Sie sind auf Seite 1von 98

Universit of Kha U ty artoum

Faculty of Medic y cine


Departmen of Bioche nt emistry

The Imm muno-Pathogenesis of Severe Plasmodi e ium falcipa arum Malaria in S M Sudanese C Children

By

Atif Has ssan Khire elsied


BSc (A Alexandria U University)., MSc (U of K). ,

A Thesis Submitted in Fulfillment of th Requirem he ments for th Degree of he f PhD in Medical Biochemistry July 2008. upervisor. Su Pr rofessor. Mustafa Idris Elbashir s MD, PhD Biochemi D. istry.

Title:

The Immuno-pathogenesis of Plasmodium falciparum Malaria in Sudanese Children:

Location:

Department of Biochemistry. Faculty of Medicine, University of Khartoum P. O. Box 102, Khartoum, Sudan.

Supervisor:

Mustafa Idris Elbashir MD, PhD, Professor of Biochemistry Faculty of Medicine, University of Khartoum.

Table of Content

Tableofcontents..............................................................................................................................3 Dedication........................................................................................................................................6 Acknowledgement............................................................................................................................7 Listofabbreviations.........................................................................................................................8 Abstract.......................................................................................................................................... 0 1 Listoffigures.................................................................................................................................. 1 1 Listoftables.................................................................................................................................... 1 1

ChapterOne:IntroductiontoMalariaImmunityandPathogenesis................................... 2 1 Theglobalburdenanddistributionofmalaria.............................................................................. 3 1 Thebiologyofmalaria.................................................................................................................... 5 1 ThelifecycleofPlasmodiumfalciparum........................................................................................ 5 1 Themalariadisease........................................................................................................................ 8 1 Uncomplicatedmalaria.................................................................................................................. 8 1 Severemalariacomplications......................................................................................................... 8 1 Cerebralmalaria............................................................................................................................. 9 1 Severemalarialanemia.................................................................................................................. 9 1 ImmunitytoMalaria...................................................................................................................... 0 2 Theinnateimmuneresponsestomalaria..................................................................................... 0 2 2 Adaptiveimmuneresponsestomalaria........................................................................................ 2 Antisporozoiteimmunity ............................................................................................................. 2 . 2 Immunitytoliverstagesofmalariaparasite................................................................................. 3 2 Immunitytobloodstagesofmalariaparasite.............................................................................. 3 2 AntibodymediatedimmunitytobloodstagesofP.falciparum.................................................. 4 2 CellmediatedimmunitytobloodstagesofP.falciparum............................................................ 5 2 TheImmunoPathogenesisofSevereMalaria............................................................................. 6 2 Thepathogenesisofcerebralmalaria........................................................................................... 6 2

Thepathogenesisofseveremalariaanemia................................................................................. 8 2 Theroleofcytokinesinimmunopathogenesisofmalaria.......................................................... 9 2 Interferon(IFN)........................................................................................................................ 0 3 Tumornecrosisfactor(TNF).................................................................................................... 0 3 Lymphotoxin(LT)........................................................................................................................ 1 3 Interleukin1(IL1).......................................................................................................................... 1 3 Interleukin4(IL4).......................................................................................................................... 2 3 3 Interleukin6(IL6).......................................................................................................................... 2 Interleukin8(IL8)andchemokines ............................................................................................. 3 . 3 Interleukin10(IL10)..................................................................................................................... 3 3 Interleukin12(IL12)...................................................................................................................... 5 3 Interleukin18(IL18)...................................................................................................................... 5 3 GranulocyteMacrophageColonyStimulatingFactor(GMCSF).................................................... 5 3 TransformingGrowthFactor(TGF)......................................................................................... 6 3 3 Cytokinegenepolymorphismsandmalariaprotectionandpathogenesis.................................... 6 Interleukin10genepolymorphisms............................................................................................. 7 3 ThePresentStudy......................................................................................................................... 9 3 Rationale........................................................................................................................................ 9 3 StudyI............................................................................................................................................. 9 3 StudyII............................................................................................................................................ 1 4 Thespecificobjectives................................................................................................................... 3 4 ChapterTwo:MaterialandMethods..................................................................................44 Thestudyareaandpopulation...................................................................................................... 4 4 Microscopyfordiagnosisandestimationofparasitedensity ...................................................... 7 . 4 ParasiteDNAextraction,andestimationofP.falciparummultiplicityofinfection(MOI)............ 7 4 Preparationofthecrudeparasiteantigensanddeterminationofspecificantibodylevels......... 8 4 5 ExtractionofhumanDNAandgenotypeofinterleukin10(1082A/G)SNP................................ 0 Determinationofcirculatinglevelsofinterleukin10 .................................................................. 1 . 5 Statisticalanalyses......................................................................................................................... 2 5

ChapterThree:Results............................................................................................................ 4 5 StudyI............................................................................................................................................ 4 5 Thedemographicandclinicalfeaturesofthestudypopulation .................................................. 4 . 5 CorrelationofIgGantibodieswithage,parasitedensity,MOIandmsp2genotypesin thestudypopulation..................................................................................................................... 6 5 5 ComparisonofIgGantibodysubclasslevelsbetweenstudygroups............................................ 6

StudyII........................................................................................................................................... 1 6 Thedemographicandmalariologicalindicesofthestudygroups................................................ 1 6 Thecirculatinglevelsofinterleukin10inthedifferentstudygroups........................................... 1 6 Thefrequencyofallelesandgenotypesininterleukin10(1082A/G)singlenucleotide polymorphismamongclinicalgroups. .......................................................................................... 2 . 6 ChapterFour:DiscussionandConclusions.................................................................................. 6 6 StudyI............................................................................................................................................. 8 6 StudyII........................................................................................................................................... 1 7 Conclusions.................................................................................................................................... 4 7 Perspectives................................................................................................................................... 4 7 References..................................................................................................................................... 5 7 Appendices................................................................................................................................... 9 9

Dedication

To the memory of my father, the great teacher who taught me the love and respect of knowledge, to my mother, my wife and children and to all sisters and brothers, without their patience, understanding, support and most of all love, the completion of this work would not have been possible.

Acknowledgements
I am very much indebted to my supervisor, Professor Mustafa Idris Elbashir, The Dean

Faculty of Medicine University of Khartoum. His wide knowledge and way of thinking have been of great value for me since I knew him fifteen years ago. His understanding, encouraging and personal guidance have provided a good basis for the present thesis. I
would like to express my deep and sincere gratitude to Dr Hayder Geha whose advice, stimulating suggestions and encouragement helped me in all the time of research and writing of this thesis. I have furthermore to thank Professor Marita Troye-Blomberge of the Department of Immunology, Wenner-Gren Institute, Stockholm University who gave me the possibility to to do the necessary research work and to use departmental resources to complete this thesis. I want to thank my colleagues in the Department of Biochemistry, University of Khartoum and the Department of Immunology, Wenner-Gren Institute, Stockholm University, for the invaluable discussions and immense technical assistance. I am especially obliged to Dr. Thorya Moh Elhassan for her great help in difficult times, and to Nnaemeka C. Iriemenam and Amre Nasr

for the invaluable assistance. I would like also to thank all patients, their guardians/families for
their cooperation and the staff at El-Gedarif and El-Obied hospitals for their collaboration and help. I express my gratitude to field team, Adil, Faiz, Ihsan and Mustafa for the great help and assistance,

This study was partially supported by a MIM/WHO/TDR grant to Malaria Immunology and Pathogenesis Consortium in Africa (MIMPAC); (Grant ID A10622), and the International University of Africa, Khartoum, Sudan.

List of abbreviations ADCI AMA-1 APCs CM CSP DNA EDTA ELISA GLURP GM-CSF GPI ICAM-1 IFN- IgG IL iRBCs LT- MHC MIF MOI MSP NK NKT PBS PCR PECAM pRBC rs SM SMA SNP TGF- Th TNF : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : : Antibody-dependent cellular inhibition Apical membrane antigen-1 Antigen presenting cells Cerebral malaria Circumsporozoite protein Deoxyribonucleic acid Ethylene diamine tetraacetic acid Enzyme-linked immunosorbent assay Glutamate-rich protein Granulocyte-macrophage colony stimulating factor Glycosylphosphatidylinositol Intracellular adhesion molecule-1 Interferon gamma Immunoglobulin-G Interleukin Infected red blood cells Lymphotoxin- Major histocompatibility complexes Macrophage migration inhibitory factor Multiplicity of infection Merozoite surface protein Natural killer cells Natural Killer T cells. Phosphate buffer saline Polymerase chain reaction Platelet-endothelial cell adhesion molecule Parasitized red blood cells Reference sequence Severe malaria Severe malarial anemia. Single nucleotide polymorphism Transforming Growth Factor- T helper types of CD4+ T cells Tumor necrosis factor 8

TRAP UM VSA WHO

: : : :

Thrombospondin related adhesive protein Uncomplicated malaria Variant surface antigens World Health Organization

Abstract Malaria causes considerable morbidity and mortality among young children in malaria endemic areas worldwide. Anti-malarial antibodies are crucial determinants of Plasmodium falciparum infection outcome. In this study the levels of circulating immunoglobulin-G antibodies specific to crude Plasmodium falciparum antigens were determined by an indirect enzyme-linked immunosorbent assay in a hospital based casecontrol study, of children with varying malaria severity living in the rural area of Gedarif eastern Sudan and El-Obied west of Sudan. The aims were to investigate if immunoglobulin-G subclasses correlated with the clinical presentation of the disease and other malariometric indices among children exposed to seasonal and unstable Plasmodium falciparum transmission. The levels of circulating interleukin-10 and the genotypes of the single nucleotide polymorphism at position -1082 of IL-10 gene were also investigated in relation to malaria severity. The results revealed that individuals with severe malaria had significantly lower prevalence and levels of IgG1 and IgG4 as compared to those with uncomplicated malaria, suggesting a protective role for these molecules against severe malaria. The levels of IL-10 were lower in severe malaria patients as compared to those who presented with uncomplicated malaria, but no significant association was found between the specific alleles or genotypes in the -1082 A/G polymorphic site and any particular form of clinical malaria. The present data argue for protective roles for IgG1 and IgG4 antibodies and IL-10 against severe malaria, but find no evidence supporting significant effect of the different alleles or genotypes at position (-1082A/G) of the IL-10 gene on protection from SM. 10

:
. . . - . -01 )-2801( .

-1 4 -01 . -01 , )-2801( -01 . 1 4 -01 )-2801( -01 .

11

List of Figures
1. Figure 1.1 2. Figure 1.2 3. Figure 3.1. The Global Distribution of Malaria Transmission Risk. The life cycle of Plasmodium falciparum.

Page
15 18

General comparisons of IgG subclass levels between the study groups. Comparisons of IgG levels between the patient groups. Levels of circulating IL10 in the study groups of malaria patients and healthy controls.

59 60

4. Figure 3.2. 5.

Figure 3.3.

67

List of Tables 1. Table 3.1. The demographic and malariological indices of the study groups Antimalarial immunoglobulin-G subclass sera reactivity and risk of severe malaria. Demographic features of study II population The Hardy-Weinberg Equilibrium for the interleukin-10 (-1082 A/G) SNP in the study groups Estimation of odds ratios, assuming different modes of A/G alleles inheritance. 66 56

2. Table 3.2.

61 64

3. Table 3.3. 4. Table 3.4 5. Table 3.5

65.

12

Chapter one

Introduction to Malaria Immunity and Pathogenesis

13

1.1.

Introduction to malaria:

1.1.1. The global burden and distribution of malaria. Malaria is by far the worlds most important parasitic disease with a staggering global impact [1]. Current estimates indicate that, malaria threatens the lives of 3.2 billion, causes between 350-500 million clinical episode and kills more than one million people each year [2, 3]. Most morbidity and mortality occur among the vulnerable children and women at child bearing age in the poorest communities of sub-Saharan Africa (Fig 1.1) [4]. Conventional approaches used to assess the impact of malaria seem inadequate, and the reported figures of incidence, morbidity, and mortality are likely to be largely underestimated [5]. Recent innovative approaches, exploiting satellite imagery and climatic modeling of malaria transmission levels, further support this view [6-9]. In addition to its direct heavy toll on health, malaria has immense socio-economic impact, and it has long been recognized that malarious communities are impoverished communities. This is clearly evidenced by the fact that, malaria endemic countries are the least developed ones with the lowest rates of economic growth [10]. The disease is responsible for around 1.3 % reduction in the overall annual economic growth rates in countries with high malaria endemicity. Malaria impedes economic through its effects on worker productivity, school performance and medical costs [10-13]. The global distribution and clinical patterns of malaria are widely variable and are largely determined by the local levels of parasite transmission which are influenced by several biological and environmental factors [7, 14]. Genetic characteristics of the host, the parasite, and the vector as well as the environmental factors including socioeconomic, and the malaria control measures contribute differently to the disease distribution. 14

Conventionally, levels of transmission are measured by estimating the entomological inoculation rate [15-18]. Alternatively, the epidemiology of malaria in a specific geographical area can be defined in terms of spleen enlargement and parasite rates among children [19]. More conveniently, the pattern of malaria transmission can be described as stable or unstable based on the fluctauation in the annual transmission. However, local epidemiologic features of malaria transmission may be drastically altered by environmental and/or social disturbances causing frequent epidemics.

Figure. 1.1 The Global Distribution of Malaria Transmission Risk; Malaria is endemic in regions of 107 contries, mainly in sub-Saharan Africa.
Source: The world malaria report 2005; http://www.rbm.who.int/wmr2005/html/map1.htm.[20]

15

1.1.2. The biology of malaria. Malaria is a mosquito-borne disease that affects several vertebrate species including rodents, reptiles, birds and primates. The infection is transmitted by the bite of an infectious female Anopheles mosquito carrying a protozoan parasite of the genus Plasmodium. There are over 175 species of plasmodia currently recognised, many of these are known to cause malaria in non-human vertebrates [21]. However, humans malaria can solely be caused by four species of Plasmodium: P. vivax, P. ovale, P. malariae, and P. falciparum. Among these, P. falciparum is the most virulent and is responsible for almost all cases of severe disease and deaths. P. falciparum is the most common parasite in tropical Africa, and exists in all malaria endemic regions worldwide, while other species are variably distributed in different geographical locations [22, 23]. More than four hundred species of Anopheles were identified, but only seventy of these are known to naturally transmit malaria to various vertebrates. [24, 25]. Human malaria parasites are transmitted by a few Anopheles species, of which, An. gambiae, An. arabiensis, and An. funestus are the most prevalent in Africa whereas other species such as An. albimanus, An. culicifacies, An. dirus, An. anthropophagus are found in other parts of the world [26].

1.1.3. Plasmodium falciparum life-cycle. Plasmodium falciparum has a remarkably complex life cycle (Figure 1.2). The infection begins when sporozoites are released into the blood stream by a female Anopheles mosquito. The inoculated sporozoites circulate briefly before they invade the liver parenchymal cells where they undergo asexual replicative phase called exo-erythrocytic

16

schizogony. This liver stage of parasite lasts for approximately a week before releasing merozoites which invade red blood cells (RBCs). Erythrocyte invasion is crucial for parasite survival, it provides a rich source of nutrients in a niche (the parasitophorous) that is largely protected from host immune defense. The erythrocytic schizogony involves rapid successive mitotic divisions and generate thousands of merozoites which are released to re-invade new erythrocytes. It is this stage that is responsible for the acute clinical episodes of malaria. After a period of 7 to 15 days of asexual replicative cycles in RBCs, a small proportion of the merozoites undergo transformation into male and female gametocytes, the micro- and macro-gametocytes respectively. The half-life of the mature gametocyte in the blood is about 2.4 days during which it could be taken up by a feeding Anopheles mosquito. A few minutes after ingestion by the mosquito, the micro-gametocytes exflagellate in the insect midgut, each forms eight micro-gametes which move to fertilize the female macro-gametes. Eighteen to 24 hours after fertilization, the zygote elongates into an ookinete which traverses the membranes and the epithelial cells of the mid-gut and transforms into oocyst. The oocyst enlarges progressively as the nucleus divides repeatedly to form sporozoites. Mature sporozoites exit the oocyst to the coelomic cavity and migrate to salivary glands awaiting transfer to a new victim [27-29].

17

Fig 1.2. The life cycle of Plasmodium falciparum.


Modified from the Center of Biologics Evaluation and Research, US, Food and Drug Administration: www.fda.gov/Cber/blood/malaria071206sk.htm. Accessed on the 3rd of June 2008.

An infectious mosquito injects sporozoites into the blood stream of the human host, which then circulate with the blood to the liver. An exo-erythrocytic stage in the liver leads to release of merozoites into the blood stream. The merozoites invade the red blood cells and undergo an asexual replicative cycle (the erythrocytic schizogony) to produce multiple merozoites capable of re-invading new red blood cells. Occasionally, erythrocytic trophozoites transform into gametocytes which are released into the blood stream and are eventually taken up by the mosquito, and then undergo the sexual stage of development (sporogenic cycle).

18

1.1.4. The malaria disease. 1.1.4.1. Uncomplicated malaria.

The clinical presentation of P. falciparum malaria is quite variable and generally mimics that of many other diseases caused by bacteria, rickettsia, and viruses. Initially malaria can have a gradual course of onset with non-specific symptoms including slight fever, anorexia, malaise and headache [30, 31]. The classical manifestations of the acute P. falciparum include fever with a feeling of coldness and chills followed by profuse sweating. Generalized symptoms such as backache, myalgia, aching joints, and gastrointestinal symptoms, i.e., anorexia, nausea, vomiting, abdominal pain, and mild diarrhea may also be found. Unless the malaria infection in young children and nonimmuned individuals is diagnosed early and promptly treated, the condition may progressively deteriorate to a severe life-threatening disease with complex multi-system dysfunction [32].
.

1.1.4.2.

Severe complications of malaria.

Severe malaria (SM) is a complex multi-system disorder presenting with a considerably variable range of clinical manifestations [33]. In children and non-immune adults, many of severe complications of P. falciparum may occur in combination or as isolated manifestations. Such complications include cerebral malaria (CM), severe malarial anemia (SMA), respiratory distress, renal failure, and serious metabolic derangements including hypoglycemia and lactic acidemia. However, CM and SMA are the leading causes of malaria-related deaths in Africa. [31, 34-36].

19

1.1.4.2.1. Cerebral Malaria (CM). Cerebral malaria (CM) is a syndrome of impaired consciousness associated with a P. falciparum malaria infection. Clinically, the term is loosely used to describe those at high risk of death due to SM. However, standardized criteria defining CM have been established by a panel of WHO experts, accordingly, the strict research definition of CM; is a condition of unarousable coma for more than 30 minutes after a generalized convulsion, confirmed by P. falciparum infection and exclusion of other causes of coma [37-41]. In malaria endemic areas, CM accounts for up to 10% of all hospitalized cases of P. falciparum malaria, and for more than 80% of fatal cases. It occurs predominantly in patients with little or no parasite specific immunity, i.e., children under five years age in endemic areas or individuals coming from malaria-free areas.

1.1.4.2.2. Severe malarial anemia (SMA). Severe malarial anemia (SMA) is defined as a haemoglobin concentration 50g/L (or a haematocrit 15%) in a patient with a P. falciparum parasitaemia with normocytic indices. The prevalence of SMA has increased during the last decade as a result of spread of chloroquine-resistant P. falciparum across the world [42-44]. SMA has a profound effects on communities in malaria endemic regions [45], anemia is associated with an increased risk of death [32], impaired cognitive and motor development [46-48], retarded growth [49], diminished immune protection [50], and reduced physical work capacity [51].

20

1.2.

Immunity to malaria.

The immune system plays an important role in the host defense against infection with P. falciparum, and in limiting the extent of infection once it occurs. Although most

individuals living in malaria endemic areas are equally exposed to infectious bites of the Anopheles mosquitoes, a large proportion of the infected individuals either spontaneously resolve the infection, or remain asymptomatic i.e. harboring some parasitemia without significant clinical disease. A small proportion of infected individuals, mostly young children, develop clinical symptoms of malaria, and only less than 5% of these clinical cases proceed into severe life-threatening complications [52]. The clinical pattern of malaria is markedly influenced by the age-dependent acquisition of anti-malaria immunity [53]. Different components of the immune system including both antibody-dependent and cell-mediated mechanisms operate at different stages of the parasites life cycle (figure 1.2) [54]. However, protective immunity to malaria takes years of exposure to infection before it develops significantly [55], and even with repeated infections, this immunity is rarely sterile, and is species-, variant-, strain-, and stage-specific [56, 57]. This inefficient and slow acquisition of anti-malaria immunity is attributed to the P. falciparum virulence and variability [54].

1.2.1. The innate immune responses to malaria. Early interactions between sporozoites and host cells produce no significant systemic immune response and disease. Only merozoites are the stages that elicit strong immunoinflammatory responses. Although not fully understood, several humoral components of innate immunity comprising the complement proteins [58], C-reactive proteins [59-61]

21

and mannose binding lectins [62-64] were implicated in protection from malaria. However, the principal innate effector mechanisms against malaria parasites are mediated by cellular components including monocytes-macrophages, natural killer cells, natural killer T cells and / T cells [65]. The activation of monocytes-macrophages is one of the first events in the innate resistance to P. falciparum infection. Circulating monocytes and tissue resident macrophages in the liver, spleen, and other tissues are directly involved in phagocytosis and clearance of the infected red blood cells (iRBCs) [66, 67]. Malarial antigens such as the hemozoin pigment and/or the membrane-anchored glycosylphosphatidylinsitol, phagocytosed by the macrophages or acting through surface receptors, activate these cells to produce a series of pro-inflammatory mediators and cytokines. The inflammatory agents produced may have a direct anti-parasite activity or contribute to the activation of other effector immune mechanisms. However, these cells are also implicated in pathogenesis of severe malaria [68, 69]. The natural killer (NK) cells interact with iRBCs and liberate pro-inflammatory cytokines, particularly IFN- [70-74]. In non-immune donors, NK cells can kill plasmodia-infected erythrocytes in the absence of opsonizing antibodies through direct cell-cell interaction [72-76]. Early production of IFN- by NK cells stimulated by IL-12 during malaria infection is associated with a better prognosis of the disease [70, 72, 77]. Additionally, the interaction of NK with iRBCs induces production of the chemotactant IL-8 [78], suggesting a role for NK cells in recruitment and activation of other leukocytes during malaria infection.

22

Natural killer T cells (NKTs) are a specialized subset of lymphocytes that share properties of both T cells and natural killer (NK) cells. They differ from conventional T cells in that; they express the distinct NK.1.1 marker and possess a unique single invariant V14+ antigen receptor that recognizes lipids and glycolipids presented by CD1d molecules, rather than peptide-MHC complexes [79-81]. Experimental models have indicated a role for NKTs in conferring protection against murine malaria, and NKT cells were linked to malaria-associated splenomegaly and expansion of the splenic B cell pool forming the parasite-specific antibody. [82-85]. The gamma/delta (/) T cells constitute a minor subset of T lymphocytes characterized by an antigen receptor formed of and chains rather than the common and chains that make up the antigen receptors on the conventional T cells [86]. Peripheral / T cells are strongly up-regulated during malaria, and they respond to malaria antigens by proliferation and lymphokine production [87-89]. Activated / T cells but not / T cells from malaria-nave donors were shown to inhibit parasite replication in eryhrocytes in vitro, indicating an important innate protective a role against malaria [90].

