Download Host susceptibility to malaria in human and mice: compatible

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Duffy antigen system wikipedia , lookup

Neonatal infection wikipedia , lookup

Plant disease resistance wikipedia , lookup

Immunomics wikipedia , lookup

Globalization and disease wikipedia , lookup

Plasmodium falciparum wikipedia , lookup

Transcript
Articles in PresS. Physiol Genomics (October 29, 2013). doi:10.1152/physiolgenomics.00044.2013
1
Host susceptibility to malaria in human and mice: compatible approaches to
2
identify potential resistant genes
3
4
5
Maria Hernandez-Valladares1, Pascal Rihet2,3 and Fuad A. Iraqi4*
6
7
1
Biomedical Research/Public Health Operations, Granada, Spain
8
2
UMR1090 TAGC, INSERM,, 163 av, Luminy, 13288 Marseille, CEDEX 09, France
9
3
Aix-Marseille University, Marseille, F-13288, France
10
4
Department of Clinical Microbiology and Immunology, Sackler Faculty of Medicine, Tel Aviv
11
University, Ramat Aviv, 69978 Tel Aviv, Israel
12
13
Running title: Genetic resistance to Malaria in human and murine
14
15
*
Correspondence to: Prof. Fuad A. Iraqi
16
Department of Clinical Microbiology and Immunology
17
Sackler Faculty of Medicine
18
Tel-Aviv University
19
Ramat Aviv, Tel-Aviv 69978
20
Israel
21
Phone: +972-3-6409321
22
Fax: +972-3-6409160
23
[email protected]
24
25
26
All the authors contributed equally to the preparation of this review
27
1
Copyright © 2013 by the American Physiological Society.
28
Abstract
29
There is a growing evidence for human genetic factors controlling the outcome of malaria infection,
30
while molecular basis of this genetic control still poorly understood. Case-control and Family-based
31
studies have been carried to identify genes underlying host susceptibility to malarial infection.
32
Parasitemia and mild malaria have been genetically linked to human chromosomes 5q31-q33 and
33
6p21.3, and several immune genes located within those regions have been associated with malaria
34
related phenotypes. Association and linkage studies of resistance to malaria are not easy to carryout in
35
human populations, due to the difficulty to survey a significant number of families. Murine models
36
were proven to be excellent genetic tool for studying host response to malaria, that allowed mapping
37
fourteen resistance loci, eight of them controlling parasitic levels and six controlling cerebral malaria.
38
Once QTL or genes have been identified then may human ortholoug can be identified. Comparative
39
mapping studies showed that couple of human and mouse might share similar genetic-controlled
40
mechanisms of resistance. In this way, char8 that controls parasitemia was mapped on chromosome 11;
41
char8 corresponds to human chromosome 5q31-q33, and contains immune genes, such as Il3, Il4, Il5,
42
Il12b, Il13, Irf1, and Csf2. Nevertheless, part of the genetic factors controlling malaria traits might
43
differ in both hosts due to specific host-pathogen interactions. Finally, novel genetic tools including
44
animal model were recently developed and will offer new opportunities for identifying genetic factors
45
underlying host phenotypic response to malaria, which will help in better therapeutic strategies
46
including vaccine and drug development.
47
48
49
50
51
52
2
53
Introduction
54
Plasmodium falciparum malaria remains a major cause of morbidity and mortality in many developing
55
countries. According to WHO, the number of cases in 2012 reached 219 millions (95% CI 154-289
56
million), among which over 660 000 (95% CI 490 000-836 000) were fatal (62, 178). Drug-resistant
57
parasite strains are spreading rapidly across the world and, in spite of substantial efforts and
58
encouraging results from recent trials, a fully protective malaria vaccine remains a distant prospect
59
(154). Vaccine development faces major difficulties partly due to the genetic control of immunity to the
60
parasite (155, 156, 163, 176) and/or to the parasitic antigenic variation (146).
61
The outcome of human malaria infection is thought to depend on both parasite and host genetic factors.
62
P. falciparum genotypes have been associated with differences in disease outcome (41) and may be
63
subject to selective pressure (34, 55). The influence of host genetic factors has been demonstrated in
64
animal models (156, 47, 49), and there is accumulating evidence of human genetic control in malarial
65
infection and disease. The development of efficient control measures calls for a better knowledge of
66
human genes controlling the infection and the disease. Genetic epidemiology studies are aimed at
67
estimating the heritability of phenotypes related to malaria, at localizing and at identifying genes
68
involved in susceptibility or resistance to malaria. Herein we will review the current knowledge in the
69
field of genetics of human malaria. We will particularly discuss the results of case-control studies and
70
family-based and the animal model studies.
71
Several malaria phenotypes have been considered in genetic epidemiology studies and summarized in
72
Table 1. The purpose of such studies is to identify genetic factors that control parasitemia, and those
73
that determine the risk of developing either mild malaria or severe malaria. Although the phenotypes
74
are related, the study of each phenotype should give a particular insight into the understanding of
75
malaria physiopathology.
76
Severe malaria is mainly composed of three sub-phenotypes: cerebral malaria, severe anemia, and
77
respiratory distress (102). Separate or combined phenotypes have been considered in association
3
78
studies. The studies of severe malaria phenotypes are based on recruitment in hospital. For mild
79
malaria, composed of parasitemia and fever attacks sub-phenotypes, the ascertainment is based on
80
population living in an endemic area. A binary phenotype and quantitative phenotypes can be used for
81
both severe and mild malaria: the presence or absence of malarial disease is a binary phenotype, while
82
the risk of developing malarial disease is a quantitative phenotype. The calculation of the quantitative
83
phenotype is generally based on a logistic regression model that takes into account significant
84
covariates, such as age.
85
Asexual blood stages are responsible for mild and severe malaria. Parasitemia without clinical
86
symptoms frequently occur in endemic area, but high parasitemia strongly increased the risk of
87
developing mild malaria attacks. This has been clearly demonstrated in longitudinal studies in Africa
88
(147). High parasitemia is often observed in case of severe malaria, and is a WHO criteria for severe
89
malaria diagnosis (182, 62, 113). However, in case of severe malaria, infected red blood cells are
90
sequestered at high levels, and low parasitemia can be observed (137, 138). In addition, it is
91
conceivable that two individuals with the same parasitemia differently respond to the parasite infection
92
mounting different pro-inflammatory responses, and whereas one could develop severe malaria, the
93
other could develop only mild malaria. Nevertheless, malaria pathogenesis is incompletely understood
94
and the study of parasitemia, which is a risk factor, should be helpful in understanding mechanisms
95
involved in clinical malaria. Since parasitemia fluctuate considerably, only one measurement will not
96
reflect the parasitic load of individuals. Therefore, cross-sectional studies (studies are based on data
97
collected by measuring a parameter in many subjects at the same time point, or without regarding
98
differences in time. Thus, a cross-sectional parasitemia survey is based on one parasitemia
99
measurement), that are designed to test association of polymorphisms with severe malaria are generally
100
not appropriate to the study of parasitemia. Conversely, longitudinal studies (in this study, subjects are
101
surveyed over time with continuous or repeating monitoring of a parameter. Thus, a longitudinal
102
parasitemia survey looks at parasitemia measured several times in the same subjects, examines
4
103
how
parasitemia
changes
over
the
course
of
time,
and
how
subjects
differ
in
104
parasitemia), that take into account more than ten parasitemia measurements will be more appropriate
105
for studying the influence of host genetic factors on parasitemia.
106
Additional phenotypes have been used in genetic studies. Since immunoglobulins G (IgG) have been
107
shown to inhibit the growth of parasites in cooperation with effector cells in vitro and to protect
108
humans in vivo (24), IgG levels were considered as pertinent phenotypes. The levels of IgG-subclasses
109
were also assessed: the levels of cytophilic IgG that have been associated with protection from mild or
110
severe malaria are the most interesting phenotypes (12, 13). IgG1 and IgG3 that bind to Fc receptors,
111
and IgG2 that binds to FcRIIA-H131, which is frequent in Africa, likely kill the parasite in cooperation
112
with cells such as monocytes, while IgG4 may block this antibody-dependent cell cytotoxicity
113
mechanism (Figure 1).
114
115
The heritability of phenotypes related to malarial infection and disease
116
Comparisons of three sympatric West African ethnic groups showed that the Fulani had lower blood
117
infection levels and fewer clinical malaria attacks than the Mossi and the Rimaibe living in the same
118
area in Burkina Faso (110). In addition, the Fulani had higher levels of IgG against malaria antigens
119
(111). Similar ethnic differences in resistance to malaria have been observed in Mali (39) and in Nepal
120
(160). Interestingly, the low frequency of protective globin variants in the Fulani group suggested that
121
unknown genetic factors were involved in malaria resistance (112). Although such differences may be
122
a result of differences in lifestyle, it is thought that the Fulani resistance to malaria had a genetic basis.
