your
42
morning encouraging cuddles and evening relaxing hugs, for believing in me and bringing my
self-confidence back when I had none left.
And for my incomparable mother and father, Mohammad E. Vafa and Sara Madadi; words
are not enough to express my love and gratitude to you. Whatever I have, whatever I am,
whatever I get, it is all because of you. This thesis is in fact yours.
Manijeh Vafa Homann
43
Human genetic factors involved in immunity to Plasmodium falciparum infection
REFERENCES
[1] A. Pluddemann, C. Neyen, S. Gordon, Macrophage scavenger receptors and host-derived
ligands, Methods 43 (2007) 207-217.
[2] M. Krieger, The other side of scavenger receptors: pattern recognition for host defense,
Curr. Opin. Lipidol. 8 (1997) 275-280.
[3] T.D. Kanneganti, M. Lamkanfi, G. Nunez, Intracellular NOD-like receptors in host
defense and disease, Immunity 27 (2007) 549-559.
[4] C. Coban, K.J. Ishii, T. Horii, S. Akira, Manipulation of host innate immune responses by
the malaria parasite, Trends Microbiol. 15 (2007) 271-278.
[5] C. Ropert, B.S. Franklin, R.T. Gazzinelli, Role of TLRs/MyD88 in host resistance and
pathogenesis during protozoan infection: lessons from malaria, Semin. Immunopathol. 30
(2008) 41-51.
[6] M. Wykes, C. Keighley, A. Pinzon-Charry, M.F. Good, Dendritic cell biology during
malaria, Cell. Microbiol. 9 (2007) 300-305.
[7] C. Gregoire, L. Chasson, C. Luci, E. Tomasello, F. Geissmann, E. Vivier, T. Walzer, The
trafficking of natural killer cells, Immunol. Rev. 220 (2007) 169-182.
[8] J.R. Schoenborn, C.B. Wilson, Regulation of interferon-gamma during innate and
adaptive immune responses, Adv. Immunol. 96 (2007) 41-101.
[9] W.M. Yokoyama, B.F. Plougastel, Immune functions encoded by the natural killer gene
complex, Nat. Rev. Immunol. 3 (2003) 304-316.
[10] T.W. Mak, The T cell antigen receptor: "The Hunting of the Snark", Eur. J. Immunol. 37
Suppl 1 (2007) S83-93.
[11] M. Ciofani, J.C. Zuniga-Pflucker, The thymus as an inductive site for T lymphopoiesis,
Annu. Rev. Cell Dev. Biol. 23 (2007) 463-493.
44
[12] W.K. Born, N. Jin, M.K. Aydintug, J.M. Wands, J.D. French, C.L. Roark, R.L. O'Brien,
gammadelta T lymphocytes-selectable cells within the innate system? J. Clin. Immunol. 27
(2007) 133-144.
[13] S.S. Percival, J.F. Bukowski, J. Milner, Bioactive food components that enhance
gammadelta T cell function may play a role in cancer prevention, J. Nutr. 138 (2008) 1-4.
[14] J.H. Ruiz, I. Becker, CD8 cytotoxic T cells in cutaneous leishmaniasis, Parasite
Immunol. 29 (2007) 671-678.
[15] Z. Chen, J.J. O'Shea, Th17 cells: a new fate for differentiating helper T cells, Immunol.
Res. (2008) .
[16] C. Mottet, D. Golshayan, CD4+CD25+Foxp3+ regulatory T cells: from basic research to
potential therapeutic use, Swiss Med. Wkly. 137 (2007) 625-634.
[17] T. Egawa, G. Eberl, I. Taniuchi, K. Benlagha, F. Geissmann, L. Hennighausen, A.
Bendelac, D.R. Littman, Genetic evidence supporting selection of the Valpha14i NKT cell
lineage from double-positive thymocyte precursors, Immunity 22 (2005) 705-716.
[18] M. El Haj, A.B. Ya'acov, G. Lalazar, Y. Ilan, Potential role of NKT regulatory cell
ligands for the treatment of immune mediated colitis, World J. Gastroenterol. 13 (2007) 57995804.
[19] L.K. Poulsen, L. Hummelshoj, Triggers of IgE class switching and allergy development,
Ann. Med. 39 (2007) 440-456.
[20] J.N. Arnold, R.A. Dwek, P.M. Rudd, R.B. Sim, Mannan binding lectin and its interaction
with immunoglobulins in health and in disease, Immunol. Lett. 106 (2006) 103-110.
[21] J. Punnonen, R. de Waal Malefyt, P. van Vlasselaer, J.F. Gauchat, J.E. de Vries, IL-10
and viral IL-10 prevent IL-4-induced IgE synthesis by inhibiting the accessory cell function
of monocytes, J. Immunol. 151 (1993) 1280-1289.
[22] A. Cerutti, X. Qiao, B. He, Plasmacytoid dendritic cells and the regulation of
immunoglobulin heavy chain class switching, Immunol. Cell Biol. 83 (2005) 554-562.
Manijeh Vafa Homann
45
Human genetic factors involved in immunity to Plasmodium falciparum infection
[23] H.L. Spiegelberg, Biological role of different antibody classes, Int. Arch. Allergy Appl.
Immunol. 90 Suppl 1 (1989) 22-27.
[24] H.L. Spiegelberg, Biological activities of immunoglobulins of different classes and
subclasses, Adv. Immunol. 19 (1974) 259-294.
[25] F. Nimmerjahn, J.V. Ravetch, Fc-receptors as regulators of immunity, Adv. Immunol. 96
(2007) 179-204.
[26] F. Nimmerjahn, J.V. Ravetch, Fcgamma receptors as regulators of immune responses,
Nat. Rev. Immunol. 8 (2008) 34-47.
[27] S.J. Hopkins, The pathophysiological role of cytokines, Leg. Med. (Tokyo) 5 Suppl 1
(2003) S45-57.
[28] W.A. Verri Jr, T.M. Cunha, C.A. Parada, S. Poole, F.Q. Cunha, S.H. Ferreira,
Hypernociceptive role of cytokines and chemokines: targets for analgesic drug development?
Pharmacol. Ther. 112 (2006) 116-138.
[29] E. De Maeyer, J. De Maeyer-Guignard, Interferon-gamma, Curr. Opin. Immunol. 4
(1992) 321-326.
[30] P.R. Burd, W.C. Thompson, E.E. Max, F.C. Mills, Activated mast cells produce
interleukin 13, J. Exp. Med. 181 (1995) 1373-1380.
[31] W.E. Paul, Interleukin 4/B cell stimulatory factor 1: one lymphokine, many functions,
FASEB J. 1 (1987) 456-461.
[32] A. Ishizaka, Y. Sakiyama, M. Nakanishi, K. Tomizawa, E. Oshika, K. Kojima, Y.
Taguchi, E. Kandil, S. Matsumoto, The inductive effect of interleukin-4 on IgG4 and IgE
synthesis in human peripheral blood lymphocytes, Clin. Exp. Immunol. 79 (1990) 392-396.
[33] C.L. King, T.B. Nutman, IgE and IgG subclass regulation by IL-4 and IFN-gamma in
human helminth infections. Assessment by B cell precursor frequencies, J. Immunol. 151
(1993) 458-465.
[34] J.H. Jansen, W.E. Fibbe, R. Willemze, J.C. Kluin-Nelemans, Interleukin-4. A regulatory
protein, Blut 60 (1990) 269-274.
46
[35] P. Conti, D. Kempuraj, K. Kandere, M. Di Gioacchino, R.C. Barbacane, M.L. Castellani,
M. Felaco, W. Boucher, R. Letourneau, T.C. Theoharides, IL-10, an inflammatory/inhibitory
cytokine, but not always, Immunol. Lett. 86 (2003) 123-129.