1.1.1. Adaptive immune responses to malaria. 1.1.1.1. Anti-sporozoite immunity.

Protective immunity against malaria can be induced in animal models and humans following immunization with irradiated sporozoites.[91-93]. Although it confers complete resistance to subsequent sporozoite challenges, such immunity is stage-specific and does not resist challenges with blood stage Plasmodium. Anti-sporozoite antibodies are the principal effectors that promote phagocytosis and block hepatocytes invasion [94].

23

Such antibodies are mainly directed against the most abundant sporozoite surface proteins, the circumsporozoite protein (CSP) and the thrombospondin related adhesive protein (TRAP) as well as several other sporozoite and liver stage antigens. Most adults in malarious areas were found to possess high titer of antibodies to the CSP of P. falciparum [95], and the immune sera obtained from sporozoite-immunized human volunteers, and naturally infected individuals, were found to block the invasion of P. falciparum sporozoite into hepatocytes [96].

1.1.1.2.

Immunity to liver stages of malaria parasite.

Several cellular mechanisms are implicated in generating protective immunity against the liver stages of Plasmodium [97-101]. T cells were shown to be involved in the generation of protective immunity in rodents immunized with irradiated sporozoites.[102, 103], and the passive transfer of CD8+ T cells and CD4+ T cells clones derived from sporozoite immunized mice was found to confer protection [104-107]. Cytotoxic T lymphocyte response to CS proteins was observed in field studies and believed to protect against preerythrocytic stages of malaria [108].

1.1.1.3.

Immunity to blood stages of malaria parasite.

As previously noted, invasion of red cells is the key step in malaria infection and is the cause of clinical disease, hence erythrocytic stages of P. falciparum are likely to be the most important targets for protective immune responses. Protective immunity against erythrocytic stages involves both antibody-dependent and cell-mediated immunity [109, 110].

24

1.1.1.3.1.

Antibody-mediated immunity to blood stages of P. falciparum.

Passive transfer of anti-Plasmodium specific antibodies to non-immune mice, and treatment of P. falciparum infected patients with immunoglobulins extracted from naturally immune individuals, result in reduction of parasitemia and clinical symptoms [111-115]. Moreover, studies in gene-targeted, and B-cell deficient mice demonstrated that B cells are essential for the resolution of blood stage infection of murine malaria, thus, protection from malaria was thought to be mainly an antibody-mediated type of response [116, 117]. In people living in malaria endemic areas, exposure to repeated infections with different variants of P. falciparum induces production of polyclonal antibodies, predominantly IgM, IgG isotypes [118-122], and the acquired protective immunity to malaria correlates with the individual age and the level of antibodies against the asexual blood stage antigens of the parasite [14, 123-125]. Such epidemiologic observations explain the slow acquisition of protective immunity to malaria during the first years of life in parallel with the expansion of the antibody repertoire against variant surface antigens (VSA). Accordingly, a clinical episode in a child living in a malaria endemic area represents an infection with a parasite expressing VSA not recognized by the existing repertoire of antibodies.[121, 126-129]. Early studies have clearly demonstrated the predominance of immunoglobulin-G (IgG) in resolution of the blood stage infection, and several antibody-mediated mechanisms of parasite growth inhibition and killing have clearly been elucidated [130-132]. IgG antibodies directed against merozoite surface proteins enhance various effector mechanisms including; opsonization, phagocytosis [132-134], complement-mediated

25

damage of merozoites [135], blockage of RBC invasion [136, 137], and antibodydependent cellular inhibition (ADCI) of parasite growth [132, 138].

1.1.1.3.2. Cell-mediated immunity to blood stages of P. falciparum. The first observations showing the importance of T lymphocytes in recovery from malaria infections were made by Brown and co-workers in their experiment with thymectomized rats [139]. Later, it was shown that B cell deficient mice can be immunized by natural infection with Plasmodium, and that adoptive transfer of CD4+ Tcell clones could confer protection against malaria [140-142]. However, because erythrocytes do not express MHC molecules, it seems unlikely that T cells could have a direct cytotoxic effects on infected erthrocytes. Alternatively, it was suggested that T cells respond to malaria antigens by cytokines secretion. In consistence with this view, it was shown that stimulation of T cells by malarial antigens presented by antigen presenting cells (APCs) results in release of various cytokines, and experiments with gene-knocked-out mice, have shown that, immune responses to blood stages of Plasmodium are mostly mediated by IFN--secreting T lymphocytes [99, 116, 143]. Both Th1 and Th2 types of CD4+ T cells were implicated in malaria protection and pathogenesis in mice. It has been shown that; CD4+ T cells of the Th1 type which produce IL-2 and IFN-, are important during the acute phase of murine infection with Plasmodium chabaudi chabaudi AS whereas Th2 type of these cells is required during later phases to provide help to B-cells through IL-4 secretion.[144, 145]. The secretion of IFN- by T lymphocytes is believed to induce the cytophilic IgG blood-stage-specific antibodies and assist in antibody-dependent cellular inhibitory mechanisms [138]. Additionally, the CD4+ Th1 cells can kill parasites in the absence of antibodies through 26

monocyte-macrophage activation with the subsequent generation of reactive oxygen and nitrogen intermediates [146, 147]. In humans, CD4+ T cells were shown to inhibit parasite growth in vitro [148], and the production of IFN- by CD4+ T cells in response to specific erythrocytic antigens was associated with protection from malaria in field studies in Africa [149, 150]. In the other hand, the CD8+ T cells seem to play no significant role in protection against blood stage malaria in mice. Consistent with this, depletion of CD8+ T cells during primary blood stage infection did not diminish the ability of the mice to resolve infection. [151, 152]. However, experiments with mice treated with anti-CD8+ mAb showed significantly higher levels of parasitemia than untreated control mice, suggesting some role for CD8+ cells in containing the blood stages of murine malaria [153]. However, in humans, CD8+ T cells may have indirect regulatory effects on the immune responses to blood stage infection during acute malaria [154, 155].

1.2.

The immuno-pathogenesis of severe malaria. Malaria is a complicated syndrome and the clinical outcome of infection with P. falciparum is determined by several factors. Such factors include the intensity of parasite transmission, the parasite virulence, together with numerous host genetic, nutritional, and immunological characteristics [35, 52, 156-165]. However, the mechanisms leading to severe life-threatening complications of malaria are not fully understood [52, 166].

1.2.1. The pathogenesis of cerebral malaria. The mechanisms underlying the pathogenesis of CM remain controversial, and several hypotheses were set to explain the cause of this deadly syndrome [167-170]. The

27

mechanical

(sequestration)

hypothesis

assumes

that

P.

falciparum-parasitized

erythrocytes bind to the post capillary venules, causing obstruction of blood flow, and thereby decrease tissue perfusion and impair wastes disposal leading to brain pathology [169, 171]. Histological examination of brain sections from fatal cases of CM revealed sequestration of iRBCs in multiple organs and tissues [172]. However, due to lack of sequestration in some patients with cerebral symptoms, and the fact that most patients with CM fully recover without evidence of ischaemic damage, the mechanical hypothesis cannot adequately explain the cause of CM. Alternatively, it has been suggested that; malaria toxins bind to and stimulate monocytes to secrete pro-inflammatory cytokines [156, 173, 174]. Although, the cytokines themselves are not harmful, they may induce production of nitric oxide (NO) [175-178] which diffuses through the blood-brain barrier and interferes with the synaptic function as do general anesthetics, leading to a state of reduced consciousness [179]. Several malarial antigens, such as the membrane-anchored glycosylphosphatidylinositol (GPI) and the by-product of parasite metabolism the hemozoin, can cause cytokines release from various host immune cells [180]. The released cytokines up-regulate the expression of various endothelial adhesion receptors such as ICAM-1, PECAM-1/CD31, and thus enhance cytoadherence of iRBCs and sequestartion [179, 181]. Sequestered parasites in the brain vasculature may locally release more toxins and induce the production of inflammatory mediators that disrupt the brain function and metabolism leading to the CM. Most recently van der Heyde et al., suggested a unified hypothesis to explain the etiology of CM [170]. They attributed CM to a combination of parasite-host specific interactions that promote iRBCs adhesion and sequestration in the microvasculature of various organs

28

and tissues causing the induction and release of cytokines and several other bioactive molecules which trigger excessive immuno-inflammatory responses that disturb the homeostasis [170, 182].

1.2.2. The pathogenesis of severe malaria anemia. Severe malaria anemia (SMA) is the most common complication of malaria in holoendemic areas and contributes substantially to the mortality and morbidity from this disease. The pathogenesis of SMA is multi-factorial, and is only partially understood. It involves both accelerated clearance of erythrocytes from the circulation and suppression of hematopoeisis [164, 183, 184]. During acute P. falciparum infections, parasite replication results in an increased rate of RBCs destruction, but this cannot solely account for the marked reduction in hemoglobin levels observed in patients with SMA. In contrast, children with SMA were shown to have lower peripheral parasite densities than parasitemic children without anemia [185, 186] Moreover, an increased rate of clearance of uninfected erythrocytes associated with low reticulocyte levels was also observed in these patients suggesting involvement of inflammatory processes in the pathogenesis of SMA [187-194]. The analysis of microarrays profile gene expression in a murine model of malaria has shown repression of erythroid-associated transcripts during plasmodial infection, providing additional confirmation for a role of the inflammatory mediators in the etiology of SMA [195]. High circulating levels of the proinflammatory cytokines TNF- and IFN- and low levels of the anti-inflammatory IL-10 were associated with SMA, supporting the hypothesis that specific cytokine disbalance may lead to SMA [196198]. Moreover, impairment of IL-12 response was shown to contribute to the development of SMA [199, 200], and low plasma levels of IL-12 have repeatedly been 29

reported in patients with SMA [201-204]. Macrophage migration inhibitory factor (MIF) has also been suggested to play an important role in the pathogenesis of malarial anemia [68], and elevated plasma MIF concentrations were associated with suppression of erythropoiesis and enhanced severity of anemia in murine models [68, 69]. However, MIF production by monocytes in children with acute malaria was shown to decline progressively with increasing anemia severity (167).

1.2.3. The role of cytokines in immuno-pathogenesis of malaria. The importance of cytokines in protection from and pathogenesis of infectious diseases is well established, and altered production of cytokines is widely held as a mechanism responsible for a variety of immuno-inflammatory diseases [33, 205-207]. The first speculations that malaria parasites elicit a systemic inflammatory response that causes disease appeared early in 1940s [208, 209]. The striking similarities between the patho-physiology of acute P. falciparum malaria and septic diseases together with the detection of elevated levels of circulating inflammatory mediators in animals receiving endotoxin and in animals infected with malaria parasites led to the notion that; macrophage derived factors may be responsible, at least in part, for the pathology observed in malaria [164]. Pro-inflammatory cytokines are now recognized causal mediators of malaria fever and low levels of anti-inflammatory cytokines such as IL-10 and transforming growth (TGF-) and/or increased ratios of pro-inflammatory to antiinflammatory factor- cytokines have repeatedly been associated with increased risk of high fever and severe malaria [196, 198, 202, 210-213] 1.2.3.1. Interferon- (IFN-).

30

There is a growing consensus that IFN- plays a central role in protection from malaria [71, 72, 147, 149, 173, 214-216]. IFN- induces the effector mechanisms essential for control of both pre-erythrocytic and blood-stage malaria infections [99]. Treatment of mice with exogenous IFN- delayed the onset of parasitemia, decreased the parasitemia and increased mice survival [173, 216]. Conversely, treatment of Plasmodia-infected mice with neutralizing monoclonal antibodies to IFN- exacerbated the course of infection [217, 218]. Moreover, studies with IFN- knocked-out mice confirmed the critical role of IFN- in the protective immunity to malaria [219, 220]. In clinical studies, elevated levels of serum IFN- were found in patients with acute malaria [221-226], and individuals heterozygotic for an IFN-R1 polymorphism had a lower incidence of CM and death in Gambia [227].

1.2.3.2.

Tumor necrosis factor (TNF-).

Tumor necrosis factor (TNF-) plays pivotal roles in many inflammatory processes, and is implicated in pathogenesis of various acute and chronic illnesses, encompassing both infectious and non-infectious diseases [164, 228-230]. Ample evidence demonstrated the importance of TNF- in the immune response to malaria infection [33, 175, 222, 231238], and several antigens of P. falciparum blood stage parasites as well as hemozoin granules induced TNF- production in peripheral blood monocytes in vitro. The glycosylphosphatidylinositol (GPI) is the major malaria toxin responsible for the TNF- release and induction of fever [180]. Raised levels of plasma TNF- were found in the general population during the malaria transmission season, and were associated with acute and severe malaria [233, 239-245]. TNF- upregulates the ligands that bind iRBCs

31

and promotes sequestration in the brain capillaries, a process that is central to pathogenesis of CM [246, 247]. In addition, hyperlactaemia [248], hypoglycaemia [222] and fever [241] in malaria patients have all been correlated with elevated levels of serum TNF-. However, the failure of TNF- neutralizing agents to decrease the incidence of clinical CM and the disappointing outcome of trials with pentoxifylline and steroid therapy for CM raise doubts about the role of TNF- in the pathogenesis of CM [249, 250]. Moreover, there is an emerging view claiming that lymphotoxin-, and not TNF-, may be the principal mediator of human CM [238, 251].

1.2.3.3.

Lymphotoxin (LT-).

There is evidence indicating a distinctive role for LT in malaria [238, 251-253]. Like TNF- , LT- has been detected in human malarial serum, and was shown to induce high levels of IL-6 and causes hypoglycemia in a mouse model [253]. LT- but not TNF- was found necessary for IFN- production during malaria infection [251].

1.2.3.4.

Interleukin-1 (IL-1).

Interleukin-1 (IL-1) is a family of pro-inflammatory cytokines which comprises IL-1, IL-1, and the IL-1 Receptor antagonist (IL-1RA) [254]. Production of IL-1 by macrophages, monocytes and dendritic cells enhances T cell-dependent antibody production and increases the expression of adhesion molecules on endothelial cells to enable transmigration of leukocytes to sites of infection [254-256]. Both IL-1 and IL-1 are potent pyrogens that are implicated in the causation of malaria fever. Administration of recombinant IL-1 reduced parasitemia and protected mice from lethal cerebral malaria,

32

and elevated serum levels of IL-1 were found in malaria patients in a holoendemic area [257].

1.2.3.5.

Interleukin-4 (IL-4).

Interleukin-4 is a pleiotropic cytokine that plays a critical role in both the proliferation of lymphocytes and regulation of lymphokine production [160, 258]. IL-4 was shown to inhibit Th1 cell differentiation and to induce a reversion of developing Th1 cells to the Th2 lineage through inhibition of IFN- [259]. Early addition of IL-4 to cultures of human CD4+ T cells was shown to impair IL-2 and IFN- synthesis by these cells suggesting an immune-suppressive effect [260]. However, addition of exogenous IL-4 to in vitro cultures is required for the differentiation of naive CD4+ T cells into IL-4secreting Th2 cells [261-264], indicating a critical role in the development of Th2-type immune responses and possibly in the regulation of immunoglobulin switching to IgE and IgG isotypes [265-267]. Thus, IL-4-producing T cells were suggested to induce the production of malaria specific IgG and elevated levels of IL-4 were detected in parasitemic individuals living in malaria endemic areas [257]. It has also been shown that; IL-4 receptor expression on CD8+ T cells is required for the development of protective memory responses against liver stages of malaria parasites [268].

1.2.3.6.

Interleukin-6 (IL-6).

Interleukin-6 is one of the most important mediators of fever and is required for the activation, proliferation, and survival of T cells and dendritic cells [269-275]. In malaria, IL-6 was shown to mediate the anti-parasitic activity of IL-1 on the liver stages of malaria parasite [276], and patients with malaria fever were found to have 33

elevated serum levels of IL-6 [240, 245, 277-279]. Interestingly, higher serum levels of IL-6 were found in severe malaria than in uncomplicated malaria patients [242, 280, 281] and raised levels of this cytokine were associated with hyperparasitemia, jaundice, and shock in Vietnamese patients [282].

1.2.3.7.

Interleukin-8 (IL-8) and chemokines.

The role of IL-8 in malaria pathogenesis is not clearly understood and conflicting results were reported [245, 283-285]. However, elevated serum levels of IL-8 have been noted during acute malarial attack [283, 284].

1.2.3.8.

Interleukin-10 (IL-10).

Interleukin-10 (IL-10) was initially identifed as a cytokine synthesis inhibitory factor (CSIF) produced by mouse Th2 cells and inhibits Th1 cells cytokine production [286]. It is primarily produced by cells of the monocyte/macrophage lineage and to a lesser extent by activated T and B cells [287-289]. IL-10 has extensively diverse effects mediated through specific cell surface receptor complex (IL-10R and IL-10R) expressed on most hemopoietic cell types. It acts as a suppressor of immune responses, principally through down-regulating the expression of the MHC class II, the costimulatory molecules and cytokines genes in the antigen presenting cells [289-292]. For, example, IL-10 is known to inhibit the secretion of IFN- by Th1 cells [286, 293], and of TNF- and MIF by monocytes and macrophages [290, 294], and it modifies the expression of chemokine receptors [295]. In various experimental models, induction or administration of IL-10 suppressed the antigen-specific T-cell proliferation and contributed to the establishment of persistent infection by a number of pathogens including HIV [296] Mycobacterium 34

tuberculosis [297], Listeria monocytogenes [298], Trypanosoma cruzi [299]. However, although generally precieved as an anti-inflammatory cytokine, IL-10 could have quite opposite effects on some specific models [300, 301]. For instance, IL-10 was found to stimulate the expression of MHC-II molecules and the production of immunoglobulins in B cells [302], and to augment the proliferation and cytotoxicity of NK cells when combined with IL-18 [303]. Increasing evidence demonstrates the role of IL-10 in malaria protection and pathogenesis [210, 226, 245, 304-307]. Murine and in vitro studies have demonstrated IL-10s ability to inhibit TNF- production in response to malarial antigens [221], and high levels early during malaria infection were found to inhbit Th1 type of immune responses in both mice and humans. These raised levels were also associated with less effective clearance of Plasmodial parasites and subsequently the development of severe malarial complications, particularly SMA [210, 245, 305, 308, 309]. In contrast, IL-10 has been suggested to play a protective role against experimental cerebral malaria [307, 310], and clinical studies associated insufficient IL-10 with acute and severe malaria [196, 198, 282]. Moreover, it was demonstrated that, in individuals with SMA, plasma levels of TNF- tend to exceed those of IL-10, and low ratios of IL-10:TNF were found to be a risk factor for both CM and SMA, whereas higher ratios were more frequent among hyperparasitemic individuals [311]. Interestingly, low circulating IL-10 associated with high pro-inflammatory cytokines levels was noted in comatose patients with CM [282], suggesting protective effect for IL-10 against SM. Therefore, while it may be detrimental by decreasing cellular immune responses to parasite, IL-10 may equally be beneficial in preventing excessive inflammation that underlies the SM. It is currently

35

assumed that not only the levels but also timely production of IL-10 is crucial for controlling the parasite growth and replication as well as for preventing excessive inflammatory responses to P. falciparum.

1.2.3.9.

Interleukin-12 (IL-12) .

Interleukin-12 plays an important regulatory role in both the cell-mediated [201-203, 213, 312-315] and antibody-mediated [316] immune responses to malaria. IL-12 has been inversely associated with disease severity in experimental and human malaria. Administration of IL-12 decreased mortality and reduced parasitemia in mice through a mechanism which appears to be critically dependent on IFN- [201, 203, 315, 317, 318]

1.2.3.10.

Interleukin-18 (IL-18). have been found in adult patients with uncomplicated P.

Elevated levels of IL-18

falciparum malaria, and these levels were found to decrease with recovery from disease [319]. Higher levels of IL-18 were found in children with mild malaria than in children with a severe malaria, suggesting a protective effect against severe malaria.[202, 320]). Impaired IL-18 response was shown to contribute to the development of SMA in areas of holoendemicity for malaria [204].

1.2.3.11.

Granulocyte-Macrophage Colony Stimulating Factor (GM-CSF).

Granulocyte-macrophage colony stimulating factor (GM-CSF) gene-knocked out mice were found to be more susceptible to infection with P. chabaudi AS than wild-type mice [321], indicating a protective role against murine malaria. GM-CSF was shown to act

36

synergistically with TNF- in priming neutrophils for enhanced phagocytosis and killing of asexual stages of P. falciparum and to up-regulate expression of cytokines, complement proteins and Fc receptors [322]. Incorporation of a plasmid encoding murine GM-CSF in a DNA vaccine against the CSP of P. yoelii enhanced the efficacy of the vaccine [323]. However, elevated serum levels of GM-CSF were recorded in severe human P. falciparum malaria [223, 324].

1.2.3.12.

Transforming Growth Factor- ( TGF-).

Transforming growth factor- (TGF-) plays a key role in limiting malaria pathology, and reduced levels of TGF- were associated with severe malaria [211, 325, 326]. Increased mortality was observed among IL-10/ mice when infected with the normally non lethal P. chabaudi [265], and treatment of P. berghei infected mice with anti-TGF- antibodies increased the parasite virulence, suggesting a protective effect for TGF- against severe disease [326, 327]. Conversely, treatment of these mice with recombinant TGF- resulted in upregulation of IL-10 and down-regulation of TNF- and extended survival time [326, 327]. However, induction, or administration of TGF- during early, acute blood-stage infection suppresses protective proinflammatory cytokine responses, leading to unrestrained parasite proliferation and increased disease severity [328-330]. 1.2.4. Cytokine gene polymorphisms and malaria protection and pathogenesis. The search for genetic variation associated with disease is an exciting area of medical investigations that has rapidly expanded during the last decades and contributed to better understanding of mechanisms of protection and pathogenesis of many diseases. Genetic markers in cytokine genes are becoming widely used in studies of immune-mediated

37

disease, and it is becoming apparent that inherited differences in cytokine genes contribute to phenotypic variations, risks of disease and responses to the environmental stimuli including infectious challenges [331-340]. Interindividual differences in cytokine profiles appear to be due mainly to allelic polymorphism within regulatory regions of the cytokine genes. The biallelic mutations, the single nucleotide polymorphisms (SNP), are the most common source of variations in the human genome [333]. Analyses of interindividual differences and genetic polymorphisms in association with malaria protection and pathogenesis may provide insight into new strategies for fighting malaria. [341, 342]. Examples of the clearly demonstrated associations of malaria risk with cytokine gene polymorphisms include two substitutions at the positions -308 [343, 344] and -376 upstream of the transcriptional start of the TNF- gene [244]. Both polymorphisms were identified as independent risk factors for cerebral malaria in African populations[243, 244, 345]. A long list of reports associating certain SNPs in different cytokines genes with malaria is currently available. However, the following discussion is confined to polymorphisms in IL-10 which is investigated in the current study.

1.2.4.1.

Interleukin-10 gene polymorphisms.