123
Furthermore, twin studies have shown that mild malaria and antibody responses to parasite antigens
124
were more concordant within monozygotic twin pairs than in dizygotic pairs (72, 73, 152). The results
125
of several family studies have shown conclusively a familial aggregation of blood infection levels (1,
126
52, 142), malaria attacks (72), severe malaria (141), and anti-malarial immune responses (12, 155,
127
158). Segregation analyses of parasitemia yielded significant sibling correlation coefficients, which
5
128
ranged from 0.16 to 0.24 (1, 52, 142). Similarly, several studies have found significant sibling
129
correlation of IgG and IgG sub-class responses to parasite antigens (12, 157, 158); sibling correlation
130
coefficients ranged from 0.13 to 0.39. Although an initial study detected a major gene controlling
131
parasitemia in Cameroun, segregation analyses of both parasitemia and antibody levels rather
132
evidenced a complex genetic model in Cameroun, in Burkina Faso, and in Papua New Guinea (1, 52,
133
142, 157). The odds of having cerebral malaria were 2.49 for a child whose sibling had a history of
134
cerebral malaria than the odds for a child whose sibling had no history of cerebral malaria (34, 141).
135
Similar results have been obtained for mild malaria and severe anemia (72, 141). In addition,
136
Mackinnon et al. (97, 98) found that genetic factors explained approximately 25% of the variation in
137
the incidence of mild malaria episodes. Interestingly, the studies attributed a small proportion of the
138
variation of mild malaria incidence to hemoglobin S (HbS) and α-thalassemia, and pointed out that
139
other genes were involved in malaria resistance (97, 98).
140
Case-control association studies have mainly identified genes associated with severe malaria
141
Case-control studies have been carried out to investigate the role of candidate genes in clinical malaria
142
and these studies are summarized in Table 1. Such studies test whether case and control groups differ
143
in marker allelic frequencies. Case-control studies are based on candidate genes, which have been
144
chosen on the basis of the high frequency of deleterious alleles at the homozygous stage, or their
145
function. Fifty years ago, Haldane and Allison reported a similarity in geographical distribution of
146
hemoglobinopathologies (thalassemia and sickle disease) and P. falciparum malaria, and proposed that
147
thalassemia and HbS might confer protection against malaria (10, 59). Thus, genes encoding alpha-
148
globin and beta-globin have been considered as candidate genes. Other genes involved in erythrocyte
149
surface, in immune responses that likely affect the outcome of infectious diseases, in the sequestration
150
of P falciparum-infected erythrocytes or in erythrocyte metabolism have also been analyzed in
151
association studies.
6
152
Erythrocyte polymorphisms and severe malaria
153
HbS carriers are strongly protected against severe malaria (2, 4, 8, 67). This clearly supports that there
154
is a significant impact of malaria in selecting HbS as a protective allele, in spite of the deleterious effect
155
of HbS at the homozygous state. Another hemoglobin variant named HbC was shown to be associated
156
with protection against severe malaria (3, 107, 112). Strikingly, the HbC and the HbS mutations occur
157
in the 6th position in the beta-globin gene, and both mutations coexist in several populations in West
158
Africa. Since HbC is clinically less severe than HbS in the homozygote state, it has been suggested that
159
HbC would replace HbS in West Africa (161). Other beta-globin and alpha-globin variants that cause
160
the thalassemias have also been shown to protect against severe malaria (7, 108, 179). In addition,
161
association studies have provided consistent evidence of a protective effect of glucose-6-phosphate
162
dehydrogenase (G6PD) deficiencies against severe malaria (56, 149) and haplotype analyses support
163
the hypothesis of recent positive selection (150, 164). Other red blood cell polymorphisms, such as the
164
deletion of glycophorin C (GYPC) exon 3 or a 27 bp-deletion in band 3 gene (SLC4A1) have also been
165
associated with protection from severe malaria (9, 54, 133, 134).
166
Cytoadherence genes, immune genes and severe malaria
167
Several studies have detected the association of CD36 antigen (CD36), complement receptor 1 (CR1),
168
intercellular adhesion molecule 1 (ICAM1) and platelet/endothelial cell adhesion molecule 1 (PECAM1)
169
variants with either resistance or susceptibility to severe malaria. However, the results have been highly
170
variable in different geographical locations. For instance, ICAM1-Kilifi allele was found to be
171
associated with protection from severe malaria in Gabon and with susceptibility in Kenya, while a large
172
study in the Gambia failed to detect any association (18, 42, 81). In the same way, the CD36-1264G
173
variant was associated with susceptibility to cerebral malaria in populations living in the Gambia and in
174
Kenya (6), and with protection against severe malaria in another Kenyan population (132). Similar
175
findings were obtained for the other candidate cytoadherence genes and variants. It has been suggested
176
that parasite factors may profoundly influence the pattern of association, since the sequestration of the
7
177
parasite depends on the constant switch between different forms of P. falciparum erythrocyte
178
membrane protein 1 (PfEMP-1) (146).
179
The nitric oxide synthase type 2A (NOS2A)-954C variant has been associated with protection against
180
severe malaria in Gabon, while no association was detected in the Gambia and in Tanzania (26, 80, 90).
181
Similarly, the human leukocyte antigen HLA-B53 was associated with protection against severe
182
malaria, but not in Kenya (67, 68). In contrast, several tumor necrosis factor (TNF) alleles have been
183
associated with severe malaria in different studies: these include TNF-308A and TNF-238A (5, 104,
184
105, 175). The Fc gamma receptor IIa (FCGR2A)-H131 variant was also found to be associated with
185
susceptibility to severe malaria in two different studies in Thailand and the Gambia (35, 129).
186
The association of erythrocyte genetic variants with severe malaria seems to be well established, while
187
the pattern of associations of cytoadherence and immune responses with severe malaria is less clear and
188
remains indicative of ongoing research. Strikingly, most of erythrocyte genetic variants that protect
189
against severe malaria have been associated also with protection against mild malaria, erythrocyte
190
resistance to invasion by the parasite or a reduced parasitemia, while a few genes involved in immune
191
responses, such as TNF and NOS2A, have been found to be associated with mild malaria or parasitemia
192
(Table 1). The absence of association with parasitemia may be partly due to the cross-sectional study
193
design, which is less appropriate than the longitudinal study design. Candidate immune genes
194
associated with severe malaria may also weakly influence parasitemia and/or mild malaria, and other
195
immune genes may be more important for these phenotypes. Nevertheless, it remains possible that the
196
association of some genes with severe malaria arise from statistical artifacts, such as biases due to
197
population stratification or population admixture.
198
Family-based association and linkage studies test whether a particular allele of the marker is
199
transmitted from parents to affected offspring in families more often than expected by chance. To date,
200
most of the candidate genes have not been analyzed by using family-based studies. Nevertheless, such
201
studies recently confirmed the association of HbAS and interferon gamma receptor 1 (IFNGR1)+2200C
8
202
with protection against severe malaria (2, 77) and revealed the association of IFNG-183T with reduced
203
risk of cerebral malaria (76, 29). In addition, HbC that protects from severe malaria has also been
204
shown to protect also against mild malaria (144).
205
Linkage analyses do not focus on a particular allele, and alleles co-inherited with disease may vary
206
from one family to another), led to the localization of genes controlling parasitemia and mild malaria
207
on chromosome 5q31-q33 (44, 53, 143) and chromosome 6p21-p23 (45, 74). Chromosome 5q31-q33
208
contains several candidate genes implicated in the regulation of immune responses (Figure 1). These
209
include genes involved in the Th1 and Th2 balance, such as interleukin 4 (IL4) and IL12B, and genes
210
involved in the activation of effector cells, such as IL3, IL5, colony stimulating factor 2 (CSF2), and
211
colony stimulating factor 1 receptor (CSF1R), and genes involved in the production of antibodies, such
212
as IL4 and IL13. Chromosome 6p21-p23 corresponds to the MHC region, and the peak of linkage was
213
found to be close to TNF (45). A systematic polymorphism-screening revealed a family-based
214
association of TNF-308, TNF-238, TNF-851 and TNF-1304 polymorphisms with mild malaria or
215
parasitemia (46).Nevertheless, TNF genetic variation did not completely explain the linkage of the
216
MHC region with mild malaria. More recently, results have shows that the natural cytotoxicity
217
triggering receptor 3 (NCR3) that is close to TNF, and we showed that a NCR3 polymorphism located
218
within the promoter was associated with mild malaria, and that this association was not due to the
219
association of TNF with mild malaria (36).
220
Interestingly, recent studies were shown that mouse resistance to P. chabaudi linked to mouse
221
chromosome 11 and mouse chromosome 17 (63, 65), which carry regions homologous to chromosome
222
5q31-q33 and the MHC region, respectively. These findings might help to dissect the human loci and to
223
demonstrate the effect of candidate genes in vivo. So far, only the effect of Tnf on resistance to P.
224
chabaudi has been demonstrated by using Tnf-deficient mice (66), which were more susceptible to the
225
infection with higher parasitic levels and mortality rates.