[36] F. Rousset, E. Garcia, T. Defrance, C. Peronne, N. Vezzio, D.H. Hsu, R. Kastelein, K.W.
Moore, J. Banchereau, Interleukin 10 is a potent growth and differentiation factor for
activated human B lymphocytes, Proc. Natl. Acad. Sci. U. S. A. 89 (1992) 1890-1893.
[37] F. Rousset, S. Peyrol, E. Garcia, N. Vezzio, M. Andujar, J.A. Grimaud, J. Banchereau,
Long-term cultured CD40-activated B lymphocytes differentiate into plasma cells in response
to IL-10 but not IL-4, Int. Immunol. 7 (1995) 1243-1253.
[38] H. Spits, R. de Waal Malefyt, Functional characterization of human IL-10, Int. Arch.
Allergy Immunol. 99 (1992) 8-15.
[39] R. Tuteja, Malaria - an overview, FEBS J. 274 (2007) 4670-4679.
[40] D.A. Joy, X. Feng, J. Mu, T. Furuya, K. Chotivanich, A.U. Krettli, M. Ho, A. Wang, N.J.
White, E. Suh, P. Beerli, X.Z. Su, Early origin and recent expansion of Plasmodium
falciparum, Science 300 (2003) 318-321.
[41] R.W. Snow, C.A. Guerra, A.M. Noor, H.Y. Myint, S.I. Hay, The global distribution of
clinical episodes of Plasmodium falciparum malaria, Nature 434 (2005) 214-217.
[42] R.W. Snow, E.L. Korenromp, E. Gouws, Pediatric mortality in Africa: plasmodium
falciparum malaria as a cause or risk? Am. J. Trop. Med. Hyg. 71 (2004) 16-24.
[43] D.I. Baruch, Adhesive receptors on malaria-parasitized red cells, Baillieres Best Pract.
Res. Clin. Haematol. 12 (1999) 747-761.
[44] P.M. Andrysiak, W.E. Collins, G.H. Campbell, Stage-specific and species-specific
antigens of Plasmodium vivax and Plasmodium ovale defined by monoclonal antibodies,
Infect. Immun. 54 (1986) 609-612.
[45] L. Malaguarnera, S. Musumeci, The immune response to Plasmodium falciparum
malaria, Lancet Infect. Dis. 2 (2002) 472-478.
Manijeh Vafa Homann
47
Human genetic factors involved in immunity to Plasmodium falciparum infection
[46] L. Schofield, I. Mueller, Clinical immunity to malaria, Curr. Mol. Med. 6 (2006) 205221.
[47] S.S. Yazdani, P. Mukherjee, V.S. Chauhan, C.E. Chitnis, Immune responses to asexual
blood-stages of malaria parasites, Curr. Mol. Med. 6 (2006) 187-203.
[48] F. Yarovinsky, D. Zhang, J.F. Andersen, G.L. Bannenberg, C.N. Serhan, M.S. Hayden,
S. Hieny, F.S. Sutterwala, R.A. Flavell, S. Ghosh, A. Sher, TLR11 activation of dendritic
cells by a protozoan profilin-like protein, Science 308 (2005) 1626-1629.
[49] L. Serghides, T.G. Smith, S.N. Patel, K.C. Kain, CD36 and malaria: friends or foes?
Trends Parasitol. 19 (2003) 461-469.
[50] B.J. Angus, K. Chotivanich, R. Udomsangpetch, N.J. White, In vivo removal of malaria
parasites from red blood cells without their destruction in acute falciparum malaria, Blood 90
(1997) 2037-2040.
[51] K. Chotivanich, R. Udomsangpetch, R. McGready, S. Proux, P. Newton, S.
Pukrittayakamee, S. Looareesuwan, N.J. White, Central role of the spleen in malaria parasite
clearance, J. Infect. Dis. 185 (2002) 1538-1541.
[52] M.F. Good, D.L. Doolan, Immune effector mechanisms in malaria, Curr. Opin. Immunol.
11 (1999) 412-419.
[53] M.F. Good, Towards a blood-stage vaccine for malaria: are we following all the leads?
Nat. Rev. Immunol. 1 (2001) 117-125.
[54] B.C. Urban, D.J. Ferguson, A. Pain, N. Willcox, M. Plebanski, J.M. Austyn, D.J.
Roberts, Plasmodium falciparum-infected erythrocytes modulate the maturation of dendritic
cells, Nature 400 (1999) 73-77.
[55] B.C. Urban, D.J. Roberts, Inhibition of T cell function during malaria: implications for
immunology and vaccinology, J. Exp. Med. 197 (2003) 137-141.
[56] B.C. Urban, N. Willcox, D.J. Roberts, A role for CD36 in the regulation of dendritic cell
function, Proc. Natl. Acad. Sci. U. S. A. 98 (2001) 8750-8755.
48
[57] B.C. Urban, T. Mwangi, A. Ross, S. Kinyanjui, M. Mosobo, O. Kai, B. Lowe, K. Marsh,
D.J. Roberts, Peripheral blood dendritic cells in children with acute Plasmodium falciparum
malaria, Blood 98 (2001) 2859-2861.
[58] J.A. Perry, A. Rush, R.J. Wilson, C.S. Olver, A.C. Avery, Dendritic cells from malariainfected mice are fully functional APC, J. Immunol. 172 (2004) 475-482.
[59] M. Ho, P. Tongtawe, J. Kriangkum, T. Wimonwattrawatee, K. Pattanapanyasat, L.
Bryant, J. Shafiq, P. Suntharsamai, S. Looareesuwan, H.K. Webster, J.F. Elliott, Polyclonal
expansion of peripheral gamma delta T cells in human Plasmodium falciparum malaria,
Infect. Immun. 62 (1994) 855-862.
[60] C. Behr, R. Poupot, M.A. Peyrat, Y. Poquet, P. Constant, P. Dubois, M. Bonneville, J.J.
Fournie, Plasmodium falciparum stimuli for human gammadelta T cells are related to
phosphorylated antigens of mycobacteria, Infect. Immun. 64 (1996) 2892-2896.
[61] M.R. Goodier, C. Lundqvist, M.L. Hammarstrom, M. Troye-Blomberg, J. Langhorne,
Cytokine profiles for human V gamma 9+ T cells stimulated by Plasmodium falciparum,
Parasite Immunol. 17 (1995) 413-423.
[62] M. Hensmann, D. Kwiatkowski, Cellular basis of early cytokine response to Plasmodium
falciparum, Infect. Immun. 69 (2001) 2364-2371.
[63] M.M. Elloso, H.C. van der Heyde, J.A. vande Waa, D.D. Manning, W.P. Weidanz,
Inhibition of Plasmodium falciparum in vitro by human gamma delta T cells, J. Immunol. 153
(1994) 1187-1194.
[64] M. Troye-Blomberg, S. Worku, P. Tangteerawatana, R. Jamshaid, K. Soderstrom, G.
Elghazali, L. Moretta, M. Hammarstrom, L. Mincheva-Nilsson, Human gamma delta T cells
that inhibit the in vitro growth of the asexual blood stages of the Plasmodium falciparum
parasite express cytolytic and proinflammatory molecules, Scand. J. Immunol. 50 (1999) 642650.
[65] S.E. Farouk, L. Mincheva-Nilsson, A.M. Krensky, F. Dieli, M. Troye-Blomberg, Gamma
delta T cells inhibit in vitro growth of the asexual blood stages of Plasmodium falciparum by
a granule exocytosis-dependent cytotoxic pathway that requires granulysin, Eur. J. Immunol.
34 (2004) 2248-2256.