The IL-10 gene promoter is highly polymorphic, it contains two informative microsatellites, IL10.G and IL10.R and eleven single nucleotide polymorphisms (SNPs), of which the three proximal ones located at positions (-1082 G/A [rs 1800896], -819 C//T [rs 5289772] and -592 C//T [rs5289771]) are the most frequent points of mutations [337, 346-355]. Differences in the genotypes and allele frequencies in these SNPs were

38

observed between different populations. These differences may influence the outcomes of certain infections in different populations [356, 357] Interestingly, the -1082G allele was associated with higher production of IL-10 compared to the A allele both in vitro and in vivo [347, 358, 359], whereas polymorphisms at positions -819 and -592 appear to have no influence on IL-10 production [347]. Several studies have demonstrated the relevance of these SNPs to susceptibility and/or severity of a number of diseases including the inflammatory bowel disease [360-362], psoriasis [363-365], primary Sjogrens syndrome [366], rheumatoid arthritis [358, 367], myocardial infarction [368], coeliac disease [369] and systemic lupus erythematosus [353, 370]. Similarly, several infectious diseases were associated with certain alleles in these SNPs. Genetic predisposition to high IL-10 expression has been reported to be associated with a higher rate of mortality in meningococcal disease [334, 371]. Chronically infected hepatitis C patients who are genetically predisposed to high IL-10 production were reportedly less likely to benefit from IFN- therapy [372]. Moreover, associations have been reported between IL-10 polymorphisms and Epstein Barr virus infection [373], leprosy [374] severe malaria [375], and recurrence of hepatitis C in liver transplant patients [376]. Of particular significance, is the A/G transition at position -1082 which has been associated with rheumatoid arthritis [367], systemic lupus erythematosus [370], psoriasis [365], coeliac disease [369], graft-versus host disease [377], and the severity of P. falciparum malaria in African children [375].

1.3.

The Present Study:

39

This thesis is based primarily on results obtained in two studies which will be referred to as (Study I and II). The studies were carried out as a part of the severe malaria project in the Malaria Research Laboratory, Department of Biochemistry, Faculty of Medicine, University of Khartoum, Sudan, in collaboration with the Department of Immunology, Wenner-Gren Institute, Stockholm University, Sweden. Manuscripts of intended publications originating from these studies, together with reports of other relevant work to which the author contributed during the study period are annexed to the thesis.

1.4.

Rationale. Study I. Natural antibody-mediated immune responses are essential for acquisition of protective immunity against malaria infection and disease in endemic areas [378-380]. Passive transfer of malaria specific immunoglobulins has been shown to reduce parasitemia and to improve the clinical condition of malaria patients [381, 382]. Early studies have clearly demonstrated the predominance of immunoglobulin-G (IgG) in resolution of the blood stage infection, and several antibody-mediated mechanisms of parasite growth inhibition and killing have clearly been elucidated [130-132]. Identification of potential targets and mechanisms of antibody-mediated protective immunity is central for rational vaccine development [383, 384] Generally, natural IgG responses in human comprise four distinct antibody subclasses with varying structural and functional properties. Previous studies which mostly investigated antibody response to either individual or limited panels of malarial antigens, have somehow associated levels of certain subclasses of IgG antibodies with some

40

parameters of protection from malaria as reviewed by Garraud et al.[385]. However, only a few field studies have investigated the role of IgG subclasses in protection against SM. For example, low levels of serum IgG3 against crude extract of P. falciparum blood-stage are associated with poor prognosis of SM, and increased IgG3 is associated with reduced risk of malaria attack in Senegal [308, 386]. In a prospective case-control study of Kenyan children, the mean levels of IgG class and subclass antibodies against crude P. falciparum antigen were comparable between children who developed SM and those who remained healthy for two consecutive transmission seasons [387]. However, in the same study, high levels of IgG1 in relation to IgG2 and IgG4 antibodies were associated with protection, whereas high levels of IgG2 in relation to IgG1 and IgG3 were associated with risk of SM [387]. In contrast, levels of IgG subclasses to P. falciparum were found to be the same in protected and non protected subjects in Madagascar [388]. Interestingly, among IgG class and subclass antibodies directed against four vaccine candidate antigens; the apical membrane antigen-1 (AMA-1), the merozoite surface protein-119 (MSP-119), MSP-3, and the glutamate-rich protein (GLURP), only levels of total IgG, IgG3 and IgG4 antibodies against GLURP and IgG1 antibodies against AMA1 were associated with reduced risk of malaria incidence in Burkina Faso [389]. In addition to the intrinsic properties of antigens, the association between IgG antibody levels and protection from malaria may vary with the age and the genetic constitution of the host, as well as the parasite density and strain. Thus, in spite of the convincing evidence for an important role of the cytophilic IgG1 and IgG3 in parasite killing, the role of IgG antibody subclasses in protection from SM complications is not clearly understood. In particular, the relevance of the quantitative variation of these antibodies and protection

41

from SM in areas of seasonal parasite transmission remains to be clearly defined. Here, we undertook analyses of levels of circulating anti-malarial IgG subclasses directed against crude malaria antigen in order to determine whether particular IgG subclasses are associated with protection from SM, and to find out how such associations might be influenced by the host age, parasite density and multiplicity of infection (MOI).

Study II. Interleukin-10 (IL-10) is an important immuno-suppressive cytokine produced by several cell types including macrophages, B and T lymphocytes and dendritic cells. The main biological function of IL-10 seems to be the restriction of inflammatory responses and the regulation of proliferation and differentiation of immune cells [300, 390]. Interleukin-10 has repetitively been associated in various ways with severe complications of P. falciparum malaria. For example, animal models, although not perfectly reproducing the human disease, have provided evidence for an important role of IL-10 in the pathogenesis of CM [163, 251, 391]. IL-10, principally from splenic T cells, was shown to play a crucial role in protection of C57BL/6 mice against the neurological syndrome associated with P. berghei ANKA infection [310]. This has further been substantiated by the demonstration of IL-10s ability to inhibit IFN- production and to protect against CM [307], and by the observation of enhanced malaria disease of P. chabaudi chabaudi infection in IL-10 knocked-out mice [307, 392, 393]. Therefore, inadequate production or defective respond to IL-10 were suggested as possible pathogenic mechanisms of severe malaria [196, 198, 282, 304]. However, the role of circulating levels of IL-10 in protection from or pathogenesis of human malaria is not

42

well understood. Several reports have associated elevated circulating levels of IL-10 with acute malaria [210, 226, 394], severe malarial anemia (SMA) [212], and CM [210, 245]. Moreover, high IL-10 levels during the early phase of infection were associated with less effective clearance of parasites and hyperparasitemia [305, 309]. Conversely, other authors have associated low levels of IL-10 with CM in deeply comatose patients [282], as well as with SMA [196, 395]. Quantitative differences in IL-10 production may partially be attributed to heritable genetic factors[334, 396]. Several nucleotide polymorphisms (SNPs) identified in the IL10 gene promoter were associated with differences in rates of the cytokine production [348]. The relevance of these SNPs to susceptibility to and/or severity of inflammatory and infectious diseases has been demonstrated [373-375]. For example, the A/G transition at position -1082 (rs 1800896) has been associated with either protection against or enhanced severity of several inflammatory and infectious diseases including EpsteinBarr virus infection [373], inflammatory bowel disease and ulcerative colitis [397], Juvenile rheumatoid arthritis [358] and chronic hepatitis C virus infection [372]. Here, I investigated whether a particular allele or genotype at position -1082 is associated with protection from or susceptibility to SM malaria, and whether this is related to the circulating levels of IL-10 in different forms of clinical malaria among Sudanese children living in areas of unstable seasonal malaria transmission.

1.5.

The General Objectives:

43

The overall objective of these studies was to investigate the role of the anti-malaria immunoglobulin-G subclasses and interleukin-10 in the immune-pathogenesis of malaria.

1.5.1. The specific objectives 1.5.1.1. Study I.

1. To measure circulating levels of anti-malarial immunoglobulin-G (IgG) isotypes in order to examine how these levels relate to malaria protection and pathogenesis. 2. To determine the multiplicity of infection (MOI) by Plasmodium falciparum in patients with severe and uncomplicated malaria, and to assess the effect of MOI in the acquisition of the antibody mediated immunity to malaria.

1.5.1.2.

Study II. 1. To measure the circulating levels of IL-10 in order to determine if these levels correlate with a particular genetic allele, and how they affect the outcome of malaria infection. 2. To determine the genotypes and the allele frequencies of the single nucleotide polymorphism (SNP) at position -1082 [rs 1800896], in order to find out how they relate to protection from severe malaria.

44

Chapter Two

Material and Methods

45

2. 2.1.

Material and Methods:


The study area and population.

These studies were carried out over a period from September to December 2003, in ElGedarif Hospital and during the same period in 2006 in El-Obied Teaching Hospital. ElGedarif town (350 to 400 thousand inhabitants) is situated on the Sudanese Savanna, at an altitude of about 600 m above sea level, and is located 350 km southeast of the capital, Khartoum. The epidemiology of febrile uncomplicated malaria episodes in the area has been reported before [398]. Malaria in this area is highly seasonal, and follows the annual June-October rains; the seasonal incidence of malaria varies considerably from year to year [399]. Anopheles arabiensis is the sole malaria vector in the area. The entomological inoculation rate in Gedarif was found too low and difficult to measure by conventional methods in the surrounding villages [400]. El-Obied, the capital of the state of North Kordofan, is 400 km south-west of Khartoum. North Kordofan State lies in arid semi-desert area, and has a population of around 2.6 million, mainly sedentary farmers and nomad livestock breeders. The rainy season extends from June to mid-September with an annual rainfall of less than 400 millimetres, allowing grazing and seasonal rain-fed agriculture. The malaria epidemiology and clinical features in this area have not been investigated and little information is available about the disease prevalence and incidence. The patients attending the out-patient clinic of El-Gedarif Hospital and the Paediatric Unit of El-Obied Teaching Hospital during the mentioned periods were carefully screened for the signs of malaria. A patient was defined to have malaria if he/she complained of fever, had body temperature exceeding 37.5oC with asexual malaria 46

parasites detected in the blood. Patients who met the criteria for severe and uncomplicated malaria as described by the World Health Organisation (WHO) [24] were selected for study. Patients classified with cerebral malaria were in unarousable coma which persisted for at least 30 minutes after a seizure while patients with severe malarial anaemia had normocytic anemia with haematocrit <15% or haemoglobin <5 g/dl in the presence of microscopically detectable parasitaemia. Uncomplicated malaria cases were described as having a positive blood smear, fever, headache, and no signs of severe manifestation. All patients were promptly treated free of charge according to the hospital treatment protocols. The study was approved by the research committee of the Faculty of Medicine, University of Khartoum, and the National Research Directorate of the Sudan's Federal Ministry of Health. Informed consents were obtained from the patient and/or their guardian before enrollment. Unrestricted immediate access to health care was granted for all patients regardless of their acceptance or refusal of enrolment in the study. Comprehensive demographic and clinical data were recorded in the study questionnaire. All participants were subjected to a series of clinical investigations including haemoglobin (Hb%), blood glucose level, and a complete haemogram. Thin and thick blood smears were prepared to investigate parasitemia, species and density. Samples of venous blood were collected in EDTA tubes before commencing treatment. Approximately, 200 l of blood were blotted onto a Whattman-1 filter paper, air-dried and stored in sealed plastic bags for later use for extraction of DNA. The remaining blood samples were immediately centrifuged and the plasma aliquots were collected into 2 ml eppendorff tubes and stored at -80 C, the RBCs-containing pellets were

47

appropriately treated and cryopreserved for further parasite analyses.

2.2.

Microscopy for diagnosis and estimation of parasite density.

Preliminary diagnosis of malaria was made by quick screen of Giemsa-stained thick peripheral blood films. Standard techniques were used to confirm diagnosis and normocytemia as well as to estimate parasite density by two independent microscopists in duplicates of the thick and thin films. A slide was deemed negative after screening 100 consecutive oil immersion fields of a light microscope (Olympus, Japan; Binocular microscope Model- CH-2; using 100x objective and l0x eye pieces). The parasite density per microliter (L) of blood was calculated from the number of parasitized red blood cells (pRBC) per 200 white blood cells (WBC), assuming an average normal WBC count of 8000/L.

2.3.

Parasite DNA extraction, and estimation of P. falciparum multiplicity of infection (MOI).

Parasite DNA was extracted from filter-papers as described elsewhere [401] with slight modifications. Briefly, three small discs each of an area of 6 mm3 were cut off the bloodblotted filter paper and soaked in 1 ml 0.5 % Saponin in phosphate buffer saline (PBS) and incubated overnight at 4 C. The discs were subsequently washed for 30 minutes with PBS and boiled in 200 l of 5 % Chelex-100 in water for 15 minutes. Samples were vortex-mixed and centrifuged for 2 minutes at 10,000 rpm, the supernatants were collected and stored at -20 C until use for DNA genotyping. The multiplicity of infection is the approximate number of parasite variants specified by distinct alleles of the highly polymorphic MSP-2 gene [402]. Nested polymerase chain 48

reaction was used to amplify specific target sequences in MSP-2 as described elsewhere [403]. Briefly, in an Eppendorf thermal cycler (Mastercycler 5330), the PCR reactions were carried out in a primary amplification program that involved DNA denaturation at 95 C for 1 min, followed by 30 cycles of 98 C/10 sec, 61 C/20 sec/ 72 C/10 sec, and final extension for 5 min. The reaction mixture consisted of 5 l of iProofTM high fidelity master mix (BIO-RAD Laboratories, Hercules, CA), 2.5 l of the DNA extract, 0.5 M of primer pairs and 2 l of sterile distilled water. The outer primers corresponding to the conserved regions of block 3 of msp-2 are detailed elsewhere [404]. The second amplification round was similarly carried out, but with 1 l of the initial PCR products in two separate tubes using allele-specific primers corresponding to the two major MSP-2 families (IC/3D7 and FC27) [405]. The final PCR products were then subjected to electrophoresis on 1.5 % agarose gel loaded with ethidium bromide. Gels were illuminated with GelDoc-It TS 2UV transilluminator (BioDoc-ItTM imaging system, Upland, CA. USA) and photographed to determine number and size of DNA fragments presumed to represent different parasite variants. The least number of fragments of both alleles per sample was used as a proxy for MOI.

2.4.

Preparation of the crude parasite antigens and determination of P. falciparum -specific antibody levels.

Plasmodium falciparum parasites were maintained in continuous in vitro cultures in human erythrocytes (group O+) in RPMI 1640 + L-glutamine-NaHCO3 (GIBCO N.Y. USA) with 0.5 % albumax supplemented with Hepes, NaHCO3, gentamycin and hypoxanthine as described by Trager and Jensen [406]. The parasite cultures were

49

synchronised with 5 % D (-) sorbitol treatment in distilled water and adjusted to 1 % parasitemia of mostly early trophozoites (rings) in 5-6 % hematocrit. Sonicates of late stage infected erythrocytes, enriched by 60 % Percoll gradient centrifugation were used as antigens as previously described [407]. The protein content of the crude extract was estimated by Bradford method [408], and the extract was aliquoted into separate sterile screw capped tubes and stored at 80 C until use.
Malaria specific total IgG and the subclasses, IgG1, IgG2, IgG3, and IgG4 were measured

by an indirect ELISA, using crude antigen extract of the F32 isolate of P. falciparum. Briefly, flat-bottomed microtiter plates (High Binding from Costar, Cat. No. 3590 Corning, NY 14831, USA) were coated with 50 l/well of crude parasite antigen at 10 g/ml in sodium bicarbonate coating buffer and incubated at 4 C overnight. The plates were then blocked with 100 l/well of sodium carbonate buffer (pH 9.6) containing 0.5 % bovine serum albumin (BSA) for 2 h at 37 C. Thereafter, the plates were washed with PBS supplemented with Tween-20 plus 0.15 % Kathon (Mabtech, Nacka, Sweden) in a microplate washer (Skan washer 300, CA, USA). After washing, 50 l/well of serum samples diluted 1:400 (IgG1-4) in incubation buffer (PBS + 0.5 % BSA) were added in duplicates and incubated for 1 h at 37 C. Sera of different clinical groups were blindly allocated into plates to reduce inter-plate error. Plates were washed as above, and bound IgG subclasses were detected with the respective mouse anti-human IgG1 1:1000 (SkyBio, Bedfordshire, UK), IgG2 1:3000 (PharMingen, Erembodegem, Belgium), IgG3 1:1000 (Caltag Laboratories, Paisley, UK), and IgG4 1:2000 (Sigma, St. Louis, USA) and incubated for 1 h at 37 C. The plates were washed again and 50 l of conjugated streptavidinALP (Mabtech) diluted 1:2000 50

in incubation buffer were added and incubated for 1 h at 37 C. The plates were washed once more and developed by the addition of 50 l/well of p-nitrophenyl phosphate (Sigma-Aldrich Chemie, Steinheim, Germany). The optical densities of plates were read at OD405 with a VmaxTM microplate reader (Molecular Devices Corporation, Menlo Park, USA). Serial standard concentrations (ranging from 0.003 to 0.3 g/ml) of IgG subclasses from Biogenesis (U.K) were used for the calibrations, and plasma samples from previously known malaria positive subjects from the Fulani tribe in Mali were used as positive controls. Sera from a few malaria unexposed Swedish blood donors were used as negative controls to assess cross-reactivity and to determine the baseline levels. The mean levels of IgG antibodies detected in these negative control sera were used as baselines above which a test sample is considered positive for the particular antibody.

2.5.

Extraction of human DNA and genotype of interleukin-10 (-1082A/G) SNP:

Extraction of human DNA from filter-paper samples has been carried out by boiling as described above for parasite DNA extraction. The -1082 A/G genotype of IL-10 was determined by analyses of fragment length polymorphisms as described elsewhere [356]. Briefly, extracted DNA was amplified by PCR using sequence-specific primers (Forward; 5-AAG GCA ACA CTA CTA AGG CTT CCT T-3, and Reverse; 5-TAA ATA TCC TCA AAG TTC C-3). The PCR reaction mixture included 5 l of 2X PCR Master Mix (ABgene, Epsom, Surrey, UK), 0.25 l of each primer, 2 l distilled water and 3 l of genomic DNA extract. The reaction was carried out in Eppendorf Mastercycle (Eppendorf AG, Hamburg, Germany) with a program of initial denaturation at 95C/5min followed by 35 cycles of 95C/30 sec, 55C /30 sec, 72C/60 sec and final extension at 51

72C/5 min. The PCR product was then treated with the endonuclease EcoN1 (# R0521L BioLab, New England) according to manufacturer instructions, and incubated at 37 C for 2 hours. EcoN1 cleavage site is the underlined C in the forward primer above. The digested product was visualized on 1.5 % agarose gel with ethidium bromide, the length and number of fragments were used to identify genotypes; where the homozygous GG yields two fragments (280 and 310 bp), the heterozygous AG yields four fragments (310, 280, 252, and 28 bp) and the homozygous AA yields three (310, 252 and 28 bp) fragments.

2.6.

Determination of circulating levels of interleukin-10 .

Circulating levels of IL-10 were determined by direct ELISA. Flat-bottomed microtiter plates (High Binding from Costar, Cat. No. 3590 Corning, NY 14831, USA) were coated with 50 l/well of anti-human IL-10 monoclonal antibody 9D7 (Mabtech AB, Nacka, Sweden) diluted to 2 g/ml in phosphate buffered saline (PBS) (pH of 7.4), and incubated overnight at 4 C. The plates were then washed twice with PBS and blocked with 100 l/well of 0.1 % bovine serum albumin (BSA) in PBS for two hours at room temperature. The plates were then washed with PBS with Tween-20 plus 0.15 % Kathon in a microplate washer (Skan washer 300, Skatron Instruments, Norway). Hundred l/well of sera samples [diluted 1:100 in 0.15 M Tris-Hanks] were added to 20 l/well of Mabtech ELISA diluent and incubated at room temperature for 2 hours and then washed again as above. Then 50 l/well of biotin-conjugated monoclonal antibody ( Mabtech 12G8 [diluted 1:1000 in Mabtech ELISA diluent]) was added and incubated for 1 h at room temperature. The plates were then washed as above followed by addition of 50 l of 52

StreptavidinALP from Mabtech diluted 1:1000 in incubation buffer with 1% BSA and incubated for 1 h at room temperature. The plates were washed, developed and read using Molecular Devices (VmaxTM Kinetic microplate reader Menlo Park, USA) at 405 nm. Serial standard concentrations of (10 to 3.160 pg/ml were used for calibration, the curve fit was log-log = {log10(Y) = A + B * log10(X)}; where A = -01.59 is the log10 Y-intercept of the line of log10 (X) = 0, and B = 0.592 is the line slope. The correlation coefficient was 0.999.

2.7.

Statistical analyses.

The data were analyzed with SPSS software version 16.0.1 (SPSS Inc., Chicago, IL, USA). Logistic regression analyses were used to assess differences in frequencies and proportions and to estimate malaria risk. Scale variables which are not normally distributed were log10 transformed, and differences of means were investigated by t-test and one way analysis of variance. Fishers exact test for contingency tables was used to test for significance in proportions of categorical data between the clinical malaria groups when the expected value was less than 5.

53

Chapter Three

The results

3. The results:

3.1.

Study I. 54

3.1.1. The demographic and clinical features of the study population. The total number of individuals enrolled from El-Gedarif was 177; of them 37 were patients with cerebral malaria (CM), 44 presented with severe malarial anemia (SMA), 55 were patients with UM (UM) and 41 were malaria free donors (MF). The demographic and clinical features and parasitological indices of the study population are given in Table 3.1. Six patients with SMA were excluded from the current analysis because of

insufficient plasma samples or incomplete set of data. Four of those who had CM were mature adults (age range 25-55 years), while all those who had SMA were young children mostly under five years. Individuals presenting with SMA had lower mean age (P = 0.002) as compared to those who had CM. Three patients (two children and one adult) representing 8% of those who presented with CM died within 24 hour after admission into hospital. Whereas parasite densities were comparable between CM and SMA, both groups had significantly higher parasitemia (P = 0.009 and 0.017 respectively) than UM. Neither the MOI nor an individual msp-2 genotype was found significantly different between SM and UM, or between the CM and SMA.

55

Table 3.1.

The demographic and malariological indices of the study groups. Study Groups Cerebral Malaria 37 (14/23) 5.0 (4.0 7.8) 4.4 (0.58) * 1.7 (0.2) 2.1 (0.2) 2.0 (1.2) * Severe Malarial Anemia 38 (12/26) 2.75 (1.5 4.0) * 4.4 (0.48) 1.4 (0.2) 1.8 (0.2) 1.5 (0.9) Uncomplicated Malaria 55 (25/30) 5.0 (3.0 7.0) 4.1 (0.54) 1.6 (0.2) 2.2 (0.2) 2.2 (1.2) Malaria Free Donors 41 (12/29) 5.0 (2.0 - 6.8) -----

Number (female/male) Age years (median, 25 and 75 percentiles) Log10 parasite count. Number of IC/3D7 alleles Number of FC27 alleles Multiplicity of infection (MOI)

Except for age, all values are given as mean (SD), and * significant values are discussed in the text.

56

3.1.2. Correlation of IgG antibodies with age, parasite density, MOI and msp-2 genotypes in the study population. Pearsons correlation analysis showed significant positive association between age and total IgG (r = 0.266, P = 0.001), IgG1 (r = 0.299, P < 0.001), IgG2 (r = 0.459, P < 0.001) and IgG3 (r = 0.350, P < 0.001) but not with IgG4 (r = 0.132, P = 0.089) in all study groups. In an age adjusted analysis, the parasite density correlated significantly inversely with total IgG (r = -0.260, P = 0.006), IgG1 (r = -0.287, P < 0.002), IgG2 (r = -0.204, P = 0.033) and IgG3 (r = -0.205, P = 0.031), but did not correlate with IgG4 (r = -0.06, P = 0.53). In contrast, no significant correlation could be found between any Ab levels and neither MOI nor the number of individual msp-2 genotypes (IC-1 or FC27). The correlation between the levels of individual IgG subclasses and the total IgG concentrations was significant and supports the accuracy of the results (not listed).