226
Current lessons from human genetic studies of resistance to malaria
9
227
Case-control studies have clearly shown the protective effect of red blood cell polymorphisms on
228
malaria (14, 40, 43, 89, 123, 140). A number of case-control studies have also detected association of
229
genes involved in cytoadherence or immune responses. However, most of these studies yielded
230
conflicting results, suggesting a possible role for hitherto unknown traits associated but not examined
231
in these studies. Family-based studies that avoid biases due to population stratification (refers to a
232
situation in which the study population includes two sub-populations genetically different. In this case,
233
the two sub-populations differ in allele frequencies at a number of genetic markers located within
234
different chromosomal regions (20, 25, 69, 75, 76, 82, 95, 96, 105, 108, 124, 125, 130, 165, 180).
235
These differences can be due to different population histories), should be useful to clarify this point. To
236
date, very few family-based association studies have been conducted, and they have confirmed the
237
protective effect of hemoglobin variant and some immune gene variants.
238
Linkage and association analyses can be viewed as a more general approach to localize and to identify
239
genes involved in malaria resistance (15, 23, 78, 114, 115, 151, 169). The first studies have identified
240
two chromosomal regions that contain malaria resistance genes (Figure 1), and these results were
241
confirmed in linkage replication studies in humans. Two genes, TNF and NCR3, which are located on
242
chromosome 6p21.3, have been shown to be independently associated with mild malaria, indicating
243
that there are at least two genes located on the central region of MHC involved in the genetic control of
244
human malaria. Preliminary analysis suggests that chromosome 5q31-q33 also contains several genes
245
controlling parasitemia (P. Rihet, unpublished data). To date, those malaria resistance genes have not
246
been identified. It is well known that the identification of genes underlying complex traits is a
247
challenging task, and strategies combining mouse and human studies may be helpful in identifying
248
malaria resistance genes. In this way, two mouse loci have been mapped to chromosomal regions that
249
correspond to human chromosomes 5q31-q33 and 6p21.3 (see the section of mouse genetic studies of
250
malaria resistance in this chapter). This suggests that mouse models might help to identify malaria
251
resistance genes in humans.
10
252
Mouse Models for Studying Malaria
253
The mouse is a leading model for studying biological processes in mammals and provides models of
254
particular importance in the study of the host response to infectious disease and understanding why
255
some naturally occurring genotypes are dramatically more sensitive to disease than others. This clearly
256
applies to the malaria disease. Human and mouse genomes are very similar. Mouse strains show
257
different degree of susceptibility and resistance to malaria as observed in human cases. There are a
258
number of different parasites infecting the laboratory mouse and while none of these behaves like P.
259
falciparum, the deadly strain of human malaria, they do model parts of the disease. There are also
260
significant inter-strain variations, which have been exploited to study host responses. Therefore, the
261
mouse model offers a powerful tool for dissecting and understanding the host response mechanism to
262
malaria, and through comparative genomics it may be possible to identify the orthologous genes in
263
humans.
264
Mouse populations for gene mapping studies
265
In crosses between genetically well-defined strains of mice, chromosomal regions responsible for the
266
genetic variance of quantitative traits (i.e. parasitemia) can be mapped to a large genetic interval as
267
quantitative trait loci (QTL) in experimental resource populations (F2 or backcross –BC- generation)
268
(87, 135). A QTL is defined as a region of DNA that is associated with a particular phenotypic trait.
269
Knowing the number of QTL that explains variation in the phenotypic trait tells us about the genetic
270
complexity of a trait. QTL mapping is the statistical study of the alleles which occur in a locus and the
271
phenotypes that they produce. Due to limit of recombination events in F2 and BC populations, QTL
272
usually maps within 20-30cM genomic regions, which may consist hundreds if not thousands of genes.
273
1 cM represents approximately 2000 Kb of DNA in the mouse genome.
274
Subsequently, fine mapping of the QTL can be achieved by a number of approaches including
275
advanced intercross lines (AIL), recombinant inbred lines (RIL), congenic strains (CS), using single
276
nucleotide polymorphisms (SNPs) (181) haplotype mapping (58), and Collaborative Cross (CC) was
11
277
designed (33, 162, 166). They are unquestionably providing an opportunity to identify host genetic
278
factors that influence disease progression.
279
Once QTL or genes have been identified, genetic analysis may then be extended successfully to
280
humans (167). For example, the genetic basis of total and HDL cholesterol concentrations was
281
examined in experimental crosses involving various pairs of inbred mouse strains maintained on
282
regular chow and high-fat diet (71, 79, 99, 171). Although these experiments discovered more than 27
283
QTL, they could explain only a fraction of the total genetic variance in cholesterol metabolism.
284
Importantly, however, comparative mapping showed that the majority of murine QTL and candidate
285
genes have known correspondents in the human genome emphasizing the relevance of QTL analysis in
286
the mouse model for understanding complex disease in humans (172).
287
Murine malaria parasites
288
Since more than 20 years, laboratory mice have been exposed to a wide range of health-threatening
289
conditions, infected with all kind of microorganisms, phenotyped, cured with testing drugs and
290
immunized with potential vaccines. The availability of mouse genetic markers (e.g. microsatellites and
291
single nucleotide polymorphisms), new mouse genomes resources (see Box 1) and the published initial
292
sequencing of the mouse C57BL/6J genome by the Mouse Genome Sequencing Consortium (117) have
293
highly encouraged the study of the genetic bases of resistance to diseases in mouse. Four mouse
294
Plasmodium species (P. chabaudi, P. berghei, P. yoelii and P. vinckei) were found to infect rodents and
295
mosquitoes from several Central African regions several decades ago (31, 32, 84, 85). Thereafter,
296
numerous studies described the morphological and developmental characteristics and iso-enzyme
297
patterns from the different species (86, 170). The malaria mouse models became popular laboratory
298
tools to study the drug (mainly pyrimethamine, cloroquine and sulphonamide) resistance issues that
299
were alarmingly worrying in the affected human communities (131, 153, 174, 184).
300
Among the four species, P. chabaudi has provided a remarkable experimental tool with many
301
similarities to human P. falciparum, such as analogous blood-stage antigens, invasion of red blood
12
302
cells, suppression of cell-mediated immune responses and parasite sequestration in target organs (159).
303
Two subspecies of P. chabaudi have been described: P. chabaudi chabaudi and P. chabaudi adami.
304
The latter is known to induce a severe form of the disease characterized with more pronounced
305
pathologies when compared to the infections by P. chabaudi. P. yoelii was a common experiment
306
model of mouse cerebral malaria some decades ago. However, it has been extensively used in the
307
development and characterization of vaccine candidates in the last decade (119). Three subspecies have
308
been defined: P. yoelii yoelii, P. yoelii killicki and P. yoelii nigeriensis. P. berghei is the preferred
309
mouse parasite to study cerebral malaria and its lethal pathologies currently (93). Researchers have
310
available five different isolates: k173, SP11, LUKA, NK65 and ANKA, although most of the mouse
311
cerebral malaria investigations have been carried out with P. berghei ANKA. Four subspecies of P.
312
vinckei are recognized: P. vinckei vinckei, P. vinckei petteri, P. vinckei lentum and P. vinckei
313
brucechwatti. Despite being the most distributed mouse parasite in nature, P. vinckei is the least
314
characterized mouse parasite. P. vinckei petteri has been used in the identification of new antimalarial
315
drugs (21).
316
Common laboratory mouse strains used in malaria studies
317
In the last few decades, up to thirteen different inbred mouse strains have been infected with several
318
Plasmodium species and their susceptibility to malaria (and to other infectious diseases) has been
319
characterized (83, 126, 136). Most of the mouse strains are susceptible to the infections by the four
320
murine parasites. However, C57BL/6J and DBA/2J mice are resistant to infections by P. chabaudi and
321
P. berghei parasites, respectively, and are commonly used to investigate the molecular basis of mouse
322
resistance to malaria. Susceptible and resistant mice to P. chabaudi infections suffer from severe
323
anemia after two weeks post infection. Susceptible animals die with high parasitic levels and low red
324
blood cell concentrations whereas resistant animals clear parasites from the bloodstream and recover
325
the pre-infection red blood cell values. Moreover, the outcomes from susceptible and resistant mice
326
after infection with mouse parasites can differ substantially. While P. berghei-susceptible mice die
13
327
from cerebral malaria after few days post infection, resistant animals die from severe anemia after
328
several weeks post infection. Table 2 summarizes the current information about the susceptibility of
329
laboratory mouse to malaria induced by the most common murine parasites in mouse studies, P.
330
chabaudi, P. berghei and P. yoelii. Since 1997, a remarkable number of publications have described
331
more than a dozen genetic loci linked to murine malaria resistance (47, 48, 92, 100).
332
Genetic loci linked to murine malaria parasitemia
333
One of the first published studies crossed two mouse strains, C3H/He and SJL, susceptible to P.
334
chabaudi adami DS-induced malaria with a common resistant strain, C57BL/6J, to produce a F2
335
generation of animals (47). Two loci on chromosome 9 (chabaudi resistance char1) and chromosome 8
336
(char2) linked to the parasite-induced death were found in both crosses. Char1 was also found in F2
337
animals from both crosses when testing for QTL contributing to peak parasitemia whereas char2 was
338
only found in animals originated from the C3H/HexC57BL/6J cross. Although the 95% confidence
339
interval of their positions spanned a wide genomic region comprised of several hundreds of genes,
340
some candidate genes were suggested to control the phenotypic traits: haptoglobin (Hp), encoding a
341
protein that functions in the binding of free hemoglobin in the bloodstream, and erythrocyte antigen 1
342
(EaI), encoding for the protein that define MN blood groups in humans, on chromosome 8; transferrin
343
(Trf), encoding a key iron transport protein, and two retinol-binding proteins (Rbp1 and Rbp2), on
344
chromosome 9.