Manijeh Vafa Homann
49
Human genetic factors involved in immunity to Plasmodium falciparum infection
[66] K. Artavanis-Tsakonas, E.M. Riley, Innate immune response to malaria: rapid induction
of IFN-gamma from human NK cells by live Plasmodium falciparum-infected erythrocytes, J.
Immunol. 169 (2002) 2956-2963.
[67] A.J. Luty, B. Lell, R. Schmidt-Ott, L.G. Lehman, D. Luckner, B. Greve, P. Matousek, K.
Herbich, D. Schmid, F. Migot-Nabias, P. Deloron, R.S. Nussenzweig, P.G. Kremsner,
Interferon-gamma responses are associated with resistance to reinfection with Plasmodium
falciparum in young African children, J. Infect. Dis. 179 (1999) 980-988.
[68] D. Dodoo, F.M. Omer, J. Todd, B.D. Akanmori, K.A. Koram, E.M. Riley, Absolute
levels and ratios of proinflammatory and anti-inflammatory cytokine production in vitro
predict clinical immunity to Plasmodium falciparum malaria, J. Infect. Dis. 185 (2002) 971979.
[69] K. Artavanis-Tsakonas, K. Eleme, K.L. McQueen, N.W. Cheng, P. Parham, D.M. Davis,
E.M. Riley, Activation of a subset of human NK cells upon contact with Plasmodium
falciparum-infected erythrocytes, J. Immunol. 171 (2003) 5396-5405.
[70] M.C. D'Ombrain, D.S. Hansen, K.M. Simpson, L. Schofield, gammadelta-T cells
expressing NK receptors predominate over NK cells and conventional T cells in the innate
IFN-gamma response to Plasmodium falciparum malaria, Eur. J. Immunol. 37 (2007) 18641873.
[71] S. COHEN, I.A. McGREGOR, S. CARRINGTON, Gamma-globulin and acquired
immunity to human malaria, Nature 192 (1961) 733-737.
[72] O. Garraud, S. Mahanty, R. Perraut, Malaria-specific antibody subclasses in immune
individuals: a key source of information for vaccine design, Trends Immunol. 24 (2003) 3035.
[73] O. Garraud, R. Perraut, G. Riveau, T.B. Nutman, Class and subclass selection in parasitespecific antibody responses, Trends Parasitol. 19 (2003) 300-304.
[74] H. Perlmann, H. Helmby, M. Hagstedt, J. Carlson, P.H. Larsson, M. Troye-Blomberg, P.
Perlmann, IgE elevation and IgE anti-malarial antibodies in Plasmodium falciparum malaria:
association of high IgE levels with cerebral malaria, Clin. Exp. Immunol. 97 (1994) 284-292.
50
[75] S. Bereczky, S.M. Montgomery, M. Troye-Blomberg, I. Rooth, M.A. Shaw, A. Farnert,
Elevated anti-malarial IgE in asymptomatic individuals is associated with reduced risk for
subsequent clinical malaria, Int. J. Parasitol. 34 (2004) 935-942.
[76] E. Israelsson, H. Balogun, N.M. Vasconcelos, J. Beser, C. Roussilhon, C. Rogier, J.F.
Trape, K. Berzins, Antibody responses to a C-terminal fragment of the Plasmodium
falciparum blood-stage antigen Pf332 in Senegalese individuals naturally primed to the
parasite, Clin. Exp. Immunol. 152 (2008) 64-71.
[77] P.H. David, M. Hommel, L.H. Miller, I.J. Udeinya, L.D. Oligino, Parasite sequestration
in Plasmodium falciparum malaria: spleen and antibody modulation of cytoadherence of
infected erythrocytes, Proc. Natl. Acad. Sci. U. S. A. 80 (1983) 5075-5079.
[78] B. Wahlin, M. Wahlgren, H. Perlmann, K. Berzins, A. Bjorkman, M.E. Patarroyo, P.
Perlmann, Human antibodies to a Mr 155,000 Plasmodium falciparum antigen efficiently
inhibit merozoite invasion, Proc. Natl. Acad. Sci. U. S. A. 81 (1984) 7912-7916.
[79] L.M. Kumaratilake, A. Ferrante, T. Jaeger, C. Rzepczyk, GM-CSF-induced priming of
human neutrophils for enhanced phagocytosis and killing of asexual blood stages of
Plasmodium falciparum: synergistic effects of GM-CSF and TNF, Parasite Immunol. 18
(1996) 115-123.
[80] H. Bouharoun-Tayoun, P. Attanath, A. Sabchareon, T. Chongsuphajaisiddhi, P. Druilhe,
Antibodies that protect humans against Plasmodium falciparum blood stages do not on their
own inhibit parasite growth and invasion in vitro, but act in cooperation with monocytes, J.
Exp. Med. 172 (1990) 1633-1641.
[81] G. Giribaldi, D. Ulliers, F. Mannu, P. Arese, F. Turrini, Growth of Plasmodium
falciparum induces stage-dependent haemichrome formation, oxidative aggregation of band
3, membrane deposition of complement and antibodies, and phagocytosis of parasitized
erythrocytes, Br. J. Haematol. 113 (2001) 492-499.
[82] R. Stephens, J. Langhorne, Priming of CD4+ T cells and development of CD4+ T cell
memory; lessons for malaria, Parasite Immunol. 28 (2006) 25-30.
[83] M. Troye-Blomberg, K. Berzins, P. Perlmann, T-cell control of immunity to the asexual
blood stages of the malaria parasite, Crit. Rev. Immunol. 14 (1994) 131-155.
Manijeh Vafa Homann
51
Human genetic factors involved in immunity to Plasmodium falciparum infection
[84] E.M. Riley, S. Wahl, D.J. Perkins, L. Schofield, Regulating immunity to malaria,
Parasite Immunol. 28 (2006) 35-49.
[85] Y. Wu, Q.H. Wang, L. Zheng, H. Feng, J. Liu, S.H. Ma, Y.M. Cao, Plasmodium yoelii:
distinct CD4(+)CD25(+) regulatory T cell responses during the early stages of infection in
susceptible and resistant mice, Exp. Parasitol. 115 (2007) 301-304.
[86] M. Walther, J.E. Tongren, L. Andrews, D. Korbel, E. King, H. Fletcher, R.F. Andersen,
P. Bejon, F. Thompson, S.J. Dunachie, F. Edele, J.B. de Souza, R.E. Sinden, S.C. Gilbert,
E.M. Riley, A.V. Hill, Upregulation of TGF-beta, FOXP3, and CD4+CD25+ regulatory T
cells correlates with more rapid parasite growth in human malaria infection, Immunity 23
(2005) 287-296.
[87] I.A. Clark, L. Schofield, Pathogenesis of malaria, Parasitol. Today 16 (2000) 451-454.
[88] A.H. Fell, N.C. Smith, Immunity to asexual blood stages of Plasmodium: is resistance to
acute malaria adaptive or innate? Parasitol. Today 14 (1998) 364-369.
[89] N. Favre, B. Ryffel, G. Bordmann, W. Rudin, The course of Plasmodium chabaudi
chabaudi infections in interferon-gamma receptor deficient mice, Parasite Immunol. 19
(1997) 375-383.
[90] P. Deloron, C. Chougnet, J.P. Lepers, S. Tallet, P. Coulanges, Protective value of
elevated levels of gamma interferon in serum against exoerythrocytic stages of Plasmodium
falciparum, J. Clin. Microbiol. 29 (1991) 1757-1760.
[91] N.M. Anstey, J.B. Weinberg, M.Y. Hassanali, E.D. Mwaikambo, D. Manyenga, M.A.
Misukonis, D.R. Arnelle, D. Hollis, M.I. McDonald, D.L. Granger, Nitric oxide in Tanzanian
children with malaria: inverse relationship between malaria severity and nitric oxide
production/nitric oxide synthase type 2 expression, J. Exp. Med. 184 (1996) 557-567.