3.1.3. Comparison of IgG antibody subclass levels between study groups. The overall comparison of levels of total and subclass IgG antibodies as shown in figures 3.1 and 3.2 revealed higher levels of IgG1 and IgG3 subclasses in the whole study population regardless of the clinical group. The levels of these two antibody subclasses were highly variable both between and within study groups. The pair-wise comparisons of mean antibody levels between the study groups revealed higher levels of total IgG (P < 0.001), IgG1 (P = 0.001) and IgG3 (P < 0.001) but not IgG2 and IgG4 in all malaria patients (SM and UM) compared to the MF individuals. Further age-adjusted analyses splitting malaria patients into SM and UM revealed that there were lower levels of total IgG (P = 0.001), IgG1 (P = 0.004) and IgG4 (P = 0.023), but not IgG2 or IgG3 in the SM compared to UM. Re-analysis of data splitting the SM group into CM and SMA showed comparable levels of total 57

IgG (P = 0.15), IgG2 (P = 0.17), IgG3 (P = 0.30) and IgG4 (P = 0.43) but not of the IgG1 (P = 0.034) subclass in SMA and CM. However, the difference in IgG1 levels between these groups disappeared after adjustment for age (P = 0.33). The SMA patients had lower levels of total IgG, (P = 0.002), IgG1 (P = 0.004) and IgG4 (P = 0.031) compared to those with UM, whereas the levels of all IgG subclasses did not differ significantly between CM and UM. Noteworthy, the differences between the groups of SMA and UM remained significant in the levels of total IgG (P = 0.009), IgG1 (P = 0.03) and IgG4 (P = 0.019) after adjustment for both age and parasite density. Association of risk of SM and levels of total IgG, IgG1 and IgG4 was carried out by calculating the odds ratio of the prevalence of negative versus positive sera for the specific antibody among the SM compared to UM as shown in Table 3.2. There was significantly lower prevalence of IgG1 (P = 0.01) and IgG4 (P = 0.002) reactive sera among the group of SM compared to UM. However, there were no significant differences in the prevalence of total IgG, IgG2 and IgG3 between these two groups.

58

Figure 3.1.

General comparisons of IgG total and subclass levels within and between

different study groups of severe malaria (SM), uncomplicated malaria (UM) and malaria free donors (MF).

59

Figure 3.2.

Comparisons of IgG total and subclass levels within and between groups of

cerebral malaria (CM), severe malarial anemia (SMA) and uncomplicated malaria (UM).

60

Table 3.2. Antibody IgG1

Antimalarial immunoglobulin-G subclass sera reactivity and risk of severe malaria. Reactivity* Severe Uncomplicated Total OR 95 % CI Negative Positive malaria 23 (31.1 %) 51 (68.9 %) 27 (36.5 %) 47 (63.5 %) 1 (1.4 %) 72 (98.6 %) 46 (63 %) 27 (37 %) 1 (1.4 %) 71 (98.6 %) malaria 6 (10.9 %) 49 (89.1 %) 16 (11.1 %) 38 (88.9 %) 3 (5.5 %) 52 (94.5 %) 19 (35.2 %) 35 (64.8 %) 0 (0.0%) 51 (100 %) 29 (22.5 %) 100 (77.5 %) 43 (33.6 %) 85 (66.4 %) 4 (3.1 %) 124 (96.9 %) 65 (51.2 %) 62 (48.8 %) 1 (0.8 %) 122 (99.2 %) 3.68 0.27 1.36 0.73 0.24 4.15 3.14 0.32 1.73 0.58 1.38 9.54 0.11 0.71 0.65 2.87 0.35 1.55 0.03 1.75 0.57 29.66 1.51 6.51 0.15 0.66 0.15 19.79 0.05 6.68

P value 0.010

IgG2

Negative Positive

0.453

IgG3

Negative Positive

0.314

IgG4

Negative Positive

0.002

Total IgG

Negative# Positive

0.418

* Antibody level greater than the cut-off point established as the average level measured in sera from malaria unexposed donors was considered positive. OR = Odds ratios and CI represents Confidence Interval. # Continuity correction was made to eliminate the zero count of negative IgG total sera among persons with uncomplicated malaria, (1/(R+1) and (R/(R+1) were respectively added to counts of SM and UM in total IgG, where R = the total number of SM/ total number of UM = 72/51.

61

3.2.

Study II.

3.2.1. The demographic and malariological indices of the study groups. The total number of individuals investigated in this study was 352 including 171 individuals from the samples collected in El-Gaderif during season 2003 and the rest were from El-Obied and the villages around the city during 2006. Those who were enrolled from El-Obied included 4 patients with CM, 23 with SMA, 82 with UM, and 72 were MF. The sex frequency, age and clinical classification of study II population are shown in Table 3.3. For the overall study population, there are no significant differences in age and sex frequencies between the different clinical groups.

3.2.2. The circulating levels of interleukin-10 in the different study groups. The distribution of the levels of IL-10 was positively skewed, thus scores of IL-10 concentration were log transformed to improve normality. One way analysis of variance (ANOVA) revealed significant differences in the circulating IL-10 between the different clinical groups (F = 7.1, P < 0.001) (Figure 3.3). Tukeys post hoc analysis, showed significantly lower concentrations of IL-10 in CM patients as compared with the MF donors (P = 0.001) or with UM patients (P < 0.001). Although not significant, patients with SMA have had lower IL-10 levels compared to those with UM or the MF donors separately, however, SMA patients had slightly higher levels of IL-10 than patients with CM. Interestingly, there was an inverse correlation between IL-10 levels and the degree of fever indicated by the body temperature (1-tailed P = 0.009) (data not shown). No correlation was observed between IL-10 levels and the patient age or parasitemia separately.

62

2.1.1. The frequencies of alleles and genotypes in interleukin-10 (-1082 A/G) single nucleotide polymorphism. Of the total number of samples compiled from the two sites, we could extract sufficient DNA and genotyped the SNP (-1082 A/G) for 226 samples. The Hardy-Weinberg distribution of alleles and genotypes among different clinical groups and the estimation of risk of severe malaria incurred by carriage of a specific allele or genotype of the -1082 SNP are given in Tables 3.4 and 3.5. The allele distribution among the patients of CM but not SMA or UM deviated significantly from Hardy-Weinberg Equilibrium (HWE). The possible gene-phenotype interactions and outcomes of either allelic or genotypic constitution on malaria outcome are given in Table 3.5. Because it is inapplicable in this situation, we excluded the multiplicative genetic model assuming a risk weight (r) for the heterozygous and (r2) for homozygous candidate allele [409]. The alternative additive model assuming a risk weight (r) for the heterozygous and (2r) for homozygous, were used to test the hypothesis that carrying the G allele influences the risk of CM. Further analysis of the latest model assuming dominant or recessive modes of the allele inheritance was performed by pooling the genotypes into different models (Table 3.5).

63

Table 3.3.
Clinical groups Number (Female + Male). Age in years, mean (SD). (CM)

Demographic features of study II population


(SMA) 61 (22 + 39) 3.28 (2.63) (UM) 137 (56 + 81 M) 4.40 (2.89) (MF) 113 (47 + 66) 4.27 (2.40) Total 352 (141 + 211 ) 4.21 (2.75)

41 (16 + 25) 4.86 (2.41)

Table 3.4. The Hardy-Weinberg Equilibrium for the interleukin-10 (-1082 A/G) SNP in the study groups . Patients groups IL-10 genotype Allele frequency (A:G) %
AA AG GG Total

P values *.

64

Cerebral Malaria . Severe Malarial Anemia . Uncomplicated Malaria. Malaria free donors. Total study population

15 18 42 07 82

8 20 54 08 90

12 14 25 03 54

35 52 121 18 226

54.3 : 45.7 53.8 : 46.2 56.2 : 43.8 61.1 : 38.9 56.2 : 43.8

0.006. 0.264 0.616 0.961 0.016

* Chi Square test with 2 df, P value less than 0.05 indicates inconsistency with Hardy-Weinberg Equilibrium.

65

Table 3.5.
Allele/Genotype 1. Co-dominant inheritance A G Total 2.

Estimation of odds ratios, assuming different modes of A/G alleles inheritance.


SM (n) 83 71 154 30 47 77 53 24 UM (n) 137 99 236 47 71 118 90 28 Total 220 170 390 77 118 195 143 52 OR (95.0% C.I.) 1.18 (0.79 to 1.78) 0.84 (0.56 to 1.27) 1.04 (0.58 to 1.87) 0.96 (0.54 to 1.74) 4.67 (2.69 to 8.09) 0.69 (0.36 to 1.31) 0.32 0.90 P value * 0.42

Dominant G allele inheritance. AA AG + GG Total

3. Recessive G allele inheritance. AA +AG GG

Total 77 118 195 * Pearson Chi-Square test comparing the frequencies among patients with severe malaria (SM) and uncomplicated malaria (UM) revealed no significant risk SM being associated with any of the alleles or genotypes of -1082 A/G single nucleotide polymorphism in IL-10.

66

12

10

Serum levels of IL-10 (pg/ml)

CM

SMA

UM

MF

Patients groups

Figure 3.3

Levels of circulating IL10 in different clinical groups of malaria patients and

healthy controls, box and error bars show the median and 95% CI of circulating IL-10 levels in the cerebral malaria (CM), severe malaria anemia (SMA), uncomplicated malaria (UM) and malaria free donors (MF).

67

Chapter Four

Discussion and conclusions

68

4.1 Discussion. 4.1.1. Study I. In study I, as previously noted in responses to many malaria antigens investigated elsewhere [410-412], the relative abundance of IgG subclasses was profoundly altered, being largely dominated by IgG1 and IgG3. There were higher levels of IgG1 and IgG3 compared to IgG2 and IgG4 in all individuals investigated regardless of the clinical status. Except for IgG2, there were prominent increases in levels of all IgG antibodies in infected individuals compared to the MF donors. The rise in levels of IgG antibodies in the infected individuals is obvious as current parasitemia induces the antibody production. However, the similar levels of IgG2 in the patients and the malaria free individuals cannot be explained as a response to the current infection, rather it may indicate the intensity of P. falciparum transmission during the specified period. Worth mentioning; we collected the samples towards the end of the rainy season after a notably intense transmission season. In spite of the altered natural balance of IgG subclasses serum levels, and apart from the notable differences between the study groups, we could still detect a clear age-dependent increase in the levels of all IgG subclasses except IgG4. This age associated elevation of IgG levels concurs with the well known age-dependent evolution of protection from malaria [306, 378]. However, the lack of association between IgG4 and age in this analysis might partly be explained by the relatively young age of the children investigated hence the limited exposure to parasite. Interestingly, the IgG4 antibodies which are known to be induced by chronic infections [413, 414], might reflect the development of immune memory as a result of cumulative exposure to different parasite strains including the relatively common variants presumed to cause severe malaria. Accordingly, the elevated levels of IgG4 detected in infected individuals as compared

69

to the MF donors, are clearly associated with the current infection. However, this response might have been induced by Th-2 polarized T-cell responses. This is supported by reports demonstrating the induction of IgG4 by interleukin-4 (IL-4) [415, 416]. The effect of parasite density on the levels of anti-malaria IgG antibodies might not be straight forward due to variation in the parasite antigenic characteristics and sequestration in micro capillaries. However, in this study, we observed significant negative associations between the parasite density and the levels of all IgG antibodies except IgG4. All of these associations were age-independent, with IgG1 exhibiting the stronger association as compared to IgG2 and IgG3. The inverse associations of IgG1 and IgG3 with the parasite density are in line with the well established anti-parasite effects of these anti-malarial antibodies. However, the association between IgG2 and parasite densities reveals an unusual anti-parasite effect for this subclass. This is interesting, because unlike the primarily cytophilic IgG1 and IgG3, IgG2 antibodies do not bind the common Fc receptor (FcR) types that mediate the antibody-dependent cellular cytotoxicity. Thus, the observed anti-parasite effect of IgG2 may be mediated through a mutated type of FcRII which was shown to be common in this population [417]. The specific role of IgG4 antibodies in malaria protection and/or pathogenesis is not well understood. Only a weak non significant negative association of IgG4 with parasite density was observed, indicating little if any anti-parasite activity in this subclass. However, the increased risk of SM associated with the low prevalence of IgG1 and IgG4 antibodies in this study suggests a protective role for these two subclasses against SM. Pair-wise comparisons of the mean levels of IgG antibodies between the groups of SM, UM, and MF, and subsequently after splitting the group of SM into CM and SMA provided further support for different levels of IgG1 and IgG4 in SM and UM. While levels of all IgG antibodies are comparable between CM and SA, the combined SM group had significantly lower levels of total IgG, IgG1 and IgG4 than 70

those with UM. Although all these differences persisted after adjustment for age, only the difference in total IgG and IgG4 remained significant after inclusion of both age and the parasite density as covariates indicating a marked effect of parasite density on production of IgG1 but not total IgG and IgG4. Taken together, the associations of the low prevalence and levels of IgG1 and IgG4 antibodies with SM indicate important protective effect for these subclasses against SM, and whereas the protective effect of IgG1 may largely be due to its known antiparasite action, the presumed protective role of IgG4 against SM may be attributed to the clearly evidenced anti-inflammatory effect of IgG antibodies [418, 419]. Although the antiinflammatory role of IgG4 is proved in its clinical use for the treatment of autoimmune diseases, it remains to be demonstrated whether antimalarial IgG4 undergo Fab arm exchange that provides the basis for the anti-inflammatory activity attributed to these antibodies [419]. Several lines of evidence argue for a possible role for this anti-inflammatory activity of IgG4 in protection from severe complications of malaria. For instance, the protection from SM conferred by infection with helminthes [420, 421]. The increased levels of IgG4 associated with helminthes infection might be the underlying cause of the observed protection from clinical malaria. Moreover, IgG4 antibodies are poorly recognized by Fc receptors, thus , do not stimulate inflammation or activate the complement system, rather they were shown to inhibit the complement activation [422]. The mechanisms by which IgG4 mediate anti-severe malaria effects is not understood. However, unlike other antibodies IgG4 exist in vivo as monovalent, bispecific molecules, and are also capable of Fc-Fc interaction with other IgG antibodies [423]. Accordingly, we speculate that, IgG4 may directly block the binding of the cytophilic-antibodycontaining immune complexes to the effector cells. Alternatively, IgG4 may cause the formation of small, non precipitating immune complexes that do not induce inflammatory responses. In conclusion, the results of study I are not only revealing significant difference in IgG1 and 71

IgG4 levels and prevalence between SM and UM but also indicate an important anti-severe malaria protective effect. Thus, it is obvious that factors other than age and parasite density are involved in determining the serum levels of the anti-malaria IgG2 and IgG4 antibodies. In addition to the intensity of P. falciparum transmission other factors such as the number of previous clinical episodes, the genetic background of the host and the parasite strains might also influence the mechanisms regulating serum levels of these antibody subclasses.

4.2. Study II. In humans, the role of IL-10 in malaria is less well defined. It seems to be down-regulating the immunopathology underlying complicated malaria. High circulating levels of IL-10 have repeatedly been associated with SM in Thailand [226], Guinea [281], Mali [245] and Senegal [308]. Interestingly, in contrast to these, our current data showed low levels of circulating IL-10 in both groups of severe malaria (CM and SAM) as compared to those with UM or to HCs. The decrease in IL-10 levels was particularly prominent in the patients of CM. This observation is so far consistent with the suggested protective role of IL-10 previously demonstrated in mice models [307, 392, 424] and accords with a similar finding reported in deeply comatose adult Vietnamese patients with CM [282]. However, it contradicts many previous reports of elevated IL-10 during CM. This contradiction can be justified on the basis of different methodology and inclusion criteria of participants in these studies as well as variation in pattern of malaria transmission in the different epidemiological settings of these studies. Moreover, the immune competency, and human genetic variation in the studied populations might have also contributed to these contradicting results. According to these data, it could be speculated that CM might have arisen as a result of

72

uncontrolled pro-inflammatory responses, and that CM patients would have defective IL-10 production. The link between the circulating IL-10 levels and SMA is not well understood. Previous studies have linked SMA with relatively low levels of IL-10 in African children [196, 198, 425]. However, this is contradicted by other reports of high IL-10 during SMA [212]. Since many previous reports have associated SMA with increased levels of proinflammatory cytokines such as TNF- and IFN- which were suggested to enhance immune mediated destruction of both infected and non-infected RBCs and to suppress the erythropoiesis, it is likely that elevated levels of IL-10 through down-regulating inflammatory processes play a protective rather than deleterious role in SMA. However, this speculation does not go with the recent finding that high IL-10 is associated with inefficient clearance of P. falciparum which could as well lead to persistent parasitemia and severe malaria [305, 309]. Yet, it can still be argued that the role of IL-10 in the immuno-pathogenesis of SMA could be of dual nature, and that the temporal and balanced production of IL-10 rather than the absolute levels of this cytokine are the crucial factors in determining the outcome of malaria infection. Investigation of how the allelic variants and genotypes of the -1082 A/G SNP of IL-10 relate to the clinical forms of malaria revealed no significant association between the risk of severe malaria and any allele or genotypes of this specific SNP (Tables 3.4 and 3.5). Interestingly, tests for the Hardy-Weinberg Equilibrium (HWE) revealed significant deviation from the expected frequencies among the patients with CM (P<0.006) while the distribution of geneotypes was conforming with HWE among the other patients who presented with SMA or UM as well as the few MF donors who were analyzed. Logistic regression analysis of the allele frequencies within clinical groups (Table 3.5) revealed no significant association between any of the alleles and a particular form of clinical 73

presentations, but a non-significant trend of decreased homozygous GG genotypes can be observed among the all patient groups. However, due to lack of information about the frequency of these specific alleles and genotypes among the Sudanese population, it is not possible to draw conclusion about how this SNP relates to malaria pathogenesis. Taking into account the previous reports of higher productivity of IL-10 by the homozygous GG [347, 358, 396, 426, 427], it was interesting to assess how the G allele relates to the circulating levels of IL-10 in the different clinical groups. Comparison of mean levels of IL-10 between the different genotypes across different clinical groups revealed no significant differences between the carriers of either allele (data not shown). However, the homozygous GG individuals had slightly higher levels of IL-10 among the patients of UM, but surprisingly, lower levels in patients of CM. In general it could also be suggested that genetic background and haplotypic patterns rather than individual SNPs may be the cause of variable IL-10 production and eventually protection from and/or susceptibility to severe malarial diseases in different humans [375].

74

4.3. Conclusions. The present data reveal significant differences in the levels of IgG1 and IgG4 among between patients who presented with severe malaria as compared to those with uncomplicated malaria, indicating a protective effect for these antibody subclasess against severe malaria.

The multiplicity of infection (MOI) defined as the number of different Plasmodium falciparum clones showed no significant effect on the circulating levels of IgG antibodies, and the number of parasite clones were similar in the different patient groups.

The IL-10 levels in children with CM were significantly lower than in those with uncomplicated malaria, and IL-10 was particularly lower in comatose patients as compared to other forms of clinical malaria, indicating a protective role for IL-10 against severe malaria.

There was no significant association between the different alleles in the IL-10 gene SNP at position -1082 and the various clinical forms of malaria, and no specific allele or genotype was associated with serum levels of IL-10.

4.4. Perspectives. Future plans include investigating: Genetic factors determing susceptibility/resistance to severe malaria, particularly severe malarial anemia among sympatric tribes (Fulani, Nuba and Hawzma) living in Kordofan region in the middle west of Sudan.

75

References. 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. Breman JG, Alilio MS, Mills A: Conquering the intolerable burden of malaria: what's new, what's needed: a summary. Am J Trop Med Hyg 2004, 71(2 Suppl):1-15. WHO: The World Malaria Report 2005: Roll Back Malaria. In: The World Malaria Report 2005: . vol. Accessed on 4 Dec 2007. Geneva: World Health Organization; 2005 Hay SI, Snow RW: The Malaria Atlas Project: Developing Global Maps of Malaria Risk. PLoS Med 2006, 3(12):e473. Guyatt HL, Snow RW: Impact of malaria during pregnancy on low birth weight in subSaharan Africa. Clin Microbiol Rev 2004, 17(4):760-769, table of contents. Abdalla SI, Malik EM, Ali KM: The burden of malaria in Sudan: incidence, mortality and disability--adjusted life--years. Malar J 2007, 6:97. Craig MH, Snow RW, le Sueur D: A climate-based distribution model of malaria transmission in sub-Saharan Africa. Parasitol Today 1999, 15(3):105-111. Snow RW, Guerra CA, Noor AM, Myint HY, Hay SI: The global distribution of clinical episodes of Plasmodium falciparum malaria. Nature 2005, 434(7030):214-217. Hay SI, Tatem AJ, Graham AJ, Goetz SJ, Rogers DJ: Global environmental data for mapping infectious disease distribution. Adv Parasitol 2006, 62:37-77. Cox J, Hay SI, Abeku TA, Checchi F, Snow RW: The uncertain burden of Plasmodium falciparum epidemics in Africa. Trends Parasitol 2007, 23(4):142-148. Sachs J, Malaney P: The economic and social burden of malaria. Nature 2002, 415(6872):680-685. Gallup JL, Sachs JD: The economic burden of malaria. Am J Trop Med Hyg 2001, 64(1-2 Suppl):85-96. Russell S: The economic burden of illness for households in developing countries: a review of studies focusing on malaria, tuberculosis, and human immunodeficiency virus/acquired immunodeficiency syndrome. Am J Trop Med Hyg 2004, 71(2 Suppl):147-155. Chima RI, Goodman CA, Mills A: The economic impact of malaria in Africa: a critical review of the evidence. Health Policy 2003, 63(1):17-36. Molineaux L: Plasmodium falciparum malaria: some epidemiological implications of parasite and host diversity. Ann Trop Med Parasitol 1996, 90(4):379-393. Smith DL, Dushoff J, Snow RW, Hay SI: The entomological inoculation rate and Plasmodium falciparum infection in African children. Nature 2005, 438(7067):492-495. Drakeley C, Schellenberg D, Kihonda J, Sousa CA, Arez AP, Lopes D, Lines J, Mshinda H, Lengeler C, Armstrong Schellenberg J et al: An estimation of the entomological inoculation rate for Ifakara: a semi-urban area in a region of intense malaria transmission in Tanzania. Trop Med Int Health 2003, 8(9):767-774. Elissa N, Migot-Nabias F, Luty A, Renaut A, Toure F, Vaillant M, Lawoko M, Yangari P, Mayombo J, Lekoulou F et al: Relationship between entomological inoculation rate, Plasmodium falciparum prevalence rate, and incidence of malaria attack in rural Gabon. Acta Trop 2003, 85(3):355-361. Sagara I, Sangare D, Dolo G, Guindo A, Sissoko M, Sogoba M, Niambele MB, Yalcoue D, Kaslow DC, Dicko A et al: A high malaria reinfection rate in children and young adults living under a low entomological inoculation rate in a periurban area of Bamako, Mali. Am J Trop Med Hyg 2002, 66(3):310-313. Metselaar D: Spleens and holoendemic malaria in West New Guinea. Bull World Health Organ 1956, 15(3-5):635-649. Guerra CA, Snow RW, Hay SI: Mapping the global extent of malaria in 2005. Trends Parasitol 2006, 22(8):353-358. Chavatte JM, Chiron F, Chabaud A, Landau I: [Probable speciations by "host-vector 'fidelity'": 14 species of Plasmodium from magpies]. Parasite 2007, 14(1):21-37.