345
Simultaneously, char2 was also mapped in a separate study using a different susceptible mouse strain,
346
A/J (49). Despite the width of the 95% confidence interval, three new genes were presented as
347
attractive candidates for char2: glycophorin A (GypA), encoding for an erythrocyte membrane protein,
348
interleukin 15 (Il15) and macrophage scavenger receptor 1 (Msr1), encoding for key proteins involved
349
in the phagocytosis of microbes by mononuclear phagocytes. Thus, char2 was a common locus to
350
control parasitic levels in two independent studies, which used different susceptible mouse strains and
351
mouse parasite subspecies to induce the infection.
14
352
Later on, an independent BC study with susceptible NC/Jic and resistant 129/SvJ mice using P. yoelii
353
17XL (148) mapped a locus on chromosome 9, pymr (Plasmodium yoelii malaria resistance),
354
controlling survival and parasitemia after infection at the same position as char1 localized in a previous
355
study using P. chabaudi adami DS (47). Since P. yoelii prefers to invade reticulocytes whereas P.
356
chabaudi invades both reticulocytes and mature red blood cells, pymr (char1) might represent a
357
relevant common locus controlling malaria traits induced by different species. The locus position is
358
homologous to the human region 3q24-26 that contains two genes involved in immune responses,
359
cytokine-inducible SH-2-containing protein (Cish) and chemokine receptor 4 (Cmkbr4).
360
Further studies with the promising loci, char1 and char2, could aim the identification of the key gene.
361
In fact, five years later, a CS approach was used to generate char2-CS using C3H/HeJ and C57BL/6J
362
as parental inbred lines (28). At each generation, mice carrying the genotype of interest at char2 locus
363
were selected for subsequent backcrosses to the parental mice to fix the background genome. A total of
364
ten char2-CS, each carrying different recombination outputs around char2, were originated from
365
intercrosses of lines at the N7 or N8 generation. Genetic analysis of P. chabaudi adami DS-infected
366
animals from each line suggested a large congenic interval on chromosome 8, which conferred
367
susceptibility on background of resistant parental mice (the CS was named Char2-7). The interval was
368
most likely comprised of several genes that could interact to control survival and parasitic levels. One
369
of them was mapped to the distal region of the chromosome when compared congenic intervals from
370
non-susceptible animals of resistant background. Because char2 was localized in the central region on
371
chromosome 8 in previous studies (47, 49), these congenic studies strongly supported the existence of
372
several genes at this locus, sufficient to decrease the survival to malaria. In order to refine the map
373
positions of the char2 genes, a series of recombinant lines from the Char2-7 CS has been recently
374
derived (91) by crossing back to the parental line C57BL/6J. Nine Char2-7 recombinant CS, targeting
375
the distal region of chromosome 8, were inbred, infected with P. chabaudi adami DS and tested for
376
susceptibility to malaria. Due to changes of the parasite lethality and the difficulties to discriminate
15
377
phenotypes among the recombinant CS, mortality rates from each strain were normalized and a
378
statistical model was used to predict the position of a single locus or multiple loci. Results from the
379
different models were compared and two critical regions showed the highest likelihood of containing
380
the char2 genes. One, on the proximal portion of chromosome 8 (35-42.6 cM), was mapped on the
381
same position as in the original studies (47, 49). The other, on the distal portion of the chromosome
382
(68.9-75.4 cM), was suggested in the first congenic study (28) and it could be now confirmed as the
383
second char2 locus controlling susceptibility to malaria.
384
A different approach to refine the position of char2 in the proximal region of chromosome 8 was
385
carried out using an F11 AIL from an F1 cross of A/J and C57BL/6J mice (63). The power of the AIL
386
approach in fine mapping resistance loci has been reported in murine trypanosomiasis and pulmonary
387
adenoma studies (70, 173), reducing 95% confidence intervals of loci positions up to 7-fold in
388
comparison with those observed in F2 studies. However the 95% confidence interval of char2 location
389
(between 13 and 16 cM) could not be narrowed down as expected from the AIL approach. This was
390
most likely due to the little phenotypic differences found in the parental mouse strains infected by P.
391
chabaudi chabaudi 54X when compared to the well distinguished phenotypes observed in infections by
392
P. chabaudi chabaudi AS (156). As it has been shown in later studies, environmental factors as well as
393
parasite polymorphisms among isolates might influence the severity of malaria between susceptible and
394
resistant mice (64). Moreover, the AIL study did not detect the presence of linked loci on the proximal
395
region of chromosome 8 as suggested by the first congenic line study (28) and supporting the results
396
from the second congenic line study (91) which, showed the location of the second char2 locus on the
397
distal region of the chromosome. Two more candidate genes were presented to control parasitemia in
398
the AIL study: chondroitin sulfate proteoglycan 3 (Cspg3), a major constituent in blood vessels, and
399
interleukin 12 receptor beta 1 (il12rb1).
16
400
While many efforts towards the positional cloning of char2 genes have been performed, refinement
401
studies on the map position of char1 have not been unfortunately carrying out in spite of the relevance
402
of the locus found in two separated studies with different mouse and parasite species.
403
The first evidence of H2 locus involved in the control of survival to malaria infections in mouse
404
appeared two decades ago (183). However, the H2 locus was not significantly linked to either survival
405
or peak parasitemia in BC and BC-F2 combined studies (49). A new study using an F2 generation of
406
C3H/HeJ and C57BL/6J mice was carried out to detect loci responsible for the control of daily
407
parasitemia levels (27). A significant locus at the proximal end of chromosome 17 (a region harboring
408
the H2 complex), named char3, was identified to control parasite clearance from the blood stream after
409
peak parasitemia, whereas char1 and char2 controlled peak parasitemia levels confirming previous
410
results (47, 49). Because char3 did not play any role in the control of the peak parasitemia, but it was
411
critical in the eradication of parasites, this locus might represent the complex H2 locus that could
412
modulate T-cell responses required for the clearance of primary and secondary parasitemias (Figure 2)
413
(88). Interestingly, human MHC alleles have been involved in the protection of West African children
414
from severe malaria (67) and TNF, a gene harbored within the H2 loci, was implicated in the immune
415
response to cerebral malaria (46). The char3 region has also been involved in the control of TRBV4
416
CD8+ T cells during mononucleosis induced by γ-herpes virus (60).
417
The AIL study with an (A/JxC57BL/6J) F11 population infected with P. chabaudi chabaudi 54X
418
resolved this locus into two loci in coupling phase, one of them was mapped near the H2 region (char3)
419
whereas the other, named in this study char7, mapped within a region containing interesting candidates
420
genes, i.e. complement component 3 (C3), interleukin 25 (Il25) and immune response 5 (Ir5). Further
421
linkage studies of both loci using daily parasitemia traits revealed that char3 and char7 acted in early
422
stages of the infection before peak parasitemia, in disagreement with previous mapping studies using P.
423
chabaudi adami DS parasites. These controversial results showed that different P. chabaudi subspecies
424
and mouse-specific parasitemia development might influence the mode of action of the resistant loci.
17
425
Another remarkable effort to narrow down the positions of char1 and char2 loci and map new malaria
426
resistance loci was carried out using a set of 37 recombinant CS derived from A/J and C57BL/6J
427
parental mice (50) The individual AcB strains, created from (F1xA/J)xA/J double backcrosses,
428
contained on average 13% of the C57BL/6J genome and the total AcB set contained almost 80% of the
429
resistant genome. Susceptibility phenotypes to P. chabaudi chabaudi AS infection and genotypes at
430
char1 and char2 loci were studied in all the AcB strains (51). In particular, the AcB55 strain was found
431
to be resistant to infection (low parasitic levels and 0% mortality), despite carrying susceptibility alleles
432
at char1 and char2 loci, indicating the presence of a novel resistance locus transferred by a small
433
C57BL/6J chromosomal region into the A/J genomic background. An (AcB55xA/J)F2 population was
434
monitored for parasitic levels and mortality and genotyped at a maximum spacing of 10 cM covering
435
all murine chromosomes containing C57BL/6J genomic segments. While genetic linkage was not
436
detected when using mortality as phenotypic trait, a significant locus was detected on chromosome 3
437
(char4) when using peak parasitemia as phenotype. A second, less significant, locus (not named in this
438
study) was mapped on chromosome 10. Both loci explained 18% of the variance of the peak
439
parasitemia. The size of char4 locus was estimated between 6 and 10 cM, a region narrow enough to
440
transfer into A/J genomic background and create congenic lines for further dissection of the traits. The
441
char4 segment, syntenic with the human 4q21-q25, contained interesting candidate genes: lymphoid
442
enhancer binding factor 1 (Lef1), a transcription factor that regulates gene expression in T
443
lymphocytes; complement component factor I (Cfi) and the p105 subunit of NfκB1, an essential
444
regulator for B cell functions.