[92] K. Artavanis-Tsakonas, J.E. Tongren, E.M. Riley, The war between the malaria parasite
and the immune system: immunity, immunoregulation and immunopathology, Clin. Exp.
Immunol. 133 (2003) 145-152.
52
[93] G.E. Grau, T.E. Taylor, M.E. Molyneux, J.J. Wirima, P. Vassalli, M. Hommel, P.H.
Lambert, Tumor necrosis factor and disease severity in children with falciparum malaria, N.
Engl. J. Med. 320 (1989) 1586-1591.
[94] A. Linke, R. Kuhn, W. Muller, N. Honarvar, C. Li, J. Langhorne, Plasmodium chabaudi
chabaudi: differential susceptibility of gene-targeted mice deficient in IL-10 to an
erythrocytic-stage infection, Exp. Parasitol. 84 (1996) 253-263.
[95] N.P. Day, T.T. Hien, T. Schollaardt, P.P. Loc, L.V. Chuong, T.T. Chau, N.T. Mai, N.H.
Phu, D.X. Sinh, N.J. White, M. Ho, The prognostic and pathophysiologic role of pro- and
antiinflammatory cytokines in severe malaria, J. Infect. Dis. 180 (1999) 1288-1297.
[96] J.A. Kurtzhals, V. Adabayeri, B.Q. Goka, B.D. Akanmori, J.O. Oliver-Commey, F.K.
Nkrumah, C. Behr, L. Hviid, Low plasma concentrations of interleukin 10 in severe malarial
anaemia compared with cerebral and uncomplicated malaria, Lancet 351 (1998) 1768-1772.
[97] C. Othoro, A.A. Lal, B. Nahlen, D. Koech, A.S. Orago, V. Udhayakumar, A low
interleukin-10 tumor necrosis factor-alpha ratio is associated with malaria anemia in children
residing in a holoendemic malaria region in western Kenya, J. Infect. Dis. 179 (1999) 279282.
[98] A. Fortin, M.M. Stevenson, P. Gros, Susceptibility to malaria as a complex trait: big
pressure from a tiny creature, Hum. Mol. Genet. 11 (2002) 2469-2478.
[99] C.J. Uneke, Plasmodium falciparum malaria and ABO blood group: is there any
relationship? Parasitol. Res. 100 (2007) 759-765.
[100] C.M. Cserti, W.H. Dzik, The ABO blood group system and Plasmodium falciparum
malaria, Blood 110 (2007) 2250-2258.
[101] J. Carlson, M. Wahlgren, Plasmodium falciparum erythrocyte rosetting is mediated by
promiscuous lectin-like interactions, J. Exp. Med. 176 (1992) 1311-1317.
[102] A.V. Hill, Molecular epidemiology of the thalassaemias (including haemoglobin E),
Baillieres Clin. Haematol. 5 (1992) 209-238.
Manijeh Vafa Homann
53
Human genetic factors involved in immunity to Plasmodium falciparum infection
[103] S.A. Sarnaik, Thalassemia and related hemoglobinopathies, Indian J. Pediatr. 72 (2005)
319-324.
[104] T.N. Williams, Human red blood cell polymorphisms and malaria, Curr. Opin.
Microbiol. 9 (2006) 388-394.
[105] J. Flint, R.M. Harding, A.J. Boyce, J.B. Clegg, The population genetics of the
haemoglobinopathies, Baillieres Clin. Haematol. 11 (1998) 1-51.
[106] M. Aidoo, D.J. Terlouw, M.S. Kolczak, P.D. McElroy, F.O. ter Kuile, S. Kariuki, B.L.
Nahlen, A.A. Lal, V. Udhayakumar, Protective effects of the sickle cell gene against malaria
morbidity and mortality, Lancet 359 (2002) 1311-1312.
[107] T.N. Williams, T.W. Mwangi, S. Wambua, N.D. Alexander, M. Kortok, R.W. Snow, K.
Marsh, Sickle cell trait and the risk of Plasmodium falciparum malaria and other childhood
diseases, J. Infect. Dis. 192 (2005) 178-186.
[108] P.E. Duffy, M. Fried, Red blood cells that do and red blood cells that don't: how to
resist a persistent parasite, Trends Parasitol. 22 (2006) 99-101.
[109] K. Ayi, F. Turrini, A. Piga, P. Arese, Enhanced phagocytosis of ring-parasitized mutant
erythrocytes: a common mechanism that may explain protection against falciparum malaria in
sickle trait and beta-thalassemia trait, Blood 104 (2004) 3364-3371.
[110] M.J. Friedman, Erythrocytic mechanism of sickle cell resistance to malaria, Proc. Natl.
Acad. Sci. U. S. A. 75 (1978) 1994-1997.
[111] G. Pasvol, D.J. Weatherall, R.J. Wilson, Cellular mechanism for the protective effect of
haemoglobin S against P. falciparum malaria, Nature 274 (1978) 701-703.
[112] H.L. Shear, E.F. Roth Jr, M.E. Fabry, F.D. Costantini, A. Pachnis, A. Hood, R.L. Nagel,
Transgenic mice expressing human sickle hemoglobin are partially resistant to rodent malaria,
Blood 81 (1993) 222-226.
[113] Y.A. Abu-Zeid, N.H. Abdulhadi, T.G. Theander, L. Hviid, B.O. Saeed, S. Jepsen, J.B.
Jensen, R.A. Bayoumi, Seasonal changes in cell mediated immune responses to soluble
54
Plasmodium falciparum antigens in children with haemoglobin AA and haemoglobin AS,
Trans. R. Soc. Trop. Med. Hyg. 86 (1992) 20-22.
[114] R.A. Bayoumi, Y.A. Abu-Zeid, N.H. Abdulhadi, B.O. Saeed, T.G. Theander, L. Hviid,
H.W. Ghalib, A.H. Nugud, S. Jepsen, J.B. Jensen, Cell-mediated immune responses to
Plasmodium falciparum purified soluble antigens in sickle-cell trait subjects, Immunol. Lett.
25 (1990) 243-249.
[115] R. Cholera, N.J. Brittain, M.R. Gillrie, T.M. Lopera-Mesa, S.A. Diakite, T. Arie, M.A.
Krause, A. Guindo, A. Tubman, H. Fujioka, D.A. Diallo, O.K. Doumbo, M. Ho, T.E.
Wellems, R.M. Fairhurst, Impaired cytoadherence of Plasmodium falciparum-infected
erythrocytes containing sickle hemoglobin, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 991996.
[116] D.R. Higgs, M.A. Vickers, A.O. Wilkie, I.M. Pretorius, A.P. Jarman, D.J. Weatherall,
A review of the molecular genetics of the human alpha-globin gene cluster, Blood 73 (1989)
1081-1104.
[117] F.P. Mockenhaupt, S. Ehrhardt, S. Gellert, R.N. Otchwemah, E. Dietz, S.D. Anemana,
U. Bienzle, Alpha(+)-thalassemia protects African children from severe malaria, Blood 104
(2004) 2003-2006.
[118] S. Wambua, T.W. Mwangi, M. Kortok, S.M. Uyoga, A.W. Macharia, J.K. Mwacharo,
D.J. Weatherall, R.W. Snow, K. Marsh, T.N. Williams, The effect of alpha+-thalassaemia on
the incidence of malaria and other diseases in children living on the coast of Kenya, PLoS
Med. 3 (2006) e158.
[119] T.N. Williams, K. Maitland, S. Bennett, M. Ganczakowski, T.E. Peto, C.I. Newbold,
D.K. Bowden, D.J. Weatherall, J.B. Clegg, High incidence of malaria in alpha-thalassaemic
children, Nature 383 (1996) 522-525.