17.

18.

19. 20. 21.

76

22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34.

35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45.

Mendis K, Sina BJ, Marchesini P, Carter R: The neglected burden of Plasmodium vivax malaria. Am J Trop Med Hyg 2001, 64(1-2 Suppl):97-106. Carter R, Mendis KN: Evolutionary and historical aspects of the burden of malaria. Clin Microbiol Rev 2002, 15(4):564-594. Day. KP: The Epidemiology of Malaria, In: Malaria; Molecular and Clinical Aspects. Amsredam: Harwood Academic Publisher; 1999. Kiszewski A, Mellinger A, Spielman A, Malaney P, Sachs SE, Sachs J: A global index representing the stability of malaria transmission. Am J Trop Med Hyg 2004, 70(5):486-498. Kafatos. LZaFC: The Anopheles Mosquito: Genomic and Transmission. In: Malaria; Molecular and Clinical Aspects. Amsterdam: Harwood Academic Publishers; 1999. Garcia JE, Puentes A, Patarroyo ME: Developmental biology of sporozoite-host interactions in Plasmodium falciparum malaria: implications for vaccine design. Clin Microbiol Rev 2006, 19(4):686-707. Talman AM, Domarle O, McKenzie FE, Ariey F, Robert V: Gametocytogenesis: the puberty of Plasmodium falciparum. Malar J 2004, 3:24. Cowman AF, Crabb BS: Invasion of red blood cells by malaria parasites. Cell 2006, 124(4):755-766. Murphy GS, Oldfield EC, 3rd: Falciparum malaria. Infect Dis Clin North Am 1996, 10(4):747775. Marsh K: Clinical Features of Malaria; In Malaria Molecular and Clinical Aspects. Amsterdam: Harwood Academic Publishers; 1999. Njuguna P, Newton C: Management of severe falciparum malaria. J Postgrad Med 2004, 50(1):45-50. Clark IA, Alleva LM, Mills AC, Cowden WB: Pathogenesis of malaria and clinically similar conditions. Clin Microbiol Rev 2004, 17(3):509-539, table of contents. Bruneel F, Hocqueloux L, Alberti C, Wolff M, Chevret S, Bedos JP, Durand R, Le Bras J, Regnier B, Vachon F: The clinical spectrum of severe imported falciparum malaria in the intensive care unit: report of 188 cases in adults. Am J Respir Crit Care Med 2003, 167(5):684-689. Idro R, Jenkins NE, Newton CR: Pathogenesis, clinical features, and neurological outcome of cerebral malaria. Lancet Neurol 2005, 4(12):827-840. Mackintosh CL, Beeson JG, Marsh K: Clinical features and pathogenesis of severe malaria. Trends Parasitol 2004, 20(12):597-603. Warrell DA, White NJ, Warrell MJ: Dexamethasone deleterious in cerebral malaria. Br Med J (Clin Res Ed) 1982, 285(6355):1652. Warrell DA: Pathophysiology of severe falciparum malaria in man. Parasitology 1987, 94 Suppl:S53-76. Warrell DA: Pathophysiology of severe falciparum malaria. P N G Med J 1992, 35(4):229232. Warrell DA: Cerebral malaria. Schweiz Med Wochenschr 1992, 122(23):879-886. Newton CR, Pasvol G, Winstanley PA, Warrell DA: Cerebral malaria: what is unarousable coma? Lancet 1990, 335(8687):472. Bloland PB, Lackritz EM, Kazembe PN, Were JB, Steketee R, Campbell CC: Beyond chloroquine: implications of drug resistance for evaluating malaria therapy efficacy and treatment policy in Africa. J Infect Dis 1993, 167(4):932-937. Zucker JR, Lackritz EM, Ruebush TK, 2nd, Hightower AW, Adungosi JE, Were JB, Metchock B, Patrick E, Campbell CC: Childhood mortality during and after hospitalization in western Kenya: effect of malaria treatment regimens. Am J Trop Med Hyg 1996, 55(6):655-660. Ekvall H: Malaria and anemia. Curr Opin Hematol 2003, 10(2):108-114. Crawley J: Reducing the burden of anemia in infants and young children in malariaendemic countries of Africa: from evidence to action. Am J Trop Med Hyg 2004, 71(2

77

46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62.

63. 64. 65.

Suppl):25-34. Lozoff B, Jimenez E, Hagen J, Mollen E, Wolf AW: Poorer behavioral and developmental outcome more than 10 years after treatment for iron deficiency in infancy. Pediatrics 2000, 105(4):E51. Grantham-McGregor S, Ani C: A review of studies on the effect of iron deficiency on cognitive development in children. J Nutr 2001, 131(2S-2):649S-666S; discussion 666S-668S. Sherriff A, Emond A, Bell JC, Golding J: Should infants be screened for anaemia? A prospective study investigating the relation between haemoglobin at 8, 12, and 18 months and development at 18 months. Arch Dis Child 2001, 84(6):480-485. Lawless JW, Latham MC, Stephenson LS, Kinoti SN, Pertet AM: Iron supplementation improves appetite and growth in anemic Kenyan primary school children. J Nutr 1994, 124(5):645-654. Oppenheimer SJ: Iron and its relation to immunity and infectious disease. J Nutr 2001, 131(2S-2):616S-633S; discussion 633S-635S. Haas JD, Brownlie Tt: Iron deficiency and reduced work capacity: a critical review of the research to determine a causal relationship. J Nutr 2001, 131(2S-2):676S-688S; discussion 688S-690S. Greenwood B, Marsh K, Snow R: Why do some African children develop severe malaria? Parasitol Today 1991, 7(10):277-281. Baird JK: Age-dependent characteristics of protection v. susceptibility to Plasmodium falciparum. Ann Trop Med Parasitol 1998, 92(4):367-390. Wipasa J, Elliott S, Xu H, Good MF: Immunity to asexual blood stage malaria and vaccine approaches. Immunol Cell Biol 2002, 80(5):401-414. Snow RW, Craig M, Deichmann U, Marsh K: Estimating mortality, morbidity and disability due to malaria among Africa's non-pregnant population. Bull World Health Organ 1999, 77(8):624-640. Andrysiak PM, Collins WE, Campbell GH: Stage-specific and species-specific antigens of Plasmodium vivax and Plasmodium ovale defined by monoclonal antibodies. Infect Immun 1986, 54(3):609-612. Fandeur T, Chalvet W: Variant- and strain-specific immunity in Saimiri infected with Plasmodium falciparum. Am J Trop Med Hyg 1998, 58(2):225-231. Rowe JA, Moulds JM, Newbold CI, Miller LH: P. falciparum rosetting mediated by a parasite-variant erythrocyte membrane protein and complement-receptor 1. Nature 1997, 388(6639):292-295. Pied S, Nussler A, Pontent M, Miltgen F, Matile H, Lambert PH, Mazier D: C-reactive protein protects against preerythrocytic stages of malaria. Infect Immun 1989, 57(1):278-282. McGuire W, D'Alessandro U, Olaleye BO, Thomson MC, Langerock P, Greenwood BM, Kwiatkowski D: C-reactive protein and haptoglobin in the evaluation of a community-based malaria control programme. Trans R Soc Trop Med Hyg 1996, 90(1):10-14. Ansar W, Bandyopadhyay SM, Chowdhury S, Habib SH, Mandal C: Role of C-reactive protein in complement-mediated hemolysis in Malaria. Glycoconj J 2006, 23(3-4):233-240. Garred P, Nielsen MA, Kurtzhals JA, Malhotra R, Madsen HO, Goka BQ, Akanmori BD, Sim RB, Hviid L: Mannose-binding lectin is a disease modifier in clinical malaria and may function as opsonin for Plasmodium falciparum-infected erythrocytes. Infect Immun 2003, 71(9):5245-5253. Klabunde J, Uhlemann AC, Tebo AE, Kimmel J, Schwarz RT, Kremsner PG, Kun JF: Recognition of plasmodium falciparum proteins by mannan-binding lectin, a component of the human innate immune system. Parasitol Res 2002, 88(2):113-117. Luty AJ, Kun JF, Kremsner PG: Mannose-binding lectin plasma levels and gene polymorphisms in Plasmodium falciparum malaria. J Infect Dis 1998, 178(4):1221-1224. Urban BC, Ing R, Stevenson MM: Early interactions between blood-stage plasmodium

78

66.

67. 68.

69. 70. 71. 72. 73. 74.

75. 76. 77. 78.

79. 80. 81. 82.

parasites and the immune system. Curr Top Microbiol Immunol 2005, 297:25-70. Turrini F, Ginsburg H, Bussolino F, Pescarmona GP, Serra MV, Arese P: Phagocytosis of Plasmodium falciparum-infected human red blood cells by human monocytes: involvement of immune and nonimmune determinants and dependence on parasite developmental stage. Blood 1992, 80(3):801-808. Celada A, Cruchaud A, Perrin LH: Opsonic activity of human immune serum on in vitro phagocytosis of Plasmodium falciparum infected red blood cells by monocytes. Clin Exp Immunol 1982, 47(3):635-644. Martiney JA, Sherry B, Metz CN, Espinoza M, Ferrer AS, Calandra T, Broxmeyer HE, Bucala R: Macrophage migration inhibitory factor release by macrophages after ingestion of Plasmodium chabaudi-infected erythrocytes: possible role in the pathogenesis of malarial anemia. Infect Immun 2000, 68(4):2259-2267. McDevitt MA, Xie J, Shanmugasundaram G, Griffith J, Liu A, McDonald C, Thuma P, Gordeuk VR, Metz CN, Mitchell R et al: A critical role for the host mediator macrophage migration inhibitory factor in the pathogenesis of malarial anemia. J Exp Med 2006, 203(5):1185-1196. Mohan K, Moulin P, Stevenson MM: Natural killer cell cytokine production, not cytotoxicity, contributes to resistance against blood-stage Plasmodium chabaudi AS infection. J Immunol 1997, 159(10):4990-4998. Artavanis-Tsakonas K, Riley EM: Innate immune response to malaria: rapid induction of IFN-gamma from human NK cells by live Plasmodium falciparum-infected erythrocytes. J Immunol 2002, 169(6):2956-2963. Artavanis-Tsakonas K, Eleme K, McQueen KL, Cheng NW, Parham P, Davis DM, Riley EM: Activation of a subset of human NK cells upon contact with Plasmodium falciparuminfected erythrocytes. J Immunol 2003, 171(10):5396-5405. Mavoungou E, Luty AJ, Kremsner PG: Natural killer (NK) cell-mediated cytolysis of Plasmodium falciparum-infected human red blood cells in vitro. Eur Cytokine Netw 2003, 14(3):134-142. Baratin M, Roetynck S, Lepolard C, Falk C, Sawadogo S, Uematsu S, Akira S, Ryffel B, Tiraby JG, Alexopoulou L et al: Natural killer cell and macrophage cooperation in MyD88dependent innate responses to Plasmodium falciparum. Proc Natl Acad Sci U S A 2005, 102(41):14747-14752. Rajalingam R, Gardiner CM, Canavez F, Vilches C, Parham P: Identification of seventeen novel KIR variants: fourteen of them from two non-Caucasian donors. Tissue Antigens 2001, 57(1):22-31. Korbel DS, Newman KC, Almeida CR, Davis DM, Riley EM: Heterogeneous human NK cell responses to Plasmodium falciparum-infected erythrocytes. J Immunol 2005, 175(11):74667473. Roetynck S, Baratin M, Johansson S, Lemmers C, Vivier E, Ugolini S: Natural killer cells and malaria. Immunol Rev 2006, 214:251-263. Sivori S, Falco M, Della Chiesa M, Carlomagno S, Vitale M, Moretta L, Moretta A: CpG and double-stranded RNA trigger human NK cells by Toll-like receptors: induction of cytokine release and cytotoxicity against tumors and dendritic cells. Proc Natl Acad Sci U S A 2004, 101(27):10116-10121. Lantz O, Bendelac A: An invariant T cell receptor alpha chain is used by a unique subset of major histocompatibility complex class I-specific CD4+ and CD4-8- T cells in mice and humans. J Exp Med 1994, 180(3):1097-1106. Taniguchi M, Harada M, Kojo S, Nakayama T, Wakao H: The regulatory role of Valpha14 NKT cells in innate and acquired immune response. Annu Rev Immunol 2003, 21:483-513. Taniguchi M, Seino K, Nakayama T: The NKT cell system: bridging innate and acquired immunity. Nat Immunol 2003, 4(12):1164-1165. Schofield L, McConville MJ, Hansen D, Campbell AS, Fraser-Reid B, Grusby MJ, Tachado SD:

79

83.

84.

85.

86. 87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99.

CD1d-restricted immunoglobulin G formation to GPI-anchored antigens mediated by NKT cells. Science 1999, 283(5399):225-229. Gonzalez JM, Peter K, Esposito F, Nebie I, Tiercy JM, Bonelo A, Arevalo-Herrera M, Valmori D, Romero P, Herrera S et al: HLA-A*0201 restricted CD8+ T-lymphocyte responses to malaria: identification of new Plasmodium falciparum epitopes by IFN-gamma ELISPOT. Parasite Immunol 2000, 22(10):501-514. Weerasinghe A, Sekikawa H, Watanabe H, Mannoor K, Morshed SR, Halder RC, Kawamura T, Kosaka T, Miyaji C, Kawamura H et al: Association of intermediate T cell receptor cells, mainly their NK1.1(-) subset, with protection from malaria. Cell Immunol 2001, 207(1):2835. Mannoor MK, Weerasinghe A, Halder RC, Reza S, Morshed M, Ariyasinghe A, Watanabe H, Sekikawa H, Abo T: Resistance to malarial infection is achieved by the cooperation of NK1.1(+) and NK1.1(-) subsets of intermediate TCR cells which are constituents of innate immunity. Cell Immunol 2001, 211(2):96-104. Haas W, Tonegawa S: Development and selection of gamma delta T cells. Curr Opin Immunol 1992, 4(2):147-155. Goodier M, Fey P, Eichmann K, Langhorne J: Human peripheral blood gamma delta T cells respond to antigens of Plasmodium falciparum. Int Immunol 1992, 4(1):33-41. Troye-Blomberg M, Berzins K, Perlmann P: T-cell control of immunity to the asexual blood stages of the malaria parasite. Crit Rev Immunol 1994, 14(2):131-155. Goodier MR, Lundqvist C, Hammarstrom ML, Troye-Blomberg M, Langhorne J: Cytokine profiles for human V gamma 9+ T cells stimulated by Plasmodium falciparum. Parasite Immunol 1995, 17(8):413-423. Elloso MM, van der Heyde HC, vande Waa JA, Manning DD, Weidanz WP: Inhibition of Plasmodium falciparum in vitro by human gamma delta T cells. J Immunol 1994, 153(3):1187-1194. Nussenzweig RS, Vanderberg J, Most H, Orton C: Protective immunity produced by the injection of x-irradiated sporozoites of plasmodium berghei. Nature 1967, 216(5111):160162. Hoffman SL, Goh LM, Luke TC, Schneider I, Le TP, Doolan DL, Sacci J, de la Vega P, Dowler M, Paul C et al: Protection of humans against malaria by immunization with radiationattenuated Plasmodium falciparum sporozoites. J Infect Dis 2002, 185(8):1155-1164. Nardin E, Zavala F, Nussenzweig V, Nussenzweig RS: Pre-erythrocytic malaria vaccine: mechanisms of protective immunity and human vaccine trials. Parassitologia 1999, 41(13):397-402. Ramirez AD, Rocha EM, Krettli AU: Antisporozoite antibodies with protective and nonprotective activities: in vitro and in vivo correlations using Plasmodium gallinaceum, an avian model. J Eukaryot Microbiol 1995, 42(6):705-708. Nardin EH, Nussenzweig RS, McGregor IA, Bryan JH: Antibodies to sporozoites: their frequent occurrence in individuals living in an area of hyperendemic malaria. Science 1979, 206(4418):597-599. Hollingdale MR, Nardin EH, Tharavanij S, Schwartz AL, Nussenzweig RS: Inhibition of entry of Plasmodium falciparum and P. vivax sporozoites into cultured cells; an in vitro assay of protective antibodies. J Immunol 1984, 132(2):909-913. Plebanski M, Aidoo M, Whittle HC, Hill AV: Precursor frequency analysis of cytotoxic T lymphocytes to pre-erythrocytic antigens of Plasmodium falciparum in West Africa. J Immunol 1997, 158(6):2849-2855. Plebanski M, Flanagan KL, Lee EA, Reece WH, Hart K, Gelder C, Gillespie G, Pinder M, Hill AV: Interleukin 10-mediated immunosuppression by a variant CD4 T cell epitope of Plasmodium falciparum. Immunity 1999, 10(6):651-660. Good MF, Doolan DL: Immune effector mechanisms in malaria. Curr Opin Immunol 1999,

80

100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120.

11(4):412-419. Doolan DL, Hoffman SL: The complexity of protective immunity against liver-stage malaria. J Immunol 2000, 165(3):1453-1462. Tsuji M, Zavala F: T cells as mediators of protective immunity against liver stages of Plasmodium. Trends Parasitol 2003, 19(2):88-93. Schofield L, Villaquiran J, Ferreira A, Schellekens H, Nussenzweig R, Nussenzweig V: Gamma interferon, CD8+ T cells and antibodies required for immunity to malaria sporozoites. Nature 1987, 330(6149):664-666. Sedegah M, Sim BK, Mason C, Nutman T, Malik A, Roberts C, Johnson A, Ochola J, Koech D, Were B et al: Naturally acquired CD8+ cytotoxic T lymphocytes against the Plasmodium falciparum circumsporozoite protein. J Immunol 1992, 149(3):966-971. Romero P, Maryanski JL, Corradin G, Nussenzweig RS, Nussenzweig V, Zavala F: Cloned cytotoxic T cells recognize an epitope in the circumsporozoite protein and protect against malaria. Nature 1989, 341(6240):323-326. Rodrigues MM, Cordey AS, Arreaza G, Corradin G, Romero P, Maryanski JL, Nussenzweig RS, Zavala F: CD8+ cytolytic T cell clones derived against the Plasmodium yoelii circumsporozoite protein protect against malaria. Int Immunol 1991, 3(6):579-585. Weiss WR, Berzofsky JA, Houghten RA, Sedegah M, Hollindale M, Hoffman SL: A T cell clone directed at the circumsporozoite protein which protects mice against both Plasmodium yoelii and Plasmodium berghei. J Immunol 1992, 149(6):2103-2109. Khusmith S, Sedegah M, Hoffman SL: Complete protection against Plasmodium yoelii by adoptive transfer of a CD8+ cytotoxic T-cell clone recognizing sporozoite surface protein 2. Infect Immun 1994, 62(7):2979-2983. Lalvani A, Hurt N, Aidoo M, Kibatala P, Tanner M, Hill AV: Cytotoxic T lymphocytes to Plasmodium falciparum epitopes in an area of intense and perennial transmission in Tanzania. Eur J Immunol 1996, 26(4):773-779. Berzins. K. PP: Malaria vaccines: attacking infected erythrocyts. . Washington, DC: American Society for Microbiology Press. ; 1996. Miller LH, Good MF, Kaslow DC: Vaccines against the blood stages of falciparum malaria. Adv Exp Med Biol 1998, 452:193-205. Cohen S, Mc GI, Carrington S: Gamma-globulin and acquired immunity to human malaria. Nature 1961, 192:733-737. McGregor I.A. CSP, Cohen S. : Treatment of East African P. falciparum malaria with West African human -globulin. . Trans R Soc Trop Med Hyg 1963, 57:170175. Sanford F. Kuvin AV: Malaria Antibody Titres of West Africans in Britain. Br Med J 1963, 24; (2(5355): ):477479. Freeman RR, Parish CR: Plasmodium yoelii: antibody and the maintenance of immunity in BALB/c mice. Exp Parasitol 1981, 52(1):18-24. Spencer Valero LM, Ogun SA, Fleck SL, Ling IT, Scott-Finnigan TJ, Blackman MJ, Holder AA: Passive immunization with antibodies against three distinct epitopes on Plasmodium yoelii merozoite surface protein 1 suppresses parasitemia. Infect Immun 1998, 66(8):3925-3930. von der Weid T, Honarvar N, Langhorne J: Gene-targeted mice lacking B cells are unable to eliminate a blood stage malaria infection. J Immunol 1996, 156(7):2510-2516. Langhorne J, Cross C, Seixas E, Li C, von der Weid T: A role for B cells in the development of T cell helper function in a malaria infection in mice. Proc Natl Acad Sci U S A 1998, 95(4):1730-1734. Mohan K. MMS: Acquired immunity to asexual blood-stages of malaria, . Washington, D.C.: ASM Press. ; 1998. Perlmann P, Troye-Blomberg M: Malaria and the immune system in humans. Chem Immunol 2002, 80:229-242. Giha HA, Staalsoe T, Dodoo D, Elhassan IM, Roper C, Satti GM, Arnot DE, Theander TG,

81

121. 122. 123. 124. 125. 126.

127. 128. 129. 130. 131. 132.

133. 134. 135. 136. 137. 138.