445
Further clinical examinations of two malaria resistant AcB congenic lines, AcB55 and AcB61 (the
446
latter carried susceptibility alleles at char1 and char2 but did not carry resistance allele at char4) were
447
carried out to detect possible biochemical mechanisms that could have conferred resistance to the
448
infection (106). In fact, mice from both CS showed significant splenomegaly caused by the expansion
449
of the red pulp at the time of necropsy, a higher expression of erythroid-specific (globin, transferrin
18
450
receptor and solute carrier family 11 member 2) and cell cycle genes, and abnormal higher reticulocyte
451
concentrations. Genetic analysis of (AcB55xA/J)xF2 and (AcB55xDBA/2J)xF2 populations using
452
reticulocytosis as the quantitative trait mapped a new A/J-derived locus proximal to char4 region.
453
Liver- and cell-specific pyruvate kinase (pklr) gene was selected as a strong candidate in the new locus
454
for playing essential roles in the synthesis of ATP in red blood cells and cause hemolytic anemia in
455
humans when it is mutated. DNA sequence analysis detected an isoleucine-to-asparagine substitution at
456
aminoacid 90 of the pklr protein in both AcB55 and AcB61 congenic mice. Moreover, association
457
studies demonstrated that homozygous pklr mutants had reduced peak parasitemia and mortality rates
458
when compared with heterozygous mice. These spectacular results showed that inactive Pklr269A
459
mutants were protected against P. chabaudi chabaudi AS malaria and that splenomegaly and an
460
enhanced erythropoiesis and reticulocytosis played as compensatory mechanisms to overcome the
461
infection. These investigations proved that the recombinant CS approach can successfully be used for
462
genetic dissection of complex malaria traits. The Pklr269A mutation controlling hemolytic anemia in
463
mice affected the related and more complex parasitic level trait, which might be controlled by
464
additional not yet detected genetic factors. Epidemiological studies have not detected a selective
465
pressure on piruvate kinase deficiency by the malaria parasites, as in the cases of sickle cell trait and
466
thalassaemias, but a possible protective role of this deficiency against the different malarias in humans
467
cannot be ruled out. Currently, the impact of pyruvate kinase deficiency in humans infected with
468
malaria parasites is being studied. So far, the finding of the Pklr269A mutant as a resistant gene in mice
469
represents the most remarkable achievement from genetic studies of malaria resistance in mouse
470
models and shows the optimal strategy that it should be carried out towards a successful identification
471
of more candidate genes from preliminary F2 or BC studies
472
An additional pair of loci of opposite additive effects, char5 and char6, controlling parasitemia curves
473
was mapped on chromosome 5 in the AIL F11 study (63). A previous study with a BC population of A/J
474
and C57BL/6J as parental mice detected char6 (49), although F2 and combined BC and F2 studies
19
475
could not confirm the presence of the locus. The lack of confirmation in the latter studies could have
476
been due to the use of software to perform genetic analysis to detect single loci, possibly missing the
477
detection of linked loci, as in this case, on chromosome 5. As in the case of linked char3 and char7 loci
478
on chromosome 17, the 95% confidence interval of char5 and char6 positions could not be determined.
479
Analysis of char5 and char6 for daily parasitemia traits showed that both loci seemed to act in the days
480
of highest parasitemia, in agreement with the results from the suggestive linkage to peak parasitemia in
481
the BC study. Linked char5 and char6 regions contained a substantial number of potential candidates
482
genes: blood antigen acetylcholinesterase (Ache), erythropoietin (Epo), heat shock 27-kDa protein 1
483
(Hspb1), correlation in cytokine production 1 (Cora1), NADPH oxidase subunit (Ncf1) and actin-
484
related gene 1 (Act1).
485
The comparative approach with human and mouse genomes
486
Human homologous genes may be detected by comparative mapping of human and mouse genomic
487
regions (22, 177). However, it is also possible to perform an inversed strategy and study the role of
488
homologous human loci of resistance to malaria in mouse models. The identification of genetic factors
489
involved in the malarial disease in both humans and mice could reveal common genes controlling
490
similar mechanisms of infection from different mammalian malarias (101).
491
Several association and linkage analysis of P. falciparum parasitemia mapped pfbi (Plasmodium
492
falciparum blood infection) locus on the 5q31-q33 chromosome region (see Table 1), flanked by DNA
493
microsatellite D5S642 and interleukin 12 gene (IL12B) (143). The segment contained several candidate
494
genes involved in the regulation of the immune TH1 and TH2 responses to the infection such as
495
interleukin 4 (IL4), IL12B and interferon regulatory factor 1 (IRF1). Interestingly, the 5q31-q33 region
496
has also been identified to control Schistosoma mansoni infections (161) and regulating
497
immunoglobulin E levels (101). In an attempt to find novel murine loci for control of parasitemia on
498
regions homologous to human chromosome 5q31-q33 (Figure 3), the (A/JxC57BL/6J)F11 population of
499
mice was genotyped on chromosomes 11 and 18, but not on chromosome 13 (65). A new locus, named
20
500
char8, was found on chromosome 11 controlling parasitic levels. The locus was more significant for
501
the trait with parasitemia scores up to day 11 post infection than for the trait with parasitemia scores up
502
to day 7 post infection, indicating that char8 might act from early to late stages of the parasitemia
503
curve. Although more than 300 genes were found within the 95% confidence interval of char8, Il3, Il4,
504
Il5, Il12b, Il13, Irf1, granulocyte-macrophage colony-stimulating factor 2 (Csf2), growth differentiation
505
factor 9 (Gdf9), heat shock protein 4 (Hspa4), T-cell phenotype modifier (Tcpm1), T-cell-specific
506
transcription factor 7 (Tcf7), il2-inducible T-cell kinase (itk) and T-cell immunoglobin domain and
507
mucin domain containing genes (Tim1 and Tim3) were presented as interesting candidate genes with
508
homologues in human chromosome 5q31-q33.
509
Moreover, one of them Csf2, has been involved in the development of splenomegaly, leukocytosis, and
510
granulocyte-macrophage hematopoiesis, controlling the resistance to P. chabaudi infections (145).
511
Because any genetic factor controlling the malaria infection was not found on mouse chromosome 18,
512
this locus comparative approach might lead further dissection studies to map pfbi gene(s) on
513
chromosome 5q31-q33.
514
Genetic loci linked to murine cerebral malaria resistance
515
P. berghei ANKA induces the symptoms that can mimic some of the aspects of human cerebral malaria
516
in susceptible mouse strains. These aspects include respiratory distress syndrome, low body
517
temperature, ataxia, paralysis and coma, followed by death. Histopathological studies showed that
518
sequestration of parasitized red blood cells in brain vessels and damage of endothelial tissues were
519
involved in the irreversible effects of the infection (120, 121, 139). Based on the percentage of
520
mortality derived from cerebral malaria, laboratory strains of mouse can be divided into highly
521
susceptible (e.g. C57BL/6J, CBA/J and 129/SvJ), weakly susceptible (e.g. BALB/cJ and C3H) and
522
resistant (e.g. DBA/2J) (16). Interestingly, susceptible mice died from neurological damages with low
523
parasitemia whereas resistant mice died with severe anemia and high parasitic levels.
21
524
The first genetic study of resistance to cerebral malaria in mouse used an F2 population of susceptible
525
C57BL/6J and resistant DBA/2J mice (118). Animals surviving to day 14 post infection were
526
phenotyped as resistant whereas those dying before that day were phenotyped as susceptible. The
527
analyses showed a significant locus (unnamed in this study) in the middle region of chromosome 18
528
controlling mortality to cerebral malaria. Less significant loci were found on chromosomes 5, 8 and 14.
529
Strikingly, the position of the locus on chromosome 8 mapped within the char2 region, suggesting a
530
common segment involved in two different malaria phenotypes, parasitemia and cerebral malaria.
531
Several candidate genes on chromosome 18 were suggested to control cerebral malaria: colony-
532
stimulating factor 1 receptor (Csf1r), the platelet-derived growth factor receptor, beta peptide (Pdgfrb)
533
and the CD14 antigen (Cd14). Interestingly, the expression of the latter appeared to increase in acute
534
respiratory distress syndrome, one of the features associated with the fatalities from cerebral malaria
535
(94).