[120] J. Carlson, G.B. Nash, V. Gabutti, F. al-Yaman, M. Wahlgren, Natural protection
against severe Plasmodium falciparum malaria due to impaired rosette formation, Blood 84
(1994) 3909-3914.
[121] I.A. Cockburn, M.J. Mackinnon, A. O'Donnell, S.J. Allen, J.M. Moulds, M. Baisor, M.
Bockarie, J.C. Reeder, J.A. Rowe, A human complement receptor 1 polymorphism that
Manijeh Vafa Homann
55
Human genetic factors involved in immunity to Plasmodium falciparum infection
reduces Plasmodium falciparum rosetting confers protection against severe malaria, Proc.
Natl. Acad. Sci. U. S. A. 101 (2004) 272-277.
[122] V. Tripathy, B.M. Reddy, Present status of understanding on the G6PD deficiency and
natural selection, J. Postgrad. Med. 53 (2007) 193-202.
[123] A.C. Allison, Glucose-6-phosphate dehydrogenase deficiency in red blood cells of East
Africans, Nature 186 (1960) 531-532.
[124] A.G. Motulsky, Metabolic polymorphisms and the role of infectious diseases in human
evolution, Hum. Biol. 32 (1960) 28-62.
[125] C. Ruwende, S.C. Khoo, R.W. Snow, S.N. Yates, D. Kwiatkowski, S. Gupta, P. Warn,
C.E. Allsopp, S.C. Gilbert, N. Peschu, Natural selection of hemi- and heterozygotes for G6PD
deficiency in Africa by resistance to severe malaria, Nature 376 (1995) 246-249.
[126] E. Beutler, W. Kuhl, J.L. Vives-Corrons, J.T. Prchal, Molecular heterogeneity of
glucose-6-phosphate dehydrogenase A-, Blood 74 (1989) 2550-2555.
[127] J. May, C.G. Meyer, L. Grossterlinden, O.G. Ademowo, F.P. Mockenhaupt, P.E.
Olumese, A.G. Falusi, L. Luzzatto, U. Bienzle, Red cell glucose-6-phosphate dehydrogenase
status and pyruvate kinase activity in a Nigerian population, Trop. Med. Int. Health 5 (2000)
119-123.
[128] E.F. Roth Jr, C. Raventos-Suarez, A. Rinaldi, R.L. Nagel, Glucose-6-phosphate
dehydrogenase deficiency inhibits in vitro growth of Plasmodium falciparum, Proc. Natl.
Acad. Sci. U. S. A. 80 (1983) 298-299.
[129] M. Cappadoro, G. Giribaldi, E. O'Brien, F. Turrini, F. Mannu, D. Ulliers, G. Simula, L.
Luzzatto, P. Arese, Early phagocytosis of glucose-6-phosphate dehydrogenase (G6PD)deficient erythrocytes parasitized by Plasmodium falciparum may explain malaria protection
in G6PD deficiency, Blood 92 (1998) 2527-2534.
[130] A.V. Hill, C.E. Allsopp, D. Kwiatkowski, N.M. Anstey, P. Twumasi, P.A. Rowe, S.
Bennett, D. Brewster, A.J. McMichael, B.M. Greenwood, Common west African HLA
antigens are associated with protection from severe malaria, Nature 352 (1991) 595-600.
56
[131] J. May, B. Lell, A.J. Luty, C.G. Meyer, P.G. Kremsner, HLA-DQB1*0501-restricted
Th1 type immune responses to Plasmodium falciparum liver stage antigen 1 protect against
malaria anemia and reinfections, J. Infect. Dis. 183 (2001) 168-172.
[132] J. May, C.G. Meyer, J.F. Kun, B. Lell, D. Luckner, A.K. Dippmann, U. Bienzle, P.G.
Kremsner, HLA class II factors associated with Plasmodium falciparum merozoite surface
antigen allele families, J. Infect. Dis. 179 (1999) 1042-1045.
[133] A.D. Osafo-Addo, K.A. Koram, A.R. Oduro, M. Wilson, A. Hodgson, W.O. Rogers,
HLA-DRB1*04 allele is associated with severe malaria in northern Ghana, Am. J. Trop. Med.
Hyg. 78 (2008) 251-255.
[134] M.J. Mackinnon, T.W. Mwangi, R.W. Snow, K. Marsh, T.N. Williams, Heritability of
malaria in Africa, PLoS Med. 2 (2005) e340.
[135] A. Dolo, D. Modiano, B. Maiga, M. Daou, G. Dolo, H. Guindo, M. Ba, H. Maiga, D.
Coulibaly, H. Perlman, M.T. Blomberg, Y.T. Toure, M. Coluzzi, O. Doumbo, Difference in
susceptibility to malaria between two sympatric ethnic groups in Mali, Am. J. Trop. Med.
Hyg. 72 (2005) 243-248.
[136] D. Modiano, V. Petrarca, B.S. Sirima, A. Bosman, I. Nebie, D. Diallo, L. Lamizana, F.
Esposito, M. Coluzzi, Plasmodium falciparum malaria in sympatric ethnic groups of Burkina
Faso, west Africa, Parassitologia 37 (1995) 255-259.
[137] J.P. Pandey, A. Nasr, K.M. Rocca, M. Troy-Blomberg, G. Elghazali, Significant
differences in GM allotype frequencies between two sympatric tribes with markedly
differential susceptibility to malaria, Parasite Immunol. 29 (2007) 267-269.
[138] D. Modiano, A. Chiucchiuini, V. Petrarca, B.S. Sirima, G. Luoni, H. Perlmann, F.
Esposito, M. Coluzzi, Humoral response to Plasmodium falciparum Pf155/ring-infected
erythrocyte surface antigen and Pf332 in three sympatric ethnic groups of Burkina Faso, Am.
J. Trop. Med. Hyg. 58 (1998) 220-224.
[139] D. Modiano, A. Chiucchiuini, V. Petrarca, B.S. Sirima, G. Luoni, M.A. Roggero, G.
Corradin, M. Coluzzi, F. Esposito, Interethnic differences in the humoral response to nonrepetitive regions of the Plasmodium falciparum circumsporozoite protein, Am. J. Trop. Med.
Hyg. 61 (1999) 663-667.
Manijeh Vafa Homann
57
Human genetic factors involved in immunity to Plasmodium falciparum infection
[140] D. Modiano, V. Petrarca, B.S. Sirima, I. Nebie, D. Diallo, F. Esposito, M. Coluzzi,
Different response to Plasmodium falciparum malaria in west African sympatric ethnic
groups, Proc. Natl. Acad. Sci. U. S. A. 93 (1996) 13206-13211.
[141] S.E. Farouk, A. Dolo, S. Bereczky, B. Kouriba, B. Maiga, A. Farnert, H. Perlmann, M.
Hayano, S.M. Montgomery, O.K. Doumbo, M. Troye-Blomberg, Different antibody- and
cytokine-mediated responses to Plasmodium falciparum parasite in two sympatric ethnic
tribes living in Mali, Microbes Infect. 7 (2005) 110-117.
[142] A. Bolad, S.E. Farouk, E. Israelsson, A. Dolo, O.K. Doumbo, I. Nebie, B. Maiga, B.
Kouriba, G. Luoni, B.S. Sirima, D. Modiano, K. Berzins, M. Troye-Blomberg, Distinct
interethnic differences in immunoglobulin G class/subclass and immunoglobulin M antibody
responses to malaria antigens but not in immunoglobulin G responses to nonmalarial antigens
in sympatric tribes living in West Africa, Scand. J. Immunol. 61 (2005) 380-386.
[143] D. Modiano, G. Luoni, B.S. Sirima, A. Lanfrancotti, V. Petrarca, F. Cruciani, J.