Hviid L: Nine-year longitudinal study of antibodies to variant antigens on the surface of Plasmodium falciparum-infected erythrocytes. Infect Immun 1999, 67(8):4092-4098. Ofori MF, Dodoo D, Staalsoe T, Kurtzhals JA, Koram K, Theander TG, Akanmori BD, Hviid L: Malaria-induced acquisition of antibodies to Plasmodium falciparum variant surface antigens. Infect Immun 2002, 70(6):2982-2988. Hviid L: Naturally acquired immunity to Plasmodium falciparum malaria in Africa. Acta Trop 2005, 95(3):270-275. Chizzolini C, Dupont A, Akue JP, Kaufmann MH, Verdini AS, Pessi A, Del Giudice G: Natural antibodies against three distinct and defined antigens of Plasmodium falciparum in residents of a mesoendemic area in Gabon. Am J Trop Med Hyg 1988, 39(2):150-156. Astagneau P, Roberts JM, Steketee RW, Wirima JJ, Lepers JP, Deloron P: Antibodies to a Plasmodium falciparum blood-stage antigen as a tool for predicting the protection levels of two malaria-exposed populations. Am J Trop Med Hyg 1995, 53(1):23-28. Piper KP, Hayward RE, Cox MJ, Day KP: Malaria transmission and naturally acquired immunity to PfEMP-1. Infect Immun 1999, 67(12):6369-6374. Riley EM, Allen SJ, Wheeler JG, Blackman MJ, Bennett S, Takacs B, Schonfeld HJ, Holder AA, Greenwood BM: Naturally acquired cellular and humoral immune responses to the major merozoite surface antigen (PfMSP1) of Plasmodium falciparum are associated with reduced malaria morbidity. Parasite Immunol 1992, 14(3):321-337. Nardin EH, Nussenzweig RS: T cell responses to pre-erythrocytic stages of malaria: role in protection and vaccine development against pre-erythrocytic stages. Annu Rev Immunol 1993, 11:687-727. Saul A: The role of variant surface antigens on malaria-infected red blood cells. Parasitol Today 1999, 15(11):455-457. Newbold C, Warn P, Black G, Berendt A, Craig A, Snow B, Msobo M, Peshu N, Marsh K: Receptor-specific adhesion and clinical disease in Plasmodium falciparum. Am J Trop Med Hyg 1997, 57(4):389-398. Groux H, Gysin J: Opsonization as an effector mechanism in human protection against asexual blood stages of Plasmodium falciparum: functional role of IgG subclasses. Res Immunol 1990, 141(6):529-542. Khusmith S, Druilhe P: Antibody-dependent ingestion of P. falciparum merozoites by human blood monocytes. Parasite Immunol 1983, 5(4):357-368. Bouharoun-Tayoun H, Attanath P, Sabchareon A, Chongsuphajaisiddhi T, Druilhe P: 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 1990, 172(6):1633-1641. Shear HL, Nussenzweig RS, Bianco C: Immune phagocytosis in murine malaria. J Exp Med 1979, 149(6):1288-1298. Mota MM, Brown KN, Holder AA, Jarra W: Acute Plasmodium chabaudi chabaudi malaria infection induces antibodies which bind to the surfaces of parasitized erythrocytes and promote their phagocytosis by macrophages in vitro. Infect Immun 1998, 66(9):4080-4086. Gabriel J, Berzins K: Specific lysis of Plasmodium yoelii infected mouse erythrocytes with antibody and complement. Clin Exp Immunol 1983, 52(1):129-134. Wahlin B, Wahlgren M, Perlmann H, Berzins K, Bjorkman A, Patarroyo ME, Perlmann P: Human antibodies to a Mr 155,000 Plasmodium falciparum antigen efficiently inhibit merozoite invasion. Proc Natl Acad Sci U S A 1984, 81(24):7912-7916. Rotman HL, Daly TM, Clynes R, Long CA: Fc receptors are not required for antibodymediated protection against lethal malaria challenge in a mouse model. J Immunol 1998, 161(4):1908-1912. Bouharoun-Tayoun H, Oeuvray C, Lunel F, Druilhe P: Mechanisms underlying the monocytemediated antibody-dependent killing of Plasmodium falciparum asexual blood stages. J

82

139. 140. 141. 142. 143. 144. 145. 146. 147. 148. 149.

150.

151. 152. 153. 154. 155. 156.

Exp Med 1995, 182(2):409-418. Brown IN, Allison AC, Taylor RB: Plasmodium berghei infections in thymectomized rats. Nature 1968, 219(5151):292-293. Brake DA, Long CA, Weidanz WP: Adoptive protection against Plasmodium chabaudi adami malaria in athymic nude mice by a cloned T cell line. J Immunol 1988, 140(6):19891993. Taylor-Robinson AW, Phillips RS, Severn A, Moncada S, Liew FY: The role of TH1 and TH2 cells in a rodent malaria infection. Science 1993, 260(5116):1931-1934. Amante FH, Good MF: Prolonged Th1-like response generated by a Plasmodium yoeliispecific T cell clone allows complete clearance of infection in reconstituted mice. Parasite Immunol 1997, 19(3):111-126. van der Heyde HC, Huszar D, Woodhouse C, Manning DD, Weidanz WP: The resolution of acute malaria in a definitive model of B cell deficiency, the JHD mouse. J Immunol 1994, 152(9):4557-4562. Langhorne J, Gillard S, Simon B, Slade S, Eichmann K: Frequencies of CD4+ T cells reactive with Plasmodium chabaudi chabaudi: distinct response kinetics for cells with Th1 and Th2 characteristics during infection. Int Immunol 1989, 1(4):416-424. Stevenson MM, Tam MF: Differential induction of helper T cell subsets during blood-stage Plasmodium chabaudi AS infection in resistant and susceptible mice. Clin Exp Immunol 1993, 92(1):77-83. Taylor-Robinson AW, Looker M: Sensitivity of malaria parasites to nitric oxide at low oxygen tensions. Lancet 1998, 351(9116):1630. Su Z, Stevenson MM: Central role of endogenous gamma interferon in protective immunity against blood-stage Plasmodium chabaudi AS infection. Infect Immun 2000, 68(8):4399-4406. Fell AH, Silins SL, Baumgarth N, Good MF: Plasmodium falciparum-specific T cell clones from non-exposed and exposed donors are highly diverse in TCR beta chain V segment usage. Int Immunol 1996, 8(12):1877-1887. Luty AJ, Lell B, Schmidt-Ott R, Lehman LG, Luckner D, Greve B, Matousek P, Herbich K, Schmid D, Migot-Nabias F et al: Interferon-gamma responses are associated with resistance to reinfection with Plasmodium falciparum in young African children. J Infect Dis 1999, 179(4):980-988. Pombo DJ, Lawrence G, Hirunpetcharat C, Rzepczyk C, Bryden M, Cloonan N, Anderson K, Mahakunkijcharoen Y, Martin LB, Wilson D et al: Immunity to malaria after administration of ultra-low doses of red cells infected with Plasmodium falciparum. Lancet 2002, 360(9333):610-617. Vinetz JM, Kumar S, Good MF, Fowlkes BJ, Berzofsky JA, Miller LH: Adoptive transfer of CD8+ T cells from immune animals does not transfer immunity to blood stage Plasmodium yoelii malaria. J Immunol 1990, 144(3):1069-1074. Kumar S, Miller LH: Cellular mechanisms in immunity to blood stage infection. Immunol Lett 1990, 25(1-3):109-114. van der Heyde HC, Manning DD, Roopenian DC, Weidanz WP: Resolution of blood-stage malarial infections in CD8+ cell-deficient beta 2-m0/0 mice. J Immunol 1993, 151(6):31873191. Riley EM, Jobe O, Whittle HC: CD8+ T cells inhibit Plasmodium falciparum-induced lymphoproliferation and gamma interferon production in cell preparations from some malaria-immune individuals. Infect Immun 1989, 57(4):1281-1284. Mshana RN, McLean S, Boulandi J: In vitro cell-mediated immune responses to Plasmodium falciparum schizont antigens in adults from a malaria endemic area: CD8+ T lymphocytes inhibit the response of low responder individuals. Int Immunol 1990, 2(12):1121-1132. Newton CR, Krishna S: Severe falciparum malaria in children: current understanding of pathophysiology and supportive treatment. Pharmacol Ther 1998, 79(1):1-53.

83

157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173. 174. 175. 176. 177. 178. 179.

Cooke BM, Tilley L: Key processes in malaria pathogenesis: keeping on top down under. Parasitol Today 1999, 15(5):176-178. Schofield L, Grau GE: Immunological processes in malaria pathogenesis. Nat Rev Immunol 2005, 5(9):722-735. Kwiatkowski D. aPP: Inflammatory Processes in the Pathogenesis of Malaria; In malaria Immunology, vol. 80. Basel, Switzerland. : KARGER,; 1999. Chen Q, Schlichtherle M, Wahlgren M: Molecular aspects of severe malaria. Clin Microbiol Rev 2000, 13(3):439-450. Kirchgatter K, Del Portillo HA: Clinical and molecular aspects of severe malaria. An Acad Bras Cienc 2005, 77(3):455-475. English M. NCRJC: Malaria Pathogenicity and Disease; In Malaria Immunology. , . Basel,: Karger.; 2002. Lou J, Lucas R, Grau GE: Pathogenesis of cerebral malaria: recent experimental data and possible applications for humans. Clin Microbiol Rev 2001, 14(4):810-820, table of contents. Clark IA, Cowden WB: The pathophysiology of falciparum malaria. Pharmacol Ther 2003, 99(2):221-260. Sergiev VP, Popov AF, Chirkov VP: [Cerebral malaria: pathogenesis, clinical picture and treatment]. Med Parazitol (Mosk) 2005(1):58-62. Snow RW, Nahlen B, Palmer A, Donnelly CA, Gupta S, Marsh K: Risk of severe malaria among African infants: direct evidence of clinical protection during early infancy. J Infect Dis 1998, 177(3):819-822. Clark IA, Rockett KA: The cytokine theory of human cerebral malaria. Parasitol Today 1994, 10(10):410-412. Grau GE, de Kossodo S: Cerebral malaria: mediators, mechanical obstruction or more? Parasitol Today 1994, 10(10):408-409. Berendt AR, Tumer GD, Newbold CI: Cerebral malaria: the sequestration hypothesis. Parasitol Today 1994, 10(10):412-414. van der Heyde HC, Nolan J, Combes V, Gramaglia I, Grau GE: A unified hypothesis for the genesis of cerebral malaria: sequestration, inflammation and hemostasis leading to microcirculatory dysfunction. Trends Parasitol 2006, 22(11):503-508. Turner G: Cerebral malaria. Brain Pathol 1997, 7(1):569-582. Haldar K, Murphy SC, Milner DA, Taylor TE: Malaria: Mechanisms of Erythrocytic Infection and Pathological Correlates of Severe Disease. Annu Rev Pathol 2007, 2:217-249. Clark IA, Hunt NH, Butcher GA, Cowden WB: Inhibition of murine malaria (Plasmodium chabaudi) in vivo by recombinant interferon-gamma or tumor necrosis factor, and its enhancement by butylated hydroxyanisole. J Immunol 1987, 139(10):3493-3496. Clark IA, Rockett KA, Cowden WB: Proposed link between cytokines, nitric oxide and human cerebral malaria. Parasitol Today 1991, 7(8):205-207. Clark IA, Rockett KA, Cowden WB: Role of TNF in cerebral malaria. Lancet 1991, 337(8736):302-303. Clark IA, Rockett KA, Cowden WB: Possible central role of nitric oxide in conditions clinically similar to cerebral malaria. Lancet 1992, 340(8824):894-896. Al Yaman FM, Mokela D, Genton B, Rockett KA, Alpers MP, Clark IA: Association between serum levels of reactive nitrogen intermediates and coma in children with cerebral malaria in Papua New Guinea. Trans R Soc Trop Med Hyg 1996, 90(3):270-273. Maneerat Y, Viriyavejakul P, Punpoowong B, Jones M, Wilairatana P, Pongponratn E, Turner GD, Udomsangpetch R: Inducible nitric oxide synthase expression is increased in the brain in fatal cerebral malaria. Histopathology 2000, 37(3):269-277. Schofield L, Novakovic S, Gerold P, Schwarz RT, McConville MJ, Tachado SD: Glycosylphosphatidylinositol toxin of Plasmodium up-regulates intercellular adhesion molecule-1, vascular cell adhesion molecule-1, and E-selectin expression in vascular

84

180. 181. 182. 183. 184. 185.

186.

187. 188. 189. 190. 191. 192. 193. 194. 195. 196. 197. 198. 199.

endothelial cells and increases leukocyte and parasite cytoadherence via tyrosine kinasedependent signal transduction. J Immunol 1996, 156(5):1886-1896. Schofield L, Hackett F: Signal transduction in host cells by a glycosylphosphatidylinositol toxin of malaria parasites. J Exp Med 1993, 177(1):145-153. Mazier D, Nitcheu J, Idrissa-Boubou M: Cerebral malaria and immunogenetics. Parasite Immunol 2000, 22(12):613-623. Wahlgren M. TCJaGJ: Cytoadherence and Rosetting in the Pathogenesis of Severe Malaria. : harwood academic publishers. ; 1999. Devine DV: The regulation of complement on cell surfaces. Transfus Med Rev 1991, 5(2):123-131. Menendez C, Fleming AF, Alonso PL: Malaria-related anaemia. Parasitol Today 2000, 16(11):469-476. McElroy PD, ter Kuile FO, Lal AA, Bloland PB, Hawley WA, Oloo AJ, Monto AS, Meshnick SR, Nahlen BL: Effect of Plasmodium falciparum parasitemia density on hemoglobin concentrations among full-term, normal birth weight children in western Kenya, IV. The Asembo Bay Cohort Project. Am J Trop Med Hyg 2000, 62(4):504-512. Ong'echa JM, Keller CC, Were T, Ouma C, Otieno RO, Landis-Lewis Z, Ochiel D, Slingluff JL, Mogere S, Ogonji GA et al: Parasitemia, anemia, and malarial anemia in infants and young children in a rural holoendemic Plasmodium falciparum transmission area. Am J Trop Med Hyg 2006, 74(3):376-385. Abdalla SH: Hematopoiesis in human malaria. Blood Cells 1990, 16(2-3):401-416; discussion 417-409. Burchard GD, Radloff P, Philipps J, Nkeyi M, Knobloch J, Kremsner PG: Increased erythropoietin production in children with severe malarial anemia. Am J Trop Med Hyg 1995, 53(5):547-551. el Hassan AM, Saeed AM, Fandrey J, Jelkmann W: Decreased erythropoietin response in Plasmodium falciparum malaria-associated anaemia. Eur J Haematol 1997, 59(5):299-304. Angus BJ, Chotivanich K, Udomsangpetch R, White NJ: In vivo removal of malaria parasites from red blood cells without their destruction in acute falciparum malaria. Blood 1997, 90(5):2037-2040. Jakeman GN, Saul A, Hogarth WL, Collins WE: Anaemia of acute malaria infections in nonimmune patients primarily results from destruction of uninfected erythrocytes. Parasitology 1999, 119 ( Pt 2):127-133. Wickramasinghe SN, Abdalla SH: Blood and bone marrow changes in malaria. Baillieres Best Pract Res Clin Haematol 2000, 13(2):277-299. Nagel RL: Malarial anemia. Hemoglobin 2002, 26(4):329-343. Chang KH, Stevenson MM: Malarial anaemia: mechanisms and implications of insufficient erythropoiesis during blood-stage malaria. Int J Parasitol 2004, 34(13-14):1501-1516. Sexton AC, Good RT, Hansen DS, D'Ombrain MC, Buckingham L, Simpson K, Schofield L: Transcriptional profiling reveals suppressed erythropoiesis, up-regulated glycolysis, and interferon-associated responses in murine malaria. J Infect Dis 2004, 189(7):1245-1256. Kurtzhals JA, Adabayeri V, Goka BQ, Akanmori BD, Oliver-Commey JO, Nkrumah FK, Behr C, Hviid L: Low plasma concentrations of interleukin 10 in severe malarial anaemia compared with cerebral and uncomplicated malaria. Lancet 1998, 351(9118):1768-1772. Biemba G, Gordeuk VR, Thuma P, Weiss G: Markers of inflammation in children with severe malarial anaemia. Trop Med Int Health 2000, 5(4):256-262. Othoro C, Lal AA, Nahlen B, Koech D, Orago AS, Udhayakumar V: 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 1999, 179(1):279-282. Mohan K, Stevenson MM: Dyserythropoiesis and severe anaemia associated with malaria correlate with deficient interleukin-12 production. Br J Haematol 1998, 103(4):942-949.

85

200. 201. 202. 203. 204.

205. 206. 207. 208. 209. 210. 211. 212.

213. 214. 215. 216. 217. 218. 219. 220.

Mohan K, Sam H, Stevenson MM: Therapy with a combination of low doses of interleukin 12 and chloroquine completely cures blood-stage malaria, prevents severe anemia, and induces immunity to reinfection. Infect Immun 1999, 67(2):513-519. Luty AJ, Perkins DJ, Lell B, Schmidt-Ott R, Lehman LG, Luckner D, Greve B, Matousek P, Herbich K, Schmid D et al: Low interleukin-12 activity in severe Plasmodium falciparum malaria. Infect Immun 2000, 68(7):3909-3915. Malaguarnera L, Pignatelli S, Simpore J, Malaguarnera M, Musumeci S: Plasma levels of interleukin-12 (IL-12), interleukin-18 (IL-18) and transforming growth factor beta (TGFbeta) in Plasmodium falciparum malaria. Eur Cytokine Netw 2002, 13(4):425-430. Perkins DJ, Weinberg JB, Kremsner PG: Reduced interleukin-12 and transforming growth factor-beta1 in severe childhood malaria: relationship of cytokine balance with disease severity. J Infect Dis 2000, 182(3):988-992. Chaisavaneeyakorn S, Othoro C, Shi YP, Otieno J, Chaiyaroj SC, Lal AA, Udhayakumar V: Relationship between plasma Interleukin-12 (IL-12) and IL-18 levels and severe malarial anemia in an area of holoendemicity in western Kenya. Clin Diagn Lab Immunol 2003, 10(3):362-366. Murray PJ, Young RA: Increased antimycobacterial immunity in interleukin-10-deficient mice. Infect Immun 1999, 67(6):3087-3095. Dinarello CA: Proinflammatory cytokines. Chest 2000, 118(2):503-508. Dinarello CA: Historical insights into cytokines. Eur J Immunol 2007, 37 Suppl 1:S34-45. Cannon P.R.: Some pathological aspects of human malaria. Washington, D.C, : American Association for the Advancement of Science. ; 1941. Maegraith B.: Pathological processes in malaria and blackwater fever,. United Kingdom: Blackwell, Oxford; 1948. Peyron F, Burdin N, Ringwald P, Vuillez JP, Rousset F, Banchereau J: High levels of circulating IL-10 in human malaria. Clin Exp Immunol 1994, 95(2):300-303. Chaiyaroj SC, Rutta AS, Muenthaisong K, Watkins P, Na Ubol M, Looareesuwan S: Reduced levels of transforming growth factor-beta1, interleukin-12 and increased migration inhibitory factor are associated with severe malaria. Acta Trop 2004, 89(3):319-327. Nussenblatt V, Mukasa G, Metzger A, Ndeezi G, Garrett E, Semba RD: Anemia and interleukin-10, tumor necrosis factor alpha, and erythropoietin levels among children with acute, uncomplicated Plasmodium falciparum malaria. Clin Diagn Lab Immunol 2001, 8(6):1164-1170. Dodoo D, Omer FM, Todd J, Akanmori BD, Koram KA, Riley EM: Absolute levels and ratios of proinflammatory and anti-inflammatory cytokine production in vitro predict clinical immunity to Plasmodium falciparum malaria. J Infect Dis 2002, 185(7):971-979. Riley EM: Is T-cell priming required for initiation of pathology in malaria infections? Immunol Today 1999, 20(5):228-233. Hunt NH, Grau GE: Cytokines: accelerators and brakes in the pathogenesis of cerebral malaria. Trends Immunol 2003, 24(9):491-499. Shear HL, Srinivasan R, Nolan T, Ng C: Role of IFN-gamma in lethal and nonlethal malaria in susceptible and resistant murine hosts. J Immunol 1989, 143(6):2038-2044. Meding SJ, Cheng SC, Simon-Haarhaus B, Langhorne J: Role of gamma interferon during infection with Plasmodium chabaudi chabaudi. Infect Immun 1990, 58(11):3671-3678. Stevenson MM, Tam MF, Belosevic M, van der Meide PH, Podoba JE: Role of endogenous gamma interferon in host response to infection with blood-stage Plasmodium chabaudi AS. Infect Immun 1990, 58(10):3225-3232. Favre N, Ryffel B, Bordmann G, Rudin W: The course of Plasmodium chabaudi chabaudi infections in interferon-gamma receptor deficient mice. Parasite Immunol 1997, 19(8):375383. van der Heyde HC, Pepper B, Batchelder J, Cigel F, Weidanz WP: The time course of selected

86

221. 222. 223. 224. 225.

226. 227. 228. 229. 230. 231. 232. 233. 234. 235. 236. 237. 238.

malarial infections in cytokine-deficient mice. Exp Parasitol 1997, 85(2):206-213. Ho M, Sexton MM, Tongtawe P, Looareesuwan S, Suntharasamai P, Webster HK: Interleukin10 inhibits tumor necrosis factor production but not antigen-specific lymphoproliferation in acute Plasmodium falciparum malaria. J Infect Dis 1995, 172(3):838-844. Kwiatkowski D, Hill AV, Sambou I, Twumasi P, Castracane J, Manogue KR, Cerami A, Brewster DR, Greenwood BM: TNF concentration in fatal cerebral, non-fatal cerebral, and uncomplicated Plasmodium falciparum malaria. Lancet 1990, 336(8725):1201-1204. Ringwald P, Peyron F, Vuillez JP, Touze JE, Le Bras J, Deloron P: Levels of cytokines in plasma during Plasmodium falciparum malaria attacks. J Clin Microbiol 1991, 29(9):20762078. Harpaz R, Edelman R, Wasserman SS, Levine MM, Davis JR, Sztein MB: Serum cytokine profiles in experimental human malaria. Relationship to protection and disease course after challenge. J Clin Invest 1992, 90(2):515-523. Elghazali G, Esposito F, Troye-Blomberg M: Comparison of the number of IL-4 and IFNgamma secreting cells in response to the malaria vaccine candidate antigen Pf155/RESA in two groups of naturally primed individuals living in a malaria endemic area in Burkina Faso. Scand J Immunol 1995, 42(1):39-45. Wenisch C, Parschalk B, Narzt E, Looareesuwan S, Graninger W: Elevated serum levels of IL10 and IFN-gamma in patients with acute Plasmodium falciparum malaria. Clin Immunol Immunopathol 1995, 74(1):115-117. Koch O, Awomoyi A, Usen S, Jallow M, Richardson A, Hull J, Pinder M, Newport M, Kwiatkowski D: IFNGR1 gene promoter polymorphisms and susceptibility to cerebral malaria. J Infect Dis 2002, 185(11):1684-1687. Gimenez F, Barraud de Lagerie S, Fernandez C, Pino P, Mazier D: Tumor necrosis factor alpha in the pathogenesis of cerebral malaria. Cell Mol Life Sci 2003, 60(8):1623-1635. Grau GE: Essential role of tumor necrosis factor and other cytokines in the pathogenesis of cerebral malaria: experimental and clinical studies. Verh K Acad Geneeskd Belg 1992, 54(2):155-175. Clark IA: How TNF was recognized as a key mechanism of disease. Cytokine Growth Factor Rev 2007, 18(3-4):335-343. Kwiatkowski D: Cytokines and anti-disease immunity to malaria. Res Immunol 1991, 142(8):707-712. Grau GE, Fajardo LF, Piguet PF, Allet B, Lambert PH, Vassalli P: Tumor necrosis factor (cachectin) as an essential mediator in murine cerebral malaria. Science 1987, 237(4819):1210-1212. Grau GE, Taylor TE, Molyneux ME, Wirima JJ, Vassalli P, Hommel M, Lambert PH: Tumor necrosis factor and disease severity in children with falciparum malaria. N Engl J Med 1989, 320(24):1586-1591. Grau GE, Piguet PF, Vassalli P, Lambert PH: Tumor-necrosis factor and other cytokines in cerebral malaria: experimental and clinical data. Immunol Rev 1989, 112:49-70. Kwiatkowski D, Molyneux ME, Stephens S, Curtis N, Klein N, Pointaire P, Smit M, Allan R, Brewster DR, Grau GE et al: Anti-TNF therapy inhibits fever in cerebral malaria. Q J Med 1993, 86(2):91-98. Clark IA, Cowden WB: Roles of TNF in malaria and other parasitic infections. Immunol Ser 1992, 56:365-407. Clark IA, Budd AC, Alleva LM, Cowden WB: Human malarial disease: a consequence of inflammatory cytokine release. Malar J 2006, 5:85. Clark K, Kulk N, Amante F, Haque A, Engwerda C: Lymphotoxin alpha and tumour necrosis factor are not required for control of parasite growth, but differentially regulate cytokine production during Plasmodium chabaudi chabaudi AS infection. Parasite Immunol 2007, 29(3):153-158.