536
In order to find out new strains of mouse resistant to cerebral malaria and make use of more genetic
537
polymorphisms, several wild-derived mouse strains from Asia, Western and Eastern Europe and
538
Northern Africa were challenged with P. berghei ANKA. Six of the twelve strains tested were resistant
539
to cerebral malaria (16). WLA, a Mus musculus domesticus strain, has been used as the resistant mouse
540
line in subsequent genetic resistance studies. In fact, a (WLAxC57BL/6J)F1 population was
541
backcrossed to C57BL/6J. The resulting animals were infected with P. berghei ANKA, genotyped
542
covering the entire genome and phenotyped as susceptible when dying with typical neurological
543
symptoms between day 5 and 12 post infection (17). Genetic analysis found two loci located on
544
chromosomes 1, berr1 (Berghei resistance locus 1), and 11, berr2, controlling cerebral malaria
545
resistance. It was noted that transforming growth factor β2 (Tgfb2), a gene known to control the
546
severity of mouse infections by P. chabaudi and P. berghei (38,128) is located 0.5 cM apart from the
547
marker with the highest significance for genetic linkage on chromosome 1. A new linkage analysis
548
genotyping for Tgfb2 showed equal association strength, suggesting this gene as a potential candidate
22
549
for cerebral malaria control. A parallel study with a series of recombinant inbred mice from susceptible
550
BALB/cJ (non mammary tumor virus 7, Mtv-7, gene carriers) and resistant DBA/2J (Mtv-7 carriers),
551
showed that animals with the Mtv-7 gene inserted in their genomes were resistant where those without
552
the insert were susceptible to cerebral malaria (57). However, because Mtv-7 was not mapped within
553
the most significant linked segment on chromosome 1 to the cerebral malaria phenotypes, this gene was
554
not presented as a potential candidate.
555
A new cerebral malaria susceptibility locus, cmsc (cerebral malaria susceptibility), within the H2
556
region on chromosome 17 was found in a BC population bred from susceptible CBA and resistant
557
DBA/2 animals infected with P. berghei ANKA (127). This study did not find previously reported loci
558
such as berr1, berr2 or the unnamed locus on chromosome 18, raising the possibility of the existence
559
of different susceptible loci that contribute to the cerebral malarias of CBA and C57BL/6 mice. In
560
agreement with a previous study on susceptibility of different mouse strains to P. berghei ANKA-
561
induced cerebral malaria (146), which showed that H2 haplotypes were not highly involved in the
562
control of the pathogenesis, cmsc could not be identified as an H2 susceptible haplotype since resistant
563
mice such as C3H/HeJ and AKR share the same CBA-susceptible haplotype. However, as char3, cmsc
564
mapped near attractive candidate genes such as Tnf and lymphotoxin a (Lta) which, are critical
565
mediators of the pathogenesis induced by cerebral malaria. Recently, a gene knock-out study has
566
shown that Tnf-deficient C57BL/6J mice had higher peak parasitemias and mortality rates than
567
proficient animals (66). The multiple roles of Tnf in the different malaria pathologies as its different
568
modes of action during the parasitic infection still remain to be fully understood. Current sequence
569
analysis of both Tnf and Lta genes from CBA and DBA/2 mice might suggest the identity of the
570
causative genetic variant within the cmsc locus in the near future.
571
Previous genetic studies to find novel loci associated with resistance to cerebral malaria used the
572
reported (WLAxC57BL/6J)F2 population of mice (30). Using survival time and cure as cerebral malaria
573
phenotypes, a resistant locus on distal chromosome 1 was found, mapping within the previously
23
574
reported berr1. Two more loci, berr3 and berr4, were mapped on chromosome 9 and 4, respectively.
575
Interestingly, berr3 mapped near to previously reported char1 and pymr loci, involved in the resistance
576
to infections by P. chabaudi and P. yoelii. This region has gained a special interest since it might
577
harbor a common resistant gene to the infections of several murine Plasmodium species and it might
578
represent a universal mechanism of defense against different parasites. Recently, a deficiency in
579
chemokine receptor Ccr5 (encoding gene is located within the berr3 region) has been associated to
580
reduced susceptibility of P. berghei cerebral malaria (19). The attractive genes in the berr4 segment are
581
a toll-like receptor (Tlr4) and the complement component 8 beta subunit.
582
Conclusions and Future Directions
583
A mapping experiment using the human parasite requires the use of mosquitoes to infect chimpanzees
584
with P. falciparum gametocytes. Murine malarias make possible to work entirely in the mouse, a far
585
more tractable and cost-effective model. Such studies have focused on increasing our understanding of
586
the parasite, including the development of drug resistance, antigenic variation, and erythrocytic
587
invasion pathways. Mouse models have also been extensively used to dissect the host response to
588
disease including the innate and adaptive immune responses. This review has shown the use of mouse
589
models to map genetic regions controlling malaria traits that need to be further dissected to find the
590
resistant genes. In the immediate future, many research groups will generate congenic lines for the
591
known QTL and the results from examinations of the biological features of these lines will be
592
presented. These could give some interesting insights into associated phenotypes ahead of the
593
identification of the underlying genes.
594
The development of microarray technology with the spotting of mouse genes on chips is presently
595
being used to follow changes in expression of all mouse genes as a response to infection. In particular
596
we now are able to contrast the response of resistant and susceptible genotypes. Recently, Delahaye et
597
al. (37) were successful in using this technology to identify gene expression profiles discriminating
598
resistant and susceptible mouse during cerebral malaria infection. It is hoped that this will illuminate
24
599
the differences between the cascades of events in the two mouse strains, and determine which genes,
600
within the QTL, participate in such pathways. This will be of particular interest to those loci isolated as
601
congenic donor interval with available isogenic control lines, and will help to identify host resistance
602
genes. Indeed, the approach will lead to identify genes that are differentially expressed between
603
resistant and susceptible mice, and that will be considered strong candidates to be tested by using gene
604
targeting. Bioinformatics and experimental analyses will further identify cis-regulatory polymorphisms
605
causing changes in expression levels; these include the analysis of the sequence to search for
606
transcription binding sites and experimental methods exploring DNA-proteins interactions, such as
607
chromatin immuno-precipitation or electro mobility shift assay. Finally, with the development of the
608
new technology, namely the next generation sequence (NGS) it may be possible to determine the gene
609
expression variations of the whole genome (11, 122, 168), and combining this technology with the
610
mouse and human genetic resources, may help in identifying both genes and gene networks involved in
611
host resistance.
612
Expanding the mapping studies to larger human populations (cohorts) in malaria endemic areas will
613
provide further statistical power for identifying genetic factors underlying human phenotypic variations
614
to the disease. Combining and conduct new of analysis of the accumulated human studies can offer
615
new tool for fine mapping and identifying host genes associated with the disease.
616
617
The novel genetic reference population, namely the Collaborative Cross (CC) mice will eventually
618
comprise a set of approximately 500 recombinant inbred lines (RIL) created from full reciprocal
619
matings of 8 divergent strains of mice, including five classical inbred lines (A/J, C57BL/6J,
620
129S1/SvImJ, NOD/LtJ, and NZO/HiLtJ) and three wild-derived strains, will offer a new opportunity
621
of mapping and subsequently indentifying new genetic factors underlying variations of host
622
susceptibility to malaria (33, 162, 166). It was shown that CC mice are promising GRP, which will
623
allow identifying genes underlying complex traits, which can be applicable to malaria.
25
624
The convergent results of linkage studies in humans and in mouse models encourages to perform a
625
genome-wide screening in humans, and to compare the identified chromosomal regions in humans with
626
those previously identified in mouse models. The identification of human polymorphisms associated
627
with resistance to malaria, and the evaluation of candidate genes in the mouse model should provide
628
consistent new insights into the mechanisms of malaria resistance. This information will open the doors
629
for understanding the communication, which must occur between host and parasite.
630
Finally, it is expected that we identifying further genetic factors underling host susceptibility to
631
malaria, it will open the venue for personalized medicine application, which will benefits a significant
632
percentage of human populations, which may not response to specific drug or future vaccines.
633
26
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
REFERENCES
1. Abel L, Cot M, Mulder L, Carnevale P, Feingold J, Segregation analysis detects a major gene
controlling blood infection levels in human malaria. Am J Hum Genet, 50, 1308-1317, 1992.
2. Ackerman H, Usen S, Jallow M, Sisay-Joof F, Pinder M, et al., A comparison of case-control and
family-based association methods: the example of sickle-cell and malaria. Ann Hum Genet, 69,
559-565, 2005.
3. Agarwal A, Guindo A, Cissoko Y, Taylor JG, Coulibaly D, et al., Hemoglobin C associated with
protection from severe malaria in the Dogon of Mali, a West African population with a low
prevalence of hemoglobin S. Blood, 96, 2358-2363, 2000.
4. Aidoo M, Terlouw DJ, Kolczak MS, McElroy PD, ter Kuile FO, et al., Protective effects of the
sickle cell gene against malaria morbidity and mortality. Lancet, 359, 1311-1312, 2002.
5. Aidoo M, McElroy PD, Kolczak MS, Terlouw DJ, ter Kuile FO, et al., Tumor necrosis factor-alpha
promoter variant 2 (TNF2) is associated with pre-term delivery, infant mortality, and malaria
morbidity in western Kenya: Asembo Bay Cohort Project IX. Genet Epidemiol, 21, 201-211,
2001.
6. Aitman TJ, Cooper LD, Norsworthy PJ, Wahid FN, Gray JK, et al., Malaria susceptibility and CD36
mutation. Nature, 405, 1015-1016, 2000.
7. Allen SJ, O'Donnell A, Alexander ND, Alpers MP, Peto T, et al., alpha+-Thalassemia protects
children against disease caused by other infections as well as malaria. PNAS, 94, 14736-14741,
1997.