Simpore, B.M. Ciminelli, E. Foglietta, P. Grisanti, I. Bianco, G. Modiano, M. Coluzzi, The
lower susceptibility to Plasmodium falciparum malaria of Fulani of Burkina Faso (west
Africa) is associated with low frequencies of classic malaria-resistance genes, Trans. R. Soc.
Trop. Med. Hyg. 95 (2001) 149-152.
[144] A. Nasr, G. Elghazali, H. Giha, M. Troye-Blomberg, K. Berzins, Interethnic differences
in carriage of haemoglobin AS and Fcgamma receptor IIa (CD32) genotypes in children
living in eastern Sudan, Acta Trop. 105 (2008) 191-195.
[145] A.D. Bryceson, A.F. Fleming, G.M. Edington, Splenomegaly in Northern Nigeria, Acta
Trop. 33 (1976) 185-214.
[146] B.M. Greenwood, F. Groenendaal, A.K. Bradley, A.M. Greenwood, F. Shenton, S.
Tulloch, R. Hayes, Ethnic differences in the prevalence of splenomegaly and malaria in The
Gambia, Ann. Trop. Med. Parasitol. 81 (1987) 345-354.
[147] M. Vafa, B. Maiga, K. Berzins, M. Hayano, S. Bereczky, A. Dolo, M. Daou, C. Arama,
B. Kouriba, A. Farnert, O.K. Doumbo, M. Troye-Blomberg, Associations between the IL-4 590 T allele and Plasmodium falciparum infection prevalence in asymptomatic Fulani of
Mali, Microbes Infect. 9 (2007) 1043-1048.
58
[148] M.G. Torcia, V. Santarlasci, L. Cosmi, A. Clemente, L. Maggi, V.D. Mangano, F.
Verra, G. Bancone, I. Nebie, B.S. Sirima, F. Liotta, F. Frosali, R. Angeli, C. Severini, A.R.
Sannella, P. Bonini, M. Lucibello, E. Maggi, E. Garaci, M. Coluzzi, F. Cozzolino, F.
Annunziato, S. Romagnani, D. Modiano, Functional deficit of T regulatory cells in Fulani, an
ethnic group with low susceptibility to Plasmodium falciparum malaria, Proc. Natl. Acad. Sci.
U. S. A. 105 (2008) 646-651.
[149] K. Omi, J. Ohashi, J. Patarapotikul, H. Hananantachai, I. Naka, S. Looareesuwan, K.
Tokunaga, Fcgamma receptor IIA and IIIB polymorphisms are associated with susceptibility
to cerebral malaria, Parasitol. Int. 51 (2002) 361-366.
[150] Y.P. Shi, B.L. Nahlen, S. Kariuki, K.B. Urdahl, P.D. McElroy, J.M. Roberts, A.A. Lal,
Fcgamma receptor IIa (CD32) polymorphism is associated with protection of infants against
high-density Plasmodium falciparum infection. VII. Asembo Bay Cohort Project, J. Infect.
Dis. 184 (2001) 107-111.
[151] A. Nasr, N.C. Iriemenam, M. Troye-Blomberg, H.A. Giha, H.A. Balogun, O.F. Osman,
S.M. Montgomery, G. ElGhazali, K. Berzins, Fc gamma receptor IIa (CD32) polymorphism
and antibody responses to asexual blood-stage antigens of Plasmodium falciparum malaria in
Sudanese patients, Scand. J. Immunol. 66 (2007) 87-96.
[152] M.R. Clatworthy, L. Willcocks, B. Urban, J. Langhorne, T.N. Williams, N. Peshu, N.A.
Watkins, R.A. Floto, K.G. Smith, Systemic lupus erythematosus-associated defects in the
inhibitory receptor FcgammaRIIb reduce susceptibility to malaria, Proc. Natl. Acad. Sci. U. S.
A. 104 (2007) 7169-7174.
[153] P.A. Warmerdam, J.G. van de Winkel, A. Vlug, N.A. Westerdaal, P.J. Capel, A single
amino acid in the second Ig-like domain of the human Fc gamma receptor II is critical for
human IgG2 binding, J. Immunol. 147 (1991) 1338-1343.
[154] E.M. Braga, K.K. Scopel, N.T. Komatsu, M. da Silva-Nunes, M.U. Ferreira,
Polymorphism of the Fcgamma receptor IIA and malaria morbidity, J. Mol. Gen. Med 1
(2005) 5-10.
Manijeh Vafa Homann
59
Human genetic factors involved in immunity to Plasmodium falciparum infection
[155] L.J. Rosenwasser, D.J. Klemm, J.K. Dresback, H. Inamura, J.J. Mascali, M. Klinnert, L.
Borish, Promoter polymorphisms in the chromosome 5 gene cluster in asthma and atopy,
Clin. Exp. Allergy 25 Suppl 2 (1995) 74-8; discussion 95-6.
[156] D.M. Turner, D.M. Williams, D. Sankaran, M. Lazarus, P.J. Sinnott, I.V. Hutchinson,
An investigation of polymorphism in the interleukin-10 gene promoter, Eur. J. Immunogenet.
24 (1997) 1-8.
[157] W. McGuire, A.V. Hill, C.E. Allsopp, B.M. Greenwood, D. Kwiatkowski, Variation in
the TNF-alpha promoter region associated with susceptibility to cerebral malaria, Nature 371
(1994) 508-510.
[158] W. McGuire, J.C. Knight, A.V. Hill, C.E. Allsopp, B.M. Greenwood, D. Kwiatkowski,
Severe malarial anemia and cerebral malaria are associated with different tumor necrosis
factor promoter alleles, J. Infect. Dis. 179 (1999) 287-290.
[159] J.C. Knight, I. Udalova, A.V. Hill, B.M. Greenwood, N. Peshu, K. Marsh, D.
Kwiatkowski, A polymorphism that affects OCT-1 binding to the TNF promoter region is
associated with severe malaria, Nat. Genet. 22 (1999) 145-150.
[160] J. Ohashi, I. Naka, J. Patarapotikul, H. Hananantachai, S. Looareesuwan, K. Tokunaga,
A single-nucleotide substitution from C to T at position -1055 in the IL-13 promoter is
associated with protection from severe malaria in Thailand, Genes Immun. 4 (2003) 528-531.
[161] F. Verra, G. Luoni, C. Calissano, M. Troye-Blomberg, P. Perlmann, H. Perlmann, B.
Arca, B.S. Sirima, A. Konate, M. Coluzzi, D. Kwiatkowski, D. Modiano, IL4-589C/T
polymorphism and IgE levels in severe malaria, Acta Trop. 90 (2004) 205-209.
[162] B.A. Gyan, B. Goka, J.T. Cvetkovic, J.L. Kurtzhals, V. Adabayeri, H. Perlmann, A.K.
Lefvert, B.D. Akanmori, M. Troye-Blomberg, Allelic polymorphisms in the repeat and
promoter regions of the interleukin-4 gene and malaria severity in Ghanaian children, Clin.
Exp. Immunol. 138 (2004) 145-150.
[163] G. Luoni, F. Verra, B. Arca, B.S. Sirima, M. Troye-Blomberg, M. Coluzzi, D.
Kwiatkowski, D. Modiano, Antimalarial antibody levels and IL4 polymorphism in the Fulani
of West Africa, Genes Immun. 2 (2001) 411-414.
60
[164] V.D. Mangano, G. Luoni, K.A. Rockett, B.S. Sirima, A. Konate, J. Forton, T.G. Clark,
G. Bancone, E.S. Akha, D.P. Kwiatkowski, D. Modiano, Interferon regulatory factor-1
polymorphisms are associated with the control of Plasmodium falciparum infection, Genes
Immun. (2008) .