87

239. 240. 241. 242.

243. 244. 245.

246.

247. 248.

249. 250.

251. 252. 253.

Scuderi P, Sterling KE, Lam KS, Finley PR, Ryan KJ, Ray CG, Petersen E, Slymen DJ, Salmon SE: Raised serum levels of tumour necrosis factor in parasitic infections. Lancet 1986, 2(8520):1364-1365. Kern P, Hemmer CJ, Van Damme J, Gruss HJ, Dietrich M: Elevated tumor necrosis factor alpha and interleukin-6 serum levels as markers for complicated Plasmodium falciparum malaria. Am J Med 1989, 87(2):139-143. McGuire W, D'Alessandro U, Stephens S, Olaleye BO, Langerock P, Greenwood BM, Kwiatkowski D: Levels of tumour necrosis factor and soluble TNF receptors during malaria fever episodes in the community. Trans R Soc Trop Med Hyg 1998, 92(1):50-53. Jakobsen PH, McKay V, Morris-Jones SD, McGuire W, van Hensbroek MB, Meisner S, Bendtzen K, Schousboe I, Bygbjerg IC, Greenwood BM: Increased concentrations of interleukin-6 and interleukin-1 receptor antagonist and decreased concentrations of beta2-glycoprotein I in Gambian children with cerebral malaria. Infect Immun 1994, 62(10):4374-4379. McGuire W, Hill AV, Allsopp CE, Greenwood BM, Kwiatkowski D: Variation in the TNFalpha promoter region associated with susceptibility to cerebral malaria. Nature 1994, 371(6497):508-510. Knight JC, Kwiatkowski D: Inherited variability of tumor necrosis factor production and susceptibility to infectious disease. Proc Assoc Am Physicians 1999, 111(4):290-298. Lyke KE, Burges R, Cissoko Y, Sangare L, Dao M, Diarra I, Kone A, Harley R, Plowe CV, Doumbo OK et al: Serum levels of the proinflammatory cytokines interleukin-1 beta (IL1beta), IL-6, IL-8, IL-10, tumor necrosis factor alpha, and IL-12(p70) in Malian children with severe Plasmodium falciparum malaria and matched uncomplicated malaria or healthy controls. Infect Immun 2004, 72(10):5630-5637. Turner GD, Morrison H, Jones M, Davis TM, Looareesuwan S, Buley ID, Gatter KC, Newbold CI, Pukritayakamee S, Nagachinta B et al: An immunohistochemical study of the pathology of fatal malaria. Evidence for widespread endothelial activation and a potential role for intercellular adhesion molecule-1 in cerebral sequestration. Am J Pathol 1994, 145(5):10571069. Prudhomme JG, Sherman IW, Land KM, Moses AV, Stenglein S, Nelson JA: Studies of Plasmodium falciparum cytoadherence using immortalized human brain capillary endothelial cells. Int J Parasitol 1996, 26(6):647-655. Krishna S, Waller DW, ter Kuile F, Kwiatkowski D, Crawley J, Craddock CF, Nosten F, Chapman D, Brewster D, Holloway PA et al: Lactic acidosis and hypoglycaemia in children with severe malaria: pathophysiological and prognostic significance. Trans R Soc Trop Med Hyg 1994, 88(1):67-73. van Hensbroek MB, Palmer A, Onyiorah E, Schneider G, Jaffar S, Dolan G, Memming H, Frenkel J, Enwere G, Bennett S et al: The effect of a monoclonal antibody to tumor necrosis factor on survival from childhood cerebral malaria. J Infect Dis 1996, 174(5):1091-1097. Hemmer CJ, Hort G, Chiwakata CB, Seitz R, Egbring R, Gaus W, Hogel J, Hassemer M, Nawroth PP, Kern P et al: Supportive pentoxifylline in falciparum malaria: no effect on tumor necrosis factor alpha levels or clinical outcome: a prospective, randomized, placebocontrolled study. Am J Trop Med Hyg 1997, 56(4):397-403. Engwerda CR, Mynott TL, Sawhney S, De Souza JB, Bickle QD, Kaye PM: Locally upregulated lymphotoxin alpha, not systemic tumor necrosis factor alpha, is the principle mediator of murine cerebral malaria. J Exp Med 2002, 195(10):1371-1377. Rockett KA, Awburn MM, Aggarwal BB, Cowden WB, Clark IA: In vivo induction of nitrite and nitrate by tumor necrosis factor, lymphotoxin, and interleukin-1: possible roles in malaria. Infect Immun 1992, 60(9):3725-3730. Clark IA, Gray KM, Rockett EJ, Cowden WB, Rockett KA, Ferrante A, Aggarwal BB: Increased lymphotoxin in human malarial serum, and the ability of this cytokine to

88

254. 255. 256. 257.

258. 259. 260. 261. 262. 263. 264. 265. 266. 267. 268. 269. 270. 271. 272. 273. 274.

increase plasma interleukin-6 and cause hypoglycaemia in mice: implications for malarial pathology. Trans R Soc Trop Med Hyg 1992, 86(6):602-607. Dinarello CA: The interleukin-1 family: 10 years of discovery. Faseb J 1994, 8(15):13141325. Dinarello CA: Biologic basis for interleukin-1 in disease. Blood 1996, 87(6):2095-2147. Nakae S, Asano M, Horai R, Sakaguchi N, Iwakura Y: IL-1 enhances T cell-dependent antibody production through induction of CD40 ligand and OX40 on T cells. J Immunol 2001, 167(1):90-97. Mshana RN, Boulandi J, Mshana NM, Mayombo J, Mendome G: 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 1991, 34(3):131-139. Nelms K, Keegan AD, Zamorano J, Ryan JJ, Paul WE: The IL-4 receptor: signaling mechanisms and biologic functions. Annu Rev Immunol 1999, 17:701-738. Wurtz O, Bajenoff M, Guerder S: IL-4-mediated inhibition of IFN-gamma production by CD4+ T cells proceeds by several developmentally regulated mechanisms. Int Immunol 2004, 16(3):501-508. Gaya A, de la Calle O, Yague J, Alsinet E, Fernandez MD, Romero M, Fabregat V, Martorell J, Vives J: IL-4 inhibits IL-2 synthesis and IL-2-induced up-regulation of IL-2R alpha but not IL-2R beta chain in CD4+ human T cells. J Immunol 1991, 146(12):4209-4214. Seder RA, Paul WE: Acquisition of lymphokine-producing phenotype by CD4+ T cells. Annu Rev Immunol 1994, 12:635-673. Seder RA: Acquisition of lymphokine-producing phenotype by CD4+ T cells. J Allergy Clin Immunol 1994, 94(6 Pt 2):1195-1202. Romagnani S: Biology of human TH1 and TH2 cells. J Clin Immunol 1995, 15(3):121-129. Wenner CA, Szabo SJ, Murphy KM: Identification of IL-4 promoter elements conferring Th2-restricted expression during T helper cell subset development. J Immunol 1997, 158(2):765-773. Li B, Tournier C, Davis RJ, Flavell RA: Regulation of IL-4 expression by the transcription factor JunB during T helper cell differentiation. Embo J 1999, 18(2):420-432. Vitetta ES, Brooks K, Chen YW, Isakson P, Jones S, Layton J, Mishra GC, Pure E, Weiss E, Word C et al: T-cell-derived lymphokines that induce IgM and IgG secretion in activated murine B cells. Immunol Rev 1984, 78:137-157. Del Prete G, Maggi E, Parronchi P, Chretien I, Tiri A, Macchia D, Ricci M, Banchereau J, De Vries J, Romagnani S: IL-4 is an essential factor for the IgE synthesis induced in vitro by human T cell clones and their supernatants. J Immunol 1988, 140(12):4193-4198. Morrot A, Hafalla JC, Cockburn IA, Carvalho LH, Zavala F: IL-4 receptor expression on CD8+ T cells is required for the development of protective memory responses against liver stages of malaria parasites. J Exp Med 2005, 202(4):551-560. Van Snick J: Interleukin-6: an overview. Annu Rev Immunol 1990, 8:253-278. Jones SA: Directing transition from innate to acquired immunity: defining a role for IL-6. J Immunol 2005, 175(6):3463-3468. Akira S, Taga T, Kishimoto T: Interleukin-6 in biology and medicine. Adv Immunol 1993, 54:1-78. Teague TK, Marrack P, Kappler JW, Vella AT: IL-6 rescues resting mouse T cells from apoptosis. J Immunol 1997, 158(12):5791-5796. Demoulin JB, Maisin D, Renauld JC: Ly-6A/E induction by interleukin-6 and interleukin-9 in T cells. Eur Cytokine Netw 1999, 10(1):49-56. La Flamme AC, Pearce EJ: The absence of IL-6 does not affect Th2 cell development in vivo, but does lead to impaired proliferation, IL-2 receptor expression, and B cell responses. J Immunol 1999, 162(10):5829-5837.

89

275. 276. 277. 278. 279. 280.

281. 282. 283. 284. 285. 286. 287. 288. 289. 290. 291. 292. 293. 294.

Teague TK, Schaefer BC, Hildeman D, Bender J, Mitchell T, Kappler JW, Marrack P: Activation-induced inhibition of interleukin 6-mediated T cell survival and signal transducer and activator of transcription 1 signaling. J Exp Med 2000, 191(6):915-926. Pied S, Civas A, Berlot-Picard F, Renia L, Miltgen F, Gentilini M, Doly J, Mazier D: IL-6 induced by IL-1 inhibits malaria pre-erythrocytic stages but its secretion is down-regulated by the parasite. J Immunol 1992, 148(1):197-201. Molyneux ME, Taylor TE, Wirima JJ, Grau GE: Tumour necrosis factor, interleukin-6, and malaria. Lancet 1991, 337(8749):1098. el-Nashar TM, el-Kholy HM, el-Shiety AG, Al-Zahaby AA: Correlation of plasma levels of tumor necrosis factor, interleukin-6 and nitric oxide with the severity of human malaria. J Egypt Soc Parasitol 2002, 32(2):525-535. Prakash D, Fesel C, Jain R, Cazenave PA, Mishra GC, Pied S: Clusters of cytokines determine malaria severity in Plasmodium falciparum-infected patients from endemic areas of Central India. J Infect Dis 2006, 194(2):198-207. Ringwald P, Peyron F, Lepers JP, Rabarison P, Rakotomalala C, Razanamparany M, Rabodonirina M, Roux J, Le Bras J: Parasite virulence factors during falciparum malaria: rosetting, cytoadherence, and modulation of cytoadherence by cytokines. Infect Immun 1993, 61(12):5198-5204. Baptista JL, Vanham G, Wery M, Van Marck E: Cytokine levels during mild and cerebral falciparum malaria in children living in a mesoendemic area. Trop Med Int Health 1997, 2(7):673-679. Day NP, Hien TT, Schollaardt T, Loc PP, Chuong LV, Chau TT, Mai NT, Phu NH, Sinh DX, White NJ et al: The prognostic and pathophysiologic role of pro- and antiinflammatory cytokines in severe malaria. J Infect Dis 1999, 180(4):1288-1297. Burgmann H, Hollenstein U, Wenisch C, Thalhammer F, Looareesuwan S, Graninger W: Serum concentrations of MIP-1 alpha and interleukin-8 in patients suffering from acute Plasmodium falciparum malaria. Clin Immunol Immunopathol 1995, 76(1 Pt 1):32-36. Friedland JS, Ho M, Remick DG, Bunnag D, White NJ, Griffin GE: Interleukin-8 and Plasmodium falciparum malaria in Thailand. Trans R Soc Trop Med Hyg 1993, 87(1):54-55. Horuk R: The interleukin-8-receptor family: from chemokines to malaria. Immunol Today 1994, 15(4):169-174. Fiorentino DF, Bond MW, Mosmann TR: Two types of mouse T helper cell. IV. Th2 clones secrete a factor that inhibits cytokine production by Th1 clones. J Exp Med 1989, 170(6):2081-2095. Lalani I, Bhol K, Ahmed AR: Interleukin-10: biology, role in inflammation and autoimmunity. Ann Allergy Asthma Immunol 1997, 79(6):469-483. Fiorentino DF, Zlotnik A, Vieira P, Mosmann TR, Howard M, Moore KW, O'Garra A: IL-10 acts on the antigen-presenting cell to inhibit cytokine production by Th1 cells. J Immunol 1991, 146(10):3444-3451. de Waal Malefyt R. MKW: In Interleukin-10,. San Diego, CA: : Academic Press. ; 1998. Fiorentino DF, Zlotnik A, Mosmann TR, Howard M, O'Garra A: IL-10 inhibits cytokine production by activated macrophages. J Immunol 1991, 147(11):3815-3822. Macatonia SE, Doherty TM, Knight SC, O'Garra A: Differential effect of IL-10 on dendritic cell-induced T cell proliferation and IFN-gamma production. J Immunol 1993, 150(9):37553765. Enk AH, Katz SI: Identification and induction of keratinocyte-derived IL-10. J Immunol 1992, 149(1):92-95. Moore KW, Vieira P, Fiorentino DF, Trounstine ML, Khan TA, Mosmann TR: Homology of cytokine synthesis inhibitory factor (IL-10) to the Epstein-Barr virus gene BCRFI. Science 1990, 248(4960):1230-1234. Wu J, Cunha FQ, Liew FY, Weiser WY: IL-10 inhibits the synthesis of migration inhibitory

90

295.

296. 297.

298. 299. 300. 301. 302. 303. 304. 305.

306.

307. 308.

309. 310.

factor and migration inhibitory factor-mediated macrophage activation. J Immunol 1993, 151(8):4325-4332. Takayama T, Morelli AE, Onai N, Hirao M, Matsushima K, Tahara H, Thomson AW: Mammalian and viral IL-10 enhance C-C chemokine receptor 5 but down-regulate C-C chemokine receptor 7 expression by myeloid dendritic cells: impact on chemotactic responses and in vivo homing ability. J Immunol 2001, 166(12):7136-7143. Clerici M, Wynn TA, Berzofsky JA, Blatt SP, Hendrix CW, Sher A, Coffman RL, Shearer GM: Role of interleukin-10 in T helper cell dysfunction in asymptomatic individuals infected with the human immunodeficiency virus. J Clin Invest 1994, 93(2):768-775. Gerosa F, Nisii C, Righetti S, Micciolo R, Marchesini M, Cazzadori A, Trinchieri G: CD4(+) T cell clones producing both interferon-gamma and interleukin-10 predominate in bronchoalveolar lavages of active pulmonary tuberculosis patients. Clin Immunol 1999, 92(3):224-234. Tripp CS, Beckerman KP, Unanue ER: Immune complexes inhibit antimicrobial responses through interleukin-10 production. Effects in severe combined immunodeficient mice during Listeria infection. J Clin Invest 1995, 95(4):1628-1634. Hunter CA, Ellis-Neyes LA, Slifer T, Kanaly S, Grunig G, Fort M, Rennick D, Araujo FG: IL10 is required to prevent immune hyperactivity during infection with Trypanosoma cruzi. J Immunol 1997, 158(7):3311-3316. Moore KW, de Waal Malefyt R, Coffman RL, O'Garra A: Interleukin-10 and the interleukin10 receptor. Annu Rev Immunol 2001, 19:683-765. Mocellin S, Marincola F, Rossi CR, Nitti D, Lise M: The multifaceted relationship between IL-10 and adaptive immunity: putting together the pieces of a puzzle. Cytokine Growth Factor Rev 2004, 15(1):61-76. Rousset F, Garcia E, Defrance T, Peronne C, Vezzio N, Hsu DH, Kastelein R, Moore KW, Banchereau J: Interleukin 10 is a potent growth and differentiation factor for activated human B lymphocytes. Proc Natl Acad Sci U S A 1992, 89(5):1890-1893. Cai G, Kastelein RA, Hunter CA: IL-10 enhances NK cell proliferation, cytotoxicity and production of IFN-gamma when combined with IL-18. Eur J Immunol 1999, 29(9):26582665. Ho M, Schollaardt T, Snape S, Looareesuwan S, Suntharasamai P, White NJ: Endogenous interleukin-10 modulates proinflammatory response in Plasmodium falciparum malaria. J Infect Dis 1998, 178(2):520-525. Walther M, Tongren JE, Andrews L, Korbel D, King E, Fletcher H, Andersen RF, Bejon P, Thompson F, Dunachie SJ et al: Upregulation of TGF-beta, FOXP3, and CD4+CD25+ regulatory T cells correlates with more rapid parasite growth in human malaria infection. Immunity 2005, 23(3):287-296. Walther M, Woodruff J, Edele F, Jeffries D, Tongren JE, King E, Andrews L, Bejon P, Gilbert SC, De Souza JB et al: Innate immune responses to human malaria: heterogeneous cytokine responses to blood-stage Plasmodium falciparum correlate with parasitological and clinical outcomes. J Immunol 2006, 177(8):5736-5745. Kossodo S, Monso C, Juillard P, Velu T, Goldman M, Grau GE: Interleukin-10 modulates susceptibility in experimental cerebral malaria. Immunology 1997, 91(4):536-540. Sarthou JL, Angel G, Aribot G, Rogier C, Dieye A, Toure Balde A, Diatta B, Seignot P, Roussilhon C: Prognostic value of anti-Plasmodium falciparum-specific immunoglobulin G3, cytokines, and their soluble receptors in West African patients with severe malaria. Infect Immun 1997, 65(8):3271-3276. Hugosson E, Montgomery SM, Premji Z, Troye-Blomberg M, Bjorkman A: Higher IL-10 levels are associated with less effective clearance of Plasmodium falciparum parasites. Parasite Immunol 2004, 26(3):111-117. Eckwalanga M, Marussig M, Tavares MD, Bouanga JC, Hulier E, Pavlovitch JH, Minoprio P,

91

311. 312. 313. 314. 315. 316. 317. 318. 319.

320. 321. 322. 323.

324. 325. 326. 327.

Portnoi D, Renia L, Mazier D: Murine AIDS protects mice against experimental cerebral malaria: down-regulation by interleukin 10 of a T-helper type 1 CD4+ cell-mediated pathology. Proc Natl Acad Sci U S A 1994, 91(17):8097-8101. May J, Lell B, Luty AJ, Meyer CG, Kremsner PG: Plasma interleukin-10:Tumor necrosis factor (TNF)-alpha ratio is associated with TNF promoter variants and predicts malarial complications. J Infect Dis 2000, 182(5):1570-1573. Stevenson MM, Tam MF, Wolf SF, Sher A: IL-12-induced protection against blood-stage Plasmodium chabaudi AS requires IFN-gamma and TNF-alpha and occurs via a nitric oxide-dependent mechanism. J Immunol 1995, 155(5):2545-2556. Doolan DL, Hoffman SL: IL-12 and NK cells are required for antigen-specific adaptive immunity against malaria initiated by CD8+ T cells in the Plasmodium yoelii model. J Immunol 1999, 163(2):884-892. Rhee MS, Akanmori BD, Waterfall M, Riley EM: Changes in cytokine production associated with acquired immunity to Plasmodium falciparum malaria. Clin Exp Immunol 2001, 126(3):503-510. Stevenson MM, Su Z, Sam H, Mohan K: Modulation of host responses to blood-stage malaria by interleukin-12: from therapy to adjuvant activity. Microbes Infect 2001, 3(1):49-59. Su Z, Stevenson MM: IL-12 is required for antibody-mediated protective immunity against blood-stage Plasmodium chabaudi AS malaria infection in mice. J Immunol 2002, 168(3):1348-1355. Sedegah M, Finkelman F, Hoffman SL: Interleukin 12 induction of interferon gammadependent protection against malaria. Proc Natl Acad Sci U S A 1994, 91(22):10700-10702. Hoffman SL, Crutcher JM, Puri SK, Ansari AA, Villinger F, Franke ED, Singh PP, Finkelman F, Gately MK, Dutta GP et al: Sterile protection of monkeys against malaria after administration of interleukin-12. Nat Med 1997, 3(1):80-83. Xing Z, Zganiacz A, Wang J, Divangahi M, Nawaz F: IL-12-independent Th1-type immune responses to respiratory viral infection: requirement of IL-18 for IFN-gamma release in the lung but not for the differentiation of viral-reactive Th1-type lymphocytes. J Immunol 2000, 164(5):2575-2584. Singh RP, Kashiwamura S, Rao P, Okamura H, Mukherjee A, Chauhan VS: The role of IL-18 in blood-stage immunity against murine malaria Plasmodium yoelii 265 and Plasmodium berghei ANKA. J Immunol 2002, 168(9):4674-4681. Riopel J, Tam M, Mohan K, Marino MW, Stevenson MM: Granulocyte-macrophage colonystimulating factor-deficient mice have impaired resistance to blood-stage malaria. Infect Immun 2001, 69(1):129-136. Kumaratilake LM, Ferrante A, Jaeger T, Rzepczyk C: 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 1996, 18(3):115-123. Weiss WR, Ishii KJ, Hedstrom RC, Sedegah M, Ichino M, Barnhart K, Klinman DM, Hoffman SL: A plasmid encoding murine granulocyte-macrophage colony-stimulating factor increases protection conferred by a malaria DNA vaccine. J Immunol 1998, 161(5):23252332. Owhashi M, Kirai N, Asami M: Temporary appearance of a circulating granulocytemacrophage colony-stimulating factor in lethal murine malaria. Southeast Asian J Trop Med Public Health 1996, 27(3):530-534. Wenisch C, Parschalk B, Burgmann H, Looareesuwan S, Graninger W: Decreased serum levels of TGF-beta in patients with acute Plasmodium falciparum malaria. J Clin Immunol 1995, 15(2):69-73. Omer FM, Riley EM: Transforming growth factor beta production is inversely correlated with severity of murine malaria infection. J Exp Med 1998, 188(1):39-48. Omer FM, Kurtzhals JA, Riley EM: Maintaining the immunological balance in parasitic

92

328. 329. 330. 331. 332. 333. 334. 335. 336. 337. 338. 339. 340. 341. 342. 343. 344. 345. 346.

infections: a role for TGF-beta? Parasitol Today 2000, 16(1):18-23. Omer FM, de Souza JB, Riley EM: Differential induction of TGF-beta regulates proinflammatory cytokine production and determines the outcome of lethal and nonlethal Plasmodium yoelii infections. J Immunol 2003, 171(10):5430-5436. Tsutsui N, Kamiyama T: Transforming growth factor beta-induced failure of resistance to infection with blood-stage Plasmodium chabaudi in mice. Infect Immun 1999, 67(5):23062311. Ocana-Morgner C, Wong KA, Lega F, Dotor J, Borras-Cuesta F, Rodriguez A: Role of TGFbeta and PGE2 in T cell responses during Plasmodium yoelii infection. Eur J Immunol 2007, 37(6):1562-1574. Wang WY, Barratt BJ, Clayton DG, Todd JA: Genome-wide association studies: theoretical and practical concerns. Nat Rev Genet 2005, 6(2):109-118. Khoury MJ, Yang Q: The future of genetic studies of complex human diseases: an epidemiologic perspective. Epidemiology 1998, 9(3):350-354. Risch N, Merikangas K: The future of genetic studies of complex human diseases. Science 1996, 273(5281):1516-1517. Westendorp RG, Langermans JA, Huizinga TW, Elouali AH, Verweij CL, Boomsma DI, Vandenbroucke JP: Genetic influence on cytokine production and fatal meningococcal disease. Lancet 1997, 349(9046):170-173. Bidwell J, Keen L, Gallagher G, Kimberly R, Huizinga T, McDermott MF, Oksenberg J, McNicholl J, Pociot F, Hardt C et al: Cytokine gene polymorphism in human disease: on-line databases. Genes Immun 1999, 1(1):3-19. Bidwell J, Keen L, Gallagher G, Kimberly R, Huizinga T, McDermott MF, Oksenberg J, McNicholl J, Pociot F, Hardt C et al: Cytokine gene polymorphism in human disease: on-line databases, supplement 1. Genes Immun 2001, 2(2):61-70. Haukim N, Bidwell JL, Smith AJ, Keen LJ, Gallagher G, Kimberly R, Huizinga T, McDermott MF, Oksenberg J, McNicholl J et al: Cytokine gene polymorphism in human disease: on-line databases, supplement 2. Genes Immun 2002, 3(6):313-330. Hollegaard MV, Bidwell JL: Cytokine gene polymorphism in human disease: on-line databases, Supplement 3. Genes Immun 2006, 7(4):269-276. Bayley JP, Ottenhoff TH, Verweij CL: Is there a future for TNF promoter polymorphisms? Genes Immun 2004, 5(5):315-329. McVean G, Spencer CC, Chaix R: Perspectives on human genetic variation from the HapMap Project. PLoS Genet 2005, 1(4):e54. Kwiatkowski D: Genetic susceptibility to malaria getting complex. Curr Opin Genet Dev 2000, 10(3):320-324. Kwiatkowski D: Genetic dissection of the molecular pathogenesis of severe infection. Intensive Care Med 2000, 26 Suppl 1:S89-97. Wilson AG, Symons JA, McDowell TL, McDevitt HO, Duff GW: Effects of a polymorphism in the human tumor necrosis factor alpha promoter on transcriptional activation. Proc Natl Acad Sci U S A 1997, 94(7):3195-3199. Warzocha K, Ribeiro P, Bienvenu J, Roy P, Charlot C, Rigal D, Coiffier B, Salles G: Genetic polymorphisms in the tumor necrosis factor locus influence non-Hodgkin's lymphoma outcome. Blood 1998, 91(10):3574-3581. McGuire W, Knight JC, Hill AV, Allsopp CE, Greenwood BM, Kwiatkowski D: Severe malarial anemia and cerebral malaria are associated with different tumor necrosis factor promoter alleles. J Infect Dis 1999, 179(1):287-290. Rieth H, Mormann M, Luty AJ, Assohou-Luty CA, Roupelieva M, Kremsner PG, Kube D: A three base pair gene variation within the distal 5'-flanking region of the interleukin-10 (IL10) gene is related to the in vitro IL-10 production capacity of lipopolysaccharidestimulated peripheral blood mononuclear cells. Eur Cytokine Netw 2004, 15(2):153-158.