8. Allen SJ, Bennett S, Riley EM, Rowe PA, Jakobsen PH, et al., Morbidity from malaria and immune
responses to defined Plasmodium falciparum antigens in children with sickle cell trait in The
Gambia. Trans R Soc Trop Med Hyg, 86, 494-498, 1992.
9. Allen SJ, O'Donnell A, Alexander ND, Mgone CS, Peto TE, et al., Prevention of cerebral malaria in
children in Papua New Guinea by southeast Asian ovalocytosis band 3. Am J Trop Med Hyg,
60, 1056-1060, 1999.
10. Allison AC, The distribution of the sickle-cell trait in East Africa and elsewhere, and its apparent
relationship with the incidence of subtertian malaria. Trans R Soc Trop Med Hyg, 48, 312-318,
1954.
11. Ansorge WJ, Next-generation DNA sequencing techniques. New Biotechnology 25(4): 195-203,
2009.
12. Aucan C, Traore Y, Fumoux F, Rihet P, Familial correlation of immunoglobulin G subclass
responses to Plasmodium falciparum antigens in Burkina Faso. Infect Immun, 69, 996-1001,
2001.
13. Aucan C, Traore Y, Tall F, Nacro B, Traore-Leroux T, et al., High immunoglobulin G2 (IgG2) and
low IgG4 levels are associated with human resistance to Plasmodium falciparum malaria. Infect
Immun, 68, 1252-1258, 2000.
14. Aucan C, Walley AJ, Greenwood BM, Hill AV, Haptoglobin genotypes are not associated with
resistance to severe malaria in The Gambia. Trans R Soc Trop Med Hyg, 96, 327-328, 2002.
15. Aucan C, Walley AJ, Hennig BJ, Fitness J, Frodsham A, et al., Interferon-alpha receptor-1
(IFNAR1) variants are associated with protection against cerebral malaria in The Gambia.
Genes Immun, 4, 275-282, 2003.
16. Bagot S, Idrissa Boubou M, Campino S, Behrschmidt C, Gorgette O, Guenet JL, Penha-Goncalves
C, Mazier D, Pied S, Cazenave PA, Susceptibility to experimental cerebral malaria induced
by Plasmodium berghei ANKA in inbred mouse strains recently derived from wild stock.
Infect Immun, 70, 2049-2056, 2002.
27
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
17. Bagot S, Campino S, Penha-Goncalves C, Pied S, Cazenave PA, Holmberg D, Identification of two
cerebral malaria resistance loci using an inbred wild-derived mouse strain. PNAS, 99, 99199923, 2002.
18. Bellamy R, Kwiatkowski D, Hill AV, Absence of an association between intercellular adhesion
molecule 1, complement receptor 1 and interleukin 1 receptor antagonist gene polymorphisms
and severe malaria in a West African population. Trans R Soc Trop Med Hyg, 92, 312-316,
1998.
19. Belnoue E, Kayibanda M, Deschemin JC, Viguier M, Mack M, Kuziel WA, Renia L, CCR5
deficiency decreases susceptibility to experimental cerebral malaria. Blood,101, 4253-4259,
2003.
20. Bellamy R, Kwiatkowski D, Hill AV, Absence of an association between intercellular adhesion
molecule 1, complement receptor 1 and interleukin 1 receptor antagonist gene polymorphisms
and severe malaria in a West African population. Trans R Soc Trop Med Hyg, 92, 312-316,
1998.
21. Benoit-Vical F, Robert A, Meunier B, In vitro and in vivo potentiation of artemisinin and synthetic
endoperoxide antimalarial drugs by metalloporphyrins. Antimicrob Agents Chemother, 44,
2836-2841, 2000.
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
22. Birney E, Andrews D, Caccamo M, Chen Y, et al., Ensembl 2006. Nucleic Acids Res, 34, D556D561, 2006.
23. Boldt AB, Luty A, Grobusch MP, Dietz K, Dzeing A, et al., Association of a new mannose-binding
lectin variant with severe malaria in Gabonese children. Genes Immun, 7, 393-400, 2006.
24. Bouharoun TH, 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, 172,
1633-1641, 1990.
25. Burgner D, Xu W, Rockett K, Gravenor M, Charles IG, et al., Inducible nitric oxide synthase
polymorphism and fatal cerebral malaria (letter) (see comments). Lancet, 352, 1193-1194,
1998.
26. Burgner D, Usen S, Rockett K, Jallow M, Ackerman H, et al., Nucleotide and haplotypic diversity
of the NOS2A promoter region and its relationship to cerebral malaria. Hum Genet, 112, 379386, 2003.
27. Burt RA, Baldwin TM, Marshall VM, Foote SJ Temporal expression of an H2-linked locus in host
response to mouse malaria. Immunogenetics, 50, 278-285, 1999.
714
715
716
28. Burt RA, Marshall VM, Wagglen J, Rodda FR, Senyschen D., Baldwin TM, Buckingham LA,
Foote SJ Mice that are congenic for the char2 locus are susceptible to malaria. Infect Immun,
70, 4750-4753, 2002.
717
718
719
720
721
722
723
724
29. Cabantous S, Poudiougou B, Traore A, Keita M, Cisse MB, et al., Evidence that interferon-gamma
plays a protective role during cerebral malaria. J Infect Dis, 192, 854-860, 2005.
30. Campino S, Bagot S, Bergman ML, Almeida P, Sepulveda N, Pied S, Penha-Goncalves C,
Holmberg D, Cazenave PA, Genetic control of parasite clearance leads to resistance to
Plasmodium berghei ANKA infection and confers immunity. Genes Immun, 6, 416-421, 2005.
31. Carter R, Walliker D, New observations on the malaria parasites of the Central African Republic:
Plasmodium vinckei petteri supsp.nov. and Plasmodium chabaudi. Ann Trop Med Parasitol, 69,
187-196, 1975.
725
726
32. Carter, R. & Diggs, C.L. Plasmodia of rodents. In: Parasitic Protozoa. New York, San Francisco,
London: Academic press, vol. III, 359-465, 1977.
727
33. Churchill GA, Airey DC, Allayee H, Angel JM, Attie AD, Beatty J, Beavis WD, Belknap
28
728
729
730
731
732
733
734
JK, Bennett B, Berrettini W, Bleich A, Bogue M, Broman KW, Buck KJ, Buckler E,
Burmeister M, Chesler EJ, Cheverud JM, Clapcote S, Cook MN, Cox RD, Crabbe JC,
Crusio WE, Darvasi A, Deschepper CF, Doerge RW, Farber CR, Forejt J, Gaile D,
Garlow SJ, Geiger H, Gershenfeld H, Gordon T, Gu J, Gu W, de Haan G, Hayes NL,
Heller C, Himmelbauer H, Hitzemann R, Hunter K, Hsu HC, Iraqi FA, Ivandic B, Jacob
HJ, et al., The Collaborative Cross, a community resource for the genetic analysis of
complex traits. Nat Genet, 36, 1133-1137, 2004.
735
736
34. Conway DJ, Tracing the dawn of Plasmodium falciparum with mitochondrial genome
sequences. Trends Genet, 19, 671-674, 2003.
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
35. Cooke GS, Aucan C, Walley AJ, Segal S, Greenwood BM, et al., Association of Fcgamma receptor
IIa (CD32) polymorphism with severe malaria in West Africa. Am J Trop Med Hyg, 69, 565568, 2003.
36. Delahaye NF, Barbier M, Fumoux F, Rihet P, Association analyses of NCR3 polymorphisms with
mild malaria. Microbes Infect: doi:10.1016/j.micinf. 2006.1011.1002, 2006.
37. Delahaye NF, Coltel N, Puthier D, Flori L, Houlgatte R, Iraqi FA, Nguyen C, Grau GE, Rihet, P,
Gene expression profiling discriminates cerebral malaria-susceptible mice and cerebral malariaresistant mice. J Infectious Diseases, 193, 312-321, 2006.
38. de Kossodo S, Grau GE, Profiles of cytokine production in relation with susceptibility to cerebral
malaria. J Immunol, 151, 4811-4820, 1993.
39. Dolo A, Modiano D, Maiga B, Daou M, Dolo G, et al., Difference in susceptibility to malaria
between two sympatric ethnic groups in Mali. Am J Trop Med Hyg, 72, 243-248, 2005.
40. Elagib AA, Kider AO, Akerstrom B, Elbashir MI (1998) Association of the haptoglobin phenotype
(1-1) with falciparum malaria in Sudan. Trans R Soc Trop Med Hyg, 92, 309-311.
41. Engelbrecht F, Felger I, Genton B, Alpers M, Beck HP, Plasmodium falciparum: malaria morbidity
is associated with specific merozoite surface antigen 2 genotypes. Exp Parasitol, 81, 90-96,
1995.
42. Fernandez RD, Craig AG, Kyes SA, Peshu N, Snow RW, et al., A high frequency African coding
polymorphism in the N-terminal domain of ICAM-1 predisposing to cerebral malaria in Kenya.
Hum Mol Genet, 6, 1357-1360, 1997.