[165] R. Dzikowski, T.J. Templeton, K. Deitsch, Variant antigen gene expression in malaria,
Cell. Microbiol. 8 (2006) 1371-1381.
[166] T. Fandeur, W. Chalvet, Variant- and strain-specific immunity in Saimiri infected with
Plasmodium falciparum, Am. J. Trop. Med. Hyg. 58 (1998) 225-231.
[167] H.L. Rotman, T.M. Daly, C.A. Long, Plasmodium: immunization with carboxylterminal regions of MSP-1 protects against homologous but not heterologous blood-stage
parasite challenge, Exp. Parasitol. 91 (1999) 78-85.
[168] G.M. Jeffery, Epidemiological significance of repeated infections with homologous and
heterologous strains and species of Plasmodium, Bull. World Health Organ. 35 (1966) 873882.
[169] J.K. Baird, Age-dependent characteristics of protection v. susceptibility to Plasmodium
falciparum, Ann. Trop. Med. Parasitol. 92 (1998) 367-390.
[170] D. Arnot, Unstable malaria in Sudan: the influence of the dry season. Clone multiplicity
of Plasmodium falciparum infections in individuals exposed to variable levels of disease
transmission, Trans. R. Soc. Trop. Med. Hyg. 92 (1998) 580-585.
[171] H.A. Babiker, L.C. Ranford-Cartwright, D. Walliker, Genetic structure and dynamics of
Plasmodium falciparum infections in the Kilombero region of Tanzania, Trans. R. Soc. Trop.
Med. Hyg. 93 Suppl 1 (1999) 11-14.
[172] T. Smith, I. Felger, M. Tanner, H.P. Beck, Premunition in Plasmodium falciparum
infection: insights from the epidemiology of multiple infections, Trans. R. Soc. Trop. Med.
Hyg. 93 Suppl 1 (1999) 59-64.
[173] F. al-Yaman, B. Genton, J.C. Reeder, R.F. Anders, T. Smith, M.P. Alpers, Reduced risk
of clinical malaria in children infected with multiple clones of Plasmodium falciparum in a
Manijeh Vafa Homann
61
Human genetic factors involved in immunity to Plasmodium falciparum infection
highly endemic area: a prospective community study, Trans. R. Soc. Trop. Med. Hyg. 91
(1997) 602-605.
[174] A. Farnert, I. Rooth, Svensson, G. Snounou, A. Bjorkman, Complexity of Plasmodium
falciparum infections is consistent over time and protects against clinical disease in Tanzanian
children, J. Infect. Dis. 179 (1999) 989-995.
[175] D.A. Muller, J.D. Charlwood, I. Felger, C. Ferreira, V. do Rosario, T. Smith,
Prospective risk of morbidity in relation to multiplicity of infection with Plasmodium
falciparum in Sao Tome, Acta Trop. 78 (2001) 155-162.
[176] O.H. Branch, S. Takala, S. Kariuki, B.L. Nahlen, M. Kolczak, W. Hawley, A.A. Lal,
Plasmodium falciparum genotypes, low complexity of infection, and resistance to subsequent
malaria in participants in the Asembo Bay Cohort Project, Infect. Immun. 69 (2001) 77837792.
[177] A. Ofosu-Okyere, M.J. Mackinnon, M.P. Sowa, K.A. Koram, F. Nkrumah, Y.D. Osei,
W.G. Hill, M.D. Wilson, D.E. Arnot, Novel Plasmodium falciparum clones and rising clone
multiplicities are associated with the increase in malaria morbidity in Ghanaian children
during the transition into the high transmission season, Parasitology 123 (2001) 113-123.
[178] M.M. Shigidi, R.A. Hashim, M.N. Idris, M.M. Mukhtar, T.E. Sokrab, Parasite diversity
in adult patients with cerebral malaria: a hospital-based, case-control study, Am. J. Trop.
Med. Hyg. 71 (2004) 754-757.
[179] O.K. Amodu, S.I. Oyedeji, F. Ntoumi, A.E. Orimadegun, R.A. Gbadegesin, P.E.
Olumese, O.O. Omotade, Complexity of the msp2 locus and the severity of childhood
malaria, in south-western Nigeria, Ann. Trop. Med. Parasitol. 102 (2008) 95-102.
[180] H.P. Beck, I. Felger, W. Huber, S. Steiger, T. Smith, N. Weiss, P. Alonso, M. Tanner,
Analysis of multiple Plasmodium falciparum infections in Tanzanian children during the
phase III trial of the malaria vaccine SPf66, J. Infect. Dis. 175 (1997) 921-926.
[181] I. Felger, L. Tavul, S. Kabintik, V. Marshall, B. Genton, M. Alpers, H.P. Beck,
Plasmodium falciparum: extensive polymorphism in merozoite surface antigen 2 alleles in an
area with endemic malaria in Papua New Guinea, Exp. Parasitol. 79 (1994) 106-116.
62
[182] F. Ntoumi, H. Contamin, C. Rogier, S. Bonnefoy, J.F. Trape, O. Mercereau-Puijalon,
Age-dependent carriage of multiple Plasmodium falciparum merozoite surface antigen-2
alleles in asymptomatic malaria infections, Am. J. Trop. Med. Hyg. 52 (1995) 81-88.
[183] J.K. Baird, Host age as a determinant of naturally acquired immunity to Plasmodium
falciparum, Parasitol. Today 11 (1995) 105-111.
[184] S. Viriyakosol, N. Siripoon, C. Petcharapirat, P. Petcharapirat, W. Jarra, S. Thaithong,
K.N. Brown, G. Snounou, Genotyping of Plasmodium falciparum isolates by the polymerase
chain reaction and potential uses in epidemiological studies, Bull. World Health Organ. 73
(1995) 85-95.
[185] J.A. Smythe, M.G. Peterson, R.L. Coppel, A.J. Saul, D.J. Kemp, R.F. Anders,
Structural diversity in the 45-kilodalton merozoite surface antigen of Plasmodium falciparum,
Mol. Biochem. Parasitol. 39 (1990) 227-234.
[186] A.W. Thomas, D.A. Carr, J.M. Carter, J.A. Lyon, Sequence comparison of allelic forms
of the Plasmodium falciparum merozoite surface antigen MSA2, Mol. Biochem. Parasitol. 43
(1990) 211-220.
[187] J.A. Smythe, R.L. Coppel, K.P. Day, R.K. Martin, A.M. Oduola, D.J. Kemp, R.F.
Anders, Structural diversity in the Plasmodium falciparum merozoite surface antigen 2, Proc.
Natl. Acad. Sci. U. S. A. 88 (1991) 1751-1755.
[188] V.K. Goel, X. Li, H. Chen, S.C. Liu, A.H. Chishti, S.S. Oh, Band 3 is a host receptor
binding merozoite surface protein 1 during the Plasmodium falciparum invasion of
erythrocytes, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 5164-5169.
[189] D. Eisen, H. Billman-Jacobe, V.F. Marshall, D. Fryauff, R.L. Coppel, Temporal
variation of the merozoite surface protein-2 gene of Plasmodium falciparum, Infect. Immun.
66 (1998) 239-246.
[190] R.N. Mshana, J. Boulandi, N.M. Mshana, J. Mayombo, G. Mendome, Cytokines in the
pathogenesis of malaria: levels of IL-I beta, IL-4, IL-6, TNF-alpha and IFN-gamma in plasma
of healthy individuals and malaria patients in a holoendemic area, J. Clin. Lab. Immunol. 34
(1991) 131-139.
Manijeh Vafa Homann
63
Human genetic factors involved in immunity to Plasmodium falciparum infection
[191] D. Vercelli, Regulation of IgE synthesis, Allergy Proc. 14 (1993) 413-416.