93

347. 348. 349. 350. 351. 352. 353. 354. 355. 356. 357. 358.

359. 360. 361. 362. 363.

364.

Turner DM, Williams DM, Sankaran D, Lazarus M, Sinnott PJ, Hutchinson IV: An investigation of polymorphism in the interleukin-10 gene promoter. Eur J Immunogenet 1997, 24(1):1-8. Eskdale J, Kube D, Tesch H, Gallagher G: Mapping of the human IL10 gene and further characterization of the 5' flanking sequence. Immunogenetics 1997, 46(2):120-128. Eskdale J, Keijsers V, Huizinga T, Gallagher G: Microsatellite alleles and single nucleotide polymorphisms (SNP) combine to form four major haplotype families at the human interleukin-10 (IL-10) locus. Genes Immun 1999, 1(2):151-155. Hurme M, Lahdenpohja N, Santtila S: Gene polymorphisms of interleukins 1 and 10 in infectious and autoimmune diseases. Ann Med 1998, 30(5):469-473. D'Alfonso S, Rampi M, Rolando V, Giordano M, Momigliano-Richiardi P: New polymorphisms in the IL-10 promoter region. Genes Immun 2000, 1(3):231-233. Kube D, Rieth H, Eskdale J, Kremsner PG, Gallagher G: Structural characterisation of the distal 5' flanking region of the human interleukin-10 gene. Genes Immun 2001, 2(4):181-190. Gibson AW, Edberg JC, Wu J, Westendorp RG, Huizinga TW, Kimberly RP: Novel single nucleotide polymorphisms in the distal IL-10 promoter affect IL-10 production and enhance the risk of systemic lupus erythematosus. J Immunol 2001, 166(6):3915-3922. Mormann M, Rieth H, Hua TD, Assohou C, Roupelieva M, Hu SL, Kremsner PG, Luty AJ, Kube D: Mosaics of gene variations in the Interleukin-10 gene promoter affect interleukin10 production depending on the stimulation used. Genes Immun 2004, 5(4):246-255. Rieth H, Assohou AC, Mormann M, Roupelieva M, Kremsner PG, Kube D: A new allelic variation within the 5'-flanking region of the interleukin-10 gene. Eur J Immunogenet 2003, 30(3):191-193. Padyukov L, Hahn-Zoric M, Lau YL, Hanson LA: Different allelic frequencies of several cytokine genes in Hong Kong Chinese and Swedish Caucasians. Genes Immun 2001, 2(5):280-283. Berglundh T, Donati M, Hahn-Zoric M, Hanson LA, Padyukov L: Association of the -1087 IL 10 gene polymorphism with severe chronic periodontitis in Swedish Caucasians. J Clin Periodontol 2003, 30(3):249-254. Crawley E, Kay R, Sillibourne J, Patel P, Hutchinson I, Woo P: Polymorphic haplotypes of the interleukin-10 5' flanking region determine variable interleukin-10 transcription and are associated with particular phenotypes of juvenile rheumatoid arthritis. Arthritis Rheum 1999, 42(6):1101-1108. Koss K, Fanning GC, Welsh KI, Jewell DP: Interleukin-10 gene promoter polymorphism in English and Polish healthy controls. Polymerase chain reaction haplotyping using 3' mismatches in forward and reverse primers. Genes Immun 2000, 1(5):321-324. Pickard C, Mann C, Sinnott P, Boggild M, Hawkins C, Strange RC, Hutchinson IV, Ollier WE, Donn RP: Interleukin-10 (IL10) promoter polymorphisms and multiple sclerosis. J Neuroimmunol 1999, 101(2):207-210. Myhr KM, Vagnes KS, Maroy TH, Aarseth JH, Nyland HI, Vedeler CA: Interleukin-10 promoter polymorphisms in patients with multiple sclerosis. J Neurol Sci 2002, 202(1-2):9397. Maurer M, Kruse N, Giess R, Toyka KV, Rieckmann P: Genetic variation at position -1082 of the interleukin 10 (IL10) promotor and the outcome of multiple sclerosis. J Neuroimmunol 2000, 104(1):98-100. Reich K, Garbe C, Blaschke V, Maurer C, Middel P, Westphal G, Lippert U, Neumann C: Response of psoriasis to interleukin-10 is associated with suppression of cutaneous type 1 inflammation, downregulation of the epidermal interleukin-8/CXCR2 pathway and normalization of keratinocyte maturation. J Invest Dermatol 2001, 116(2):319-329. Craven NM, Jackson CW, Kirby B, Perrey C, Pravica V, Hutchinson IV, Griffiths CE: Cytokine gene polymorphisms in psoriasis. Br J Dermatol 2001, 144(4):849-853.

94

365. 366. 367. 368. 369. 370. 371. 372. 373. 374.

375. 376. 377. 378. 379. 380. 381. 382.

Asadullah K, Eskdale J, Wiese A, Gallagher G, Friedrich M, Sterry W: Interleukin-10 promoter polymorphism in psoriasis. J Invest Dermatol 2001, 116(6):975-978. Hulkkonen J, Pertovaara M, Antonen J, Lahdenpohja N, Pasternack A, Hurme M: Genetic association between interleukin-10 promoter region polymorphisms and primary Sjogren's syndrome. Arthritis Rheum 2001, 44(1):176-179. Eskdale J, McNicholl J, Wordsworth P, Jonas B, Huizinga T, Field M, Gallagher G: Interleukin-10 microsatellite polymorphisms and IL-10 locus alleles in rheumatoid arthritis susceptibility. Lancet 1998, 352(9136):1282-1283. Donger C, Georges JL, Nicaud V, Morrison C, Evans A, Kee F, Arveiler D, Tiret L, Cambien F: New polymorphisms in the interleukin-10 gene--relationships to myocardial infarction. Eur J Clin Invest 2001, 31(1):9-14. Hahn-Zoric M, Hytonen AM, Hanson LA, Nilsson LA, Padyukov L: Association of -1087 IL10 and -308 TNFA gene polymorphisms with serological markers of coeliac disease. J Clin Immunol 2003, 23(4):291-296. D'Alfonso S, Rampi M, Bocchio D, Colombo G, Scorza-Smeraldi R, Momigliano-Richardi P: Systemic lupus erythematosus candidate genes in the Italian population: evidence for a significant association with interleukin-10. Arthritis Rheum 2000, 43(1):120-128. Westendorp RG, Langermans JA, Huizinga TW, Verweij CL, Sturk A: Genetic influence on cytokine production in meningococcal disease. Lancet 1997, 349(9069):1912-1913. Edwards-Smith CJ, Jonsson JR, Purdie DM, Bansal A, Shorthouse C, Powell EE: Interleukin10 promoter polymorphism predicts initial response of chronic hepatitis C to interferon alfa. Hepatology 1999, 30(2):526-530. Helminen M, Lahdenpohja N, Hurme M: Polymorphism of the interleukin-10 gene is associated with susceptibility to Epstein-Barr virus infection. J Infect Dis 1999, 180(2):496499. Santos AR, Suffys PN, Vanderborght PR, Moraes MO, Vieira LM, Cabello PH, Bakker AM, Matos HJ, Huizinga TW, Ottenhoff TH et al: Role of tumor necrosis factor-alpha and interleukin-10 promoter gene polymorphisms in leprosy. J Infect Dis 2002, 186(11):16871691. Wilson JN, Rockett K, Jallow M, Pinder M, Sisay-Joof F, Newport M, Newton J, Kwiatkowski D: Analysis of IL10 haplotypic associations with severe malaria. Genes Immun 2005, 6(6):462-466. Tambur AR, Ortegel JW, Ben-Ari Z, Shabtai E, Klein T, Michowiz R, Tur-Kaspa R, Mor E: Role of cytokine gene polymorphism in hepatitis C recurrence and allograft rejection among liver transplant recipients. Transplantation 2001, 71(10):1475-1480. Lin MT, Storer B, Martin PJ, Tseng LH, Gooley T, Chen PJ, Hansen JA: Relation of an interleukin-10 promoter polymorphism to graft-versus-host disease and survival after hematopoietic-cell transplantation. N Engl J Med 2003, 349(23):2201-2210. Taylor RR, Allen SJ, Greenwood BM, Riley EM: IgG3 antibodies to Plasmodium falciparum merozoite surface protein 2 (MSP2): increasing prevalence with age and association with clinical immunity to malaria. Am J Trop Med Hyg 1998, 58(4):406-413. Wang L, Richie TL, Stowers A, Nhan DH, Coppel RL: Naturally acquired antibody responses to Plasmodium falciparum merozoite surface protein 4 in a population living in an area of endemicity in Vietnam. Infect Immun 2001, 69(7):4390-4397. Yone CL, Kremsner PG, Luty AJ: Immunoglobulin G isotype responses to erythrocyte surface-expressed variant antigens of Plasmodium falciparum predict protection from malaria in African children. Infect Immun 2005, 73(4):2281-2287. McGregor IA, S. Carrington, and S. Cohen. : Treatment of East-African Plasmodium falciparum malaria with West-African human immunoglobulin. . Trans R Soc Trop Med Hyg 1963. , 50:170-. Sabchareon A, Burnouf T, Ouattara D, Attanath P, Bouharoun-Tayoun H, Chantavanich P,

95

383. 384. 385. 386.

387.

388. 389.

390. 391. 392. 393. 394. 395. 396. 397. 398.

Foucault C, Chongsuphajaisiddhi T, Druilhe P: Parasitologic and clinical human response to immunoglobulin administration in falciparum malaria. Am J Trop Med Hyg 1991, 45(3):297-308. Marsh K, Kinyanjui S: Immune effector mechanisms in malaria. Parasite Immunol 2006, 28(1-2):51-60. Good MF, Stanisic D, Xu H, Elliott S, Wykes M: The immunological challenge to developing a vaccine to the blood stages of malaria parasites. Immunol Rev 2004, 201:254-267. Garraud O, Mahanty S, Perraut R: Malaria-specific antibody subclasses in immune individuals: a key source of information for vaccine design. Trends Immunol 2003, 24(1):3035. Aribot G, Rogier C, Sarthou JL, Trape JF, Balde AT, Druilhe P, Roussilhon C: Pattern of immunoglobulin isotype response to Plasmodium falciparum blood-stage antigens in individuals living in a holoendemic area of Senegal (Dielmo, west Africa). Am J Trop Med Hyg 1996, 54(5):449-457. Ndungu FM, Bull PC, Ross A, Lowe BS, Kabiru E, Marsh K: Naturally acquired immunoglobulin (Ig)G subclass antibodies to crude asexual Plasmodium falciparum lysates: evidence for association with protection for IgG1 and disease for IgG2. Parasite Immunol 2002, 24(2):77-82. Dubois B, Deloron P, Astagneau P, Chougnet C, Lepers JP: Isotypic analysis of Plasmodium falciparum-specific antibodies and their relation to protection in Madagascar. Infect Immun 1993, 61(10):4498-4500. Nebie I, Diarra A, Ouedraogo A, Soulama I, Bougouma EC, Tiono AB, Konate AT, Chilengi R, Theisen M, Dodoo D et al: Humoral responses to Plasmodium falciparum blood-stage antigens and association with incidence of clinical malaria in children living in an area of seasonal malaria transmission in Burkina Faso, West Africa. Infect Immun 2008, 76(2):759766. Asadullah K, Sterry W, Volk HD: Interleukin-10 therapy--review of a new approach. Pharmacol Rev 2003, 55(2):241-269. Engwerda C, Belnoue E, Gruner AC, Renia L: Experimental models of cerebral malaria. Curr Top Microbiol Immunol 2005, 297:103-143. Linke A, Kuhn R, Muller W, Honarvar N, Li C, Langhorne J: Plasmodium chabaudi chabaudi: differential susceptibility of gene-targeted mice deficient in IL-10 to an erythrocytic-stage infection. Exp Parasitol 1996, 84(2):253-263. Sanni LA, Jarra W, Li C, Langhorne J: Cerebral edema and cerebral hemorrhages in interleukin-10-deficient mice infected with Plasmodium chabaudi. Infect Immun 2004, 72(5):3054-3058. Wroczynska A, Nahorski W, Bakowska A, Pietkiewicz H: Cytokines and clinical manifestations of malaria in adults with severe and uncomplicated disease. Int Marit Health 2005, 56(1-4):103-114. Awandare GA, Goka B, Boeuf P, Tetteh JK, Kurtzhals JA, Behr C, Akanmori BD: Increased levels of inflammatory mediators in children with severe Plasmodium falciparum malaria with respiratory distress. J Infect Dis 2006, 194(10):1438-1446. Reuss E, Fimmers R, Kruger A, Becker C, Rittner C, Hohler T: Differential regulation of interleukin-10 production by genetic and environmental factors--a twin study. Genes Immun 2002, 3(7):407-413. Tagore A, Gonsalkorale WM, Pravica V, Hajeer AH, McMahon R, Whorwell PJ, Sinnott PJ, Hutchinson IV: Interleukin-10 (IL-10) genotypes in inflammatory bowel disease. Tissue Antigens 1999, 54(4):386-390. Giha HA, Rosthoj S, Dodoo D, Hviid L, Satti GM, Scheike T, Arnot DE, Theander TG: The epidemiology of febrile malaria episodes in an area of unstable and seasonal transmission. Trans R Soc Trop Med Hyg 2000, 94(6):645-651.

96

399. 400. 401. 402.

403. 404. 405. 406. 407. 408. 409. 410. 411. 412.

413. 414. 415. 416.

Theander TG: Unstable malaria in Sudan: the influence of the dry season. Malaria in areas of unstable and seasonal transmission. Lessons from Daraweesh. Trans R Soc Trop Med Hyg 1998, 92(6):589-592. Hamad AA, Nugud Ael H, Arnot DE, Giha HA, Abdel-Muhsin AM, Satti GM, Theander TG, Creasey AM, Babiker HA, Elnaiem DE: A marked seasonality of malaria transmission in two rural sites in eastern Sudan. Acta Trop 2002, 83(1):71-82. Sakihama N, Mitamura T, Kaneko A, Horii T, Tanabe K: Long PCR amplification of Plasmodium falciparum DNA extracted from filter paper blots. Exp Parasitol 2001, 97(1):50-54. Hoffmann KF, McCarty TC, Segal DH, Chiaramonte M, Hesse M, Davis EM, Cheever AW, Meltzer PS, Morse HC, 3rd, Wynn TA: Disease fingerprinting with cDNA microarrays reveals distinct gene expression profiles in lethal type 1 and type 2 cytokine-mediated inflammatory reactions. Faseb J 2001, 15(13):2545-2547. Wooden J, Gould EE, Paull AT, Sibley CH: Plasmodium falciparum: a simple polymerase chain reaction method for differentiating strains. Exp Parasitol 1992, 75(2):207-212. Snounou G, Zhu X, Siripoon N, Jarra W, Thaithong S, Brown KN, Viriyakosol S: Biased distribution of msp1 and msp2 allelic variants in Plasmodium falciparum populations in Thailand. Trans R Soc Trop Med Hyg 1999, 93(4):369-374. Smythe JA, Peterson MG, Coppel RL, Saul AJ, Kemp DJ, Anders RF: Structural diversity in the 45-kilodalton merozoite surface antigen of Plasmodium falciparum. Mol Biochem Parasitol 1990, 39(2):227-234. Trager W, Jensen JB: Human malaria parasites in continuous culture. Science 1976, 193(4254):673-675. Troye-Blomberg M, Perlmann H, Patarroyo ME, Perlmann P: Regulation of the immune response in Plasmodium falciparum malaria. II. Antigen specific proliferative responses in vitro. Clin Exp Immunol 1983, 53(2):345-353. Bradford MM: A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 1976, 72:248-254. Lewis CM: Genetic association studies: design, analysis and interpretation. Brief Bioinform 2002, 3(2):146-153. Nguer CM, Diallo TO, Diouf A, Tall A, Dieye A, Perraut R, Garraud O: Plasmodium falciparum- and merozoite surface protein 1-specific antibody isotype balance in immune Senegalese adults. Infect Immun 1997, 65(11):4873-4876. Bouharoun-Tayoun H, Druilhe P: Plasmodium falciparum malaria: evidence for an isotype imbalance which may be responsible for delayed acquisition of protective immunity. Infect Immun 1992, 60(4):1473-1481. Smith JD, Chitnis CE, Craig AG, Roberts DJ, Hudson-Taylor DE, Peterson DS, Pinches R, Newbold CI, Miller LH: Switches in expression of Plasmodium falciparum var genes correlate with changes in antigenic and cytoadherent phenotypes of infected erythrocytes. Cell 1995, 82(1):101-110. Tomee JF, Dubois AE, Koeter GH, Beaumont F, van der Werf TS, Kauffman HF: Specific IgG4 responses during chronic and transient antigen exposure in aspergillosis. Am J Respir Crit Care Med 1996, 153(6 Pt 1):1952-1957. Aalberse RC, van der Gaag R, van Leeuwen J: Serologic aspects of IgG4 antibodies. I. Prolonged immunization results in an IgG4-restricted response. J Immunol 1983, 130(2):722-726. Gascan H, Gauchat JF, Aversa G, Van Vlasselaer P, de Vries JE: Anti-CD40 monoclonal antibodies or CD4+ T cell clones and IL-4 induce IgG4 and IgE switching in purified human B cells via different signaling pathways. J Immunol 1991, 147(1):8-13. Lundgren M, Persson U, Larsson P, Magnusson C, Smith CI, Hammarstrom L, Severinson E: Interleukin 4 induces synthesis of IgE and IgG4 in human B cells. Eur J Immunol 1989,

97

417.

418.

419.

420.

421. 422. 423. 424. 425.

426. 427.

19(7):1311-1315. Nasr A, Iriemenam NC, Troye-Blomberg M, Giha HA, Balogun HA, Osman OF, Montgomery SM, ElGhazali G, Berzins K: Fc gamma receptor IIa (CD32) polymorphism and antibody responses to asexual blood-stage antigens of Plasmodium falciparum malaria in Sudanese patients. Scand J Immunol 2007, 66(1):87-96. Marchioni E, Marinou-Aktipi K, Uggetti C, Bottanelli M, Pichiecchio A, Soragna D, PiccoloG, Imbesi F, Romani A, Ceroni M: Effectiveness of intravenous immunoglobulin treatment in adult patients with steroid-resistant monophasic or recurrent acute disseminated encephalomyelitis. J Neurol 2002, 249(1):100-104. van der Neut Kolfschoten M, Schuurman J, Losen M, Bleeker WK, Martinez-Martinez P, Vermeulen E, den Bleker TH, Wiegman L, Vink T, Aarden LA et al: Anti-inflammatory activity of human IgG4 antibodies by dynamic Fab arm exchange. Science 2007, 317(5844):1554-1557. Lyke KE, Dicko A, Dabo A, Sangare L, Kone A, Coulibaly D, Guindo A, Traore K, Daou M, Diarra I et al: Association of Schistosoma haematobium infection with protection against acute Plasmodium falciparum malaria in Malian children. Am J Trop Med Hyg 2005, 73(6):1124-1130. Calandra T, Echtenacher B, Roy DL, Pugin J, Metz CN, Hultner L, Heumann D, Mannel D, Bucala R, Glauser MP: Protection from septic shock by neutralization of macrophage migration inhibitory factor. Nat Med 2000, 6(2):164-170. van der Zee JS, van Swieten P, Aalberse RC: Inhibition of complement activation by IgG4 antibodies. Clin Exp Immunol 1986, 64(2):415-422. Kawa S, Kitahara K, Hamano H, Ozaki Y, Arakura N, Yoshizawa K, Umemura T, Ota M, Mizoguchi S, Shimozuru Y et al: A novel immunoglobulin-immunoglobulin interaction in autoimmunity. PLoS ONE 2008, 3(2):e1637. Li C, Corraliza I, Langhorne J: A defect in interleukin-10 leads to enhanced malarial disease in Plasmodium chabaudi chabaudi infection in mice. Infect Immun 1999, 67(9):4435-4442. Mordmuller BG, Metzger WG, Juillard P, Brinkman BM, Verweij CL, Grau GE, Kremsner PG: Tumor necrosis factor in Plasmodium falciparum malaria: high plasma level is associated with fever, but high production capacity is associated with rapid fever clearance. Eur Cytokine Netw 1997, 8(1):29-35. Suarez A, Castro P, Alonso R, Mozo L, Gutierrez C: Interindividual variations in constitutive interleukin-10 messenger RNA and protein levels and their association with genetic polymorphisms. Transplantation 2003, 75(5):711-717. Galley HF, Lowe PR, Carmichael RL, Webster NR: Genotype and interleukin-10 responses after cardiopulmonary bypass. Br J Anaesth 2003, 91(3):424-426.

98

Das könnte Ihnen auch gefallen