43. Fischer PR, Boone P, Short report: severe malaria associated with blood group. Am J Trop Med
Hyg, 58, 122-123, 1998.
44. Flori L, Kumulungui B, Aucan C, Esnault C, Traore AS, et al., Linkage and association between
Plasmodium falciparum blood infection levels and chromosome 5q31-q33. Genes Immun, 4,
265-268, 2003.
45. Flori L, Sawadogo S, Esnault C, Delahaye NF, Fumoux F, et al., Linkage of mild malaria to the
major histocompatibility complex in families living in Burkina Faso. Hum Mol Genet, 12, 375378, 2003.
46. Flori L, Delahaye NF, Iraqi FA, Hernandez-Valladares M, Fumoux F, et al., TNF as a malaria
candidate gene: polymorphism-screening and family-based association analysis of mild malaria
attack and parasitemia in Burkina Faso. Genes Immun, 6, 472-480, 2005.
47. Foote SJ, Burt RA, Baldwin TM, Presente A, Roberts AW, et al., Mouse loci for malaria-induced
mortality and the control of parasitaemia. Nat Genet, 17, 380-381, 1997.
48. Foote SJ, Iraqi F, Kemp SJ. Controlling malaria and African trypanosomiasis: the role of the
mouse. Brief Funct Genomic Proteomic, 4(3), 214-24, 2005.
49. Fortin A, Belouchi A, Tam MF, Cardon L, Skamene E, et al., Genetic control of blood parasitaemia
in mouse malaria maps to chromosome 8. Nat Genet, 17, 382-383, 1997.
29
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
50. Fortin A, Diez E, Rochefort D, Laroche L, Malo D, Rouleau GA, Gros P, Skamene E, Recombinant
congenic strains derived from A/J and C57BL/6J: a tool for genetic dissection of complex traits.
Genomics, 74, 21-35, 2001.
51. Fortin, A., Cardon, L.R., Tam, M., Skamene, E., Stevenson, M.M. & Gros, P. Identification of a
new malaria susceptibility locus (Char4) in recombinant congenic strains of mice. PNAS, 98,
10793-10798, 2001.
52. Garcia A, Cot M, Chippaux JP, Ranques S, Feingold J, et al., Genetic control of bloof infection
levels in human malaria: evidence for a complex genetic model. Am J Trop Med Hyg, 58, 480488, 1998.
53. Garcia A, Marquet S, Bucheton B, Hillaire D, Cot M, et al., Linkage analysis of blood Plasmodium
falciparum levels: interest of the 5q31-q33 chromosome region. Am J Trop Med Hyg, 58, 705709, 1998.
54. Genton B, al-Yaman F, Mgone CS, Alexander N, Paniu MM, et al., Ovalocytosis and cerebral
malaria. Nature, 378, 564-565, 1995.
55. Gilbert SC, Plebanski M, Gupta S, Morris J, Cox M, et al., Association of malaria parasite
population structure, HLA, and immunological antagonism. Science, 279, 1173-1177, 1998.
56. Gilles HM, Fletcher KA, Hendrickse RG, Lindner R, Reddy S, et al., Glucose-6-phosphatedehydrogenase deficiency, sickling, and malaria in African children in South Western Nigeria.
Lancet, 1, 138-140, 1967.
57. Gorgette O, Existe A, Boubou MI, Bagot, S, Guenet JL, Mazier D, Cazenave PA, Pied S, Deletion
of T cells bearing the V beta8.1 T-cell receptor following mouse mammary tumor virus 7
integration confers resistance to murine cerebral malaria. Infect Immun, 70, 3701-3706, 2002.
58. Grupe A, Germer S, Usuka J, Aud D, Belknap JK, Klein RF, Ahluwalia MK, Higuchi R, Peltz G,
In silico mapping of complex disease-related traits in mice. Science, 292, 1915-1918, 2001.
59. Haldane JBS, The rate of mutation of human genes. Proc VIII Int Cong Genethereditas, 35, 267273, 1949.
60. Hardy, CL, Lu L, Nguyen, P, Woodland, DL, Williams RW, Blackman MA, Identification of
quantitative trait loci controlling activation of TRBV4 CD8+ T cells during murine gammaherpesvirus-induced infectious mononucleosis. Immunogenetics, 53, 395-400, 2001.
61. Hay SI, Guerra CA, Tatem AJ, Atkinson PM, Snow RW, Urbanization, malaria transmission and
disease burden in Africa. Nat Rev Microbiol, 3, 81-90, 2005.
62. Helbok R, Dent W, Nacher M, Lackner P, Treeprasertsuk S, et al., The use of the multi-organdysfunction score to discriminate different levels of severity in severe and complicated
Plasmodium falciparum malaria. Am J Trop Med Hyg, 72, 150-154, 2005.
63. Hernandez-Valladares M, Naessens J, Gibson JP, Musoke AJ, Nagda S, et al., Confirmation and
dissection of QTL controlling resistance to malaria in mice. Mamm Genome, 15, 390-398,
2004.
64. Hernandez-Valladares M, Naessens J, Nagda S, Musoke AJ, Rihet P, Ole-Moiyoi, OK, Iraqi FA,
Comparison of pathology in susceptible A/J and resistant C57BL/6J mice after infection with
different sub-strains of Plasmodium chabaudi. Exp Parasitol, 108, 134-141, 2004.
816
817
818
65. Hernandez-Valladares M, Rihet P, ole-MoiYoi OK, Iraqi FA, Mapping of a new quantitative trait
locus for resistance to malaria in mice by a comparative mapping approach with human
Chromosome 5q31-q33. Immunogenetics, 56, 115-117, 2004.
819
820
821
66. Hernandez-Valladares M, Naessens J, Musoke AJ, Sekikawa K, Rihet P, et al., Pathology of Tnfdeficient mice infected with Plasmodium chabaudi adami 408XZ. Exp Parasitol, 114, 271-278,
2006.
30
822
823
824
825
826
827
828
829
830
831
67. Hill AV, Allsopp CE, Kwiatkowski D, Anstey NM, Twumasi P, et al., Common west African HLA
antigens are associated with protection from severe malaria (see comments). Nature, 352, 595600, 1991.
68. Hill AV, The immunogenetics of resistance to malaria. Proc Assoc Am Physicians, 111, 272-277,
1999.
69. Hobbs MR, Udhayakumar V, Levesque MC, Booth J, Roberts JM, et al., A new NOS2 promoter
polymorphism associated with increased nitric oxide production and protection from severe
malaria in Tanzanian and Kenyan children. Lancet, 360, 1468-1475, 2002.
70. Iraqi F, Clapcott SJ, Kumari P, Haley CS, Kemp SJ, Teale AJ, Fine mapping of trypanosomiasis
resistance loci in murine advanced intercross lines. Mamm. Genome, 11, 645-648, 2000.
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
71. Ishimori N, Li R, Kelmenson PM, Korstanje R, Walsh KA, Churchill GA, Forsman-Semb K,
Paigen B, Quantitative trait loci analysis for plasma HDL-cholesterol concentrations and
atherosclerosis susceptibility between inbred mouse strains C57BL/6J and 129S1/SvImJ.
Arterioscler Thromb Vasc Biol, 24, 161-166, 2004.
72. Jepson AP, Banya WA, Sisay JF, Hassan KM, Bennett S, et al., Genetic regulation of fever in
Plasmodium falciparum malaria in Gambian twin children. J Infect Dis, 172, 316-319, 1995.
73. Jepson A, Banya W, Sisay JF, Hassan KM, Nunes C, et al., Quantification of the relative
contribution of major histocompatibility complex (MHC) and non-MHC genes to human
immune responses to foreign antigens. Infect Immun, 65, 872-876, 1997.
74. Jepson A, Sisay JF, Banya W, Hassan KM, Frodsham A, et al., Genetic linkage of mild malaria to
the major histocompatibility complex in Gambian children: study of affected sibling pairs. Br
Med J, 315, 96-97, 1997.
75. Kikuchi M, Looareesuwan S, Ubalee R, Tasanor O, Suzuki F, Wattanagoon Y, Na-Bangchang K,
Kimura A, Aikawa M, Hirayama K, Association of adhesion molecule PECAM-1/CD31
polymorphism with susceptibility to cerebral malaria in Thais. Parasitol Int, 50, 235-239, 2001.
76. Knight JC, Udalova I, Hill AV, Greenwood BM, Peshu N, et al. A polymorphism that affects OCT1 binding to the TNF promoter region is associated with severe malaria (see comments). Nat
Genet, 22, 145-150, 1999.
77. Koch O, Awomoyi A, Usen S, Jallow M, Richardson A, et al., IFNGR1 gene promoter
polymorphisms and susceptibility to cerebral malaria. J Infect Dis, 185, 1684-1687, 2002.
78. Koch O, Rockett K, Jallow M, Pinder M, Sisay-Joof F, et al., Investigation of malaria susceptibility
determinants in the IFNG/IL26/IL22 genomic region. Genes Immun, 6, 312-318, 2005.
79. Korstanje R, Li R, Howard T, Kelmenson P, Marshall J, Paigen B, Churchill G, Influence of sex