[192] T.R. Mosmann, Properties and functions of interleukin-10, Adv. Immunol. 56 (1994) 126.
[193] H. Watier, C. Auriault, A. Capron, Does Epstein-Barr virus infection confer selective
advantage to malaria-infected children? Lancet 341 (1993) 612-613.
[194] D. Modiano, G. Luoni, V. Petrarca, B. Sodiomon Sirima, M. De Luca, J. Simpore, M.
Coluzzi, J.G. Bodmer, G. Modiano, HLA class I in three West African ethnic groups: genetic
distances from sub-Saharan and Caucasoid populations, Tissue Antigens 57 (2001) 128-137.
[195] S. Bereczky, A. Dolo, B. Maiga, M. Hayano, F. Granath, S.M. Montgomery, M. Daou,
C. Arama, M. Troye-Blomberg, O.K. Doumbo, A. Farnert, Spleen enlargement and genetic
diversity of Plasmodium falciparum infection in two ethnic groups with different malaria
susceptibility in Mali, West Africa, Trans. R. Soc. Trop. Med. Hyg. 100 (2006) 248-257.
[196] M. Troye-Blomberg, P. Perlmann, T cell functions in Plasmodium falciparum and other
malarias, Prog. Allergy 41 (1988) 253-287.
[197] P. Perlmann, H. Perlmann, S. Looareesuwan, S. Krudsood, S. Kano, Y. Matsumoto, G.
Brittenham, M. Troye-Blomberg, M. Aikawa, Contrasting functions of IgG and IgE
antimalarial antibodies in uncomplicated and severe Plasmodium falciparum malaria, Am. J.
Trop. Med. Hyg. 62 (2000) 373-377.
[198] X. Ma, J.M. Chow, G. Gri, G. Carra, F. Gerosa, S.F. Wolf, R. Dzialo, G. Trinchieri,
The interleukin 12 p40 gene promoter is primed by interferon gamma in monocytic cells, J.
Exp. Med. 183 (1996) 147-157.
[199] S. Maruyama, K. Sumita, H. Shen, M. Kanoh, X. Xu, M. Sato, M. Matsumoto, H.
Shinomiya, Y. Asano, Identification of IFN regulatory factor-1 binding site in IL-12 p40 gene
promoter, J. Immunol. 170 (2003) 997-1001.
[200] J. Liu, S. Cao, L.M. Herman, X. Ma, Differential regulation of interleukin (IL)-12 p35
and p40 gene expression and interferon (IFN)-gamma-primed IL-12 production by IFN
regulatory factor 1, J. Exp. Med. 198 (2003) 1265-1276.
64
[201] S. Taki, T. Sato, K. Ogasawara, T. Fukuda, M. Sato, S. Hida, G. Suzuki, M. Mitsuyama,
E.H. Shin, S. Kojima, T. Taniguchi, Y. Asano, Multistage regulation of Th1-type immune
responses by the transcription factor IRF-1, Immunity 6 (1997) 673-679.
[202] D.L. McElligott, J.A. Phillips, C.A. Stillman, R.J. Koch, D.E. Mosier, M.V. Hobbs,
CD4+ T cells from IRF-1-deficient mice exhibit altered patterns of cytokine expression and
cell subset homeostasis, J. Immunol. 159 (1997) 4180-4186.
[203] M. Vidyarani, P. Selvaraj, S. Prabhu Anand, M.S. Jawahar, A.R. Adhilakshmi, P.R.
Narayanan, Interferon gamma (IFNgamma) & interleukin-4 (IL-4) gene variants & cytokine
levels in pulmonary tuberculosis, Indian J. Med. Res. 124 (2006) 403-410.
[204] S.P. Wiertsema, G. Baynam, S.K. Khoo, R.H. Veenhoven, N. van Heerbeek, G. Zhang,
I.A. Laing, G.T. Rijkers, J. Goldblatt, E.A. Sanders, P.N. Le Souef, Impact of genetic variants
in IL-4, IL-4 RA and IL-13 on the anti-pneumococcal antibody response, Vaccine 25 (2007)
306-313.
[205] H. Contamin, T. Fandeur, S. Bonnefoy, F. Skouri, F. Ntoumi, O. Mercereau-Puijalon,
PCR typing of field isolates of Plasmodium falciparum, J. Clin. Microbiol. 33 (1995) 944951.
[206] A. Farnert, A.P. Arez, H.A. Babiker, H.P. Beck, A. Benito, A. Bjorkman, M.C. Bruce,
D.J. Conway, K.P. Day, L. Henning, O. Mercereau-Puijalon, L.C. Ranford-Cartwright, J.M.
Rubio, G. Snounou, D. Walliker, J. Zwetyenga, V.E. do Rosario, Genotyping of Plasmodium
falciparum infections by PCR: a comparative multicentre study, Trans. R. Soc. Trop. Med.
Hyg. 95 (2001) 225-232.
[207] G.A. Luzzi, A.H. Merry, C.I. Newbold, K. Marsh, G. Pasvol, D.J. Weatherall, Surface
antigen expression on Plasmodium falciparum-infected erythrocytes is modified in alpha- and
beta-thalassemia, J. Exp. Med. 173 (1991) 785-791.
[208] T.N. Williams, D.J. Weatherall, C.I. Newbold, The membrane characteristics of
Plasmodium
falciparum-infected
and
-uninfected
heterozygous
alpha(0)thalassaemic
erythrocytes, Br. J. Haematol. 118 (2002) 663-670.
[209] F.P. Mockenhaupt, S. Ehrhardt, R. Otchwemah, T.A. Eggelte, S.D. Anemana, K. Stark,
U. Bienzle, E. Kohne, Limited influence of haemoglobin variants on Plasmodium falciparum
Manijeh Vafa Homann
65
Human genetic factors involved in immunity to Plasmodium falciparum infection
msp1 and msp2 alleles in symptomatic malaria, Trans. R. Soc. Trop. Med. Hyg. 98 (2004)
302-310.
[210] L. Konate, J. Zwetyenga, C. Rogier, E. Bischoff, D. Fontenille, A. Tall, A. Spiegel, J.F.
Trape, O. Mercereau-Puijalon, Variation of Plasmodium falciparum msp1 block 2 and msp2
allele prevalence and of infection complexity in two neighbouring Senegalese villages with
different transmission conditions, Trans. R. Soc. Trop. Med. Hyg. 93 Suppl 1 (1999) 21-28.
[211] F. Ntoumi, O. Mercereau-Puijalon, S. Ossari, A. Luty, J. Reltien, A. Georges, P. Millet,
Plasmodium falciparum: sickle-cell trait is associated with higher prevalence of multiple
infections in Gabonese children with asymptomatic infections, Exp. Parasitol. 87 (1997) 3946.
[212] J. Zwetyenga, C. Rogier, A. Tall, D. Fontenille, G. Snounou, J.F. Trape, O. MercereauPuijalon, No influence of age on infection complexity and allelic distribution in Plasmodium
falciparum infections in Ndiop, a Senegalese village with seasonal, mesoendemic malaria,
Am. J. Trop. Med. Hyg. 59 (1998) 726-735.
[213] I. Felger, T. Smith, D. Edoh, A. Kitua, P. Alonso, M. Tanner, H.P. Beck, Multiple
Plasmodium falciparum infections in Tanzanian infants, Trans. R. Soc. Trop. Med. Hyg. 93
Suppl 1 (1999) 29-34.
[214] T. Smith, I. Felger, A. Kitua, M. Tanner, H.P. Beck, Dynamics of multiple Plasmodium
falciparum infections in infants in a highly endemic area of Tanzania, Trans. R. Soc. Trop.
Med. Hyg. 93 Suppl 1 (1999) 35-39.
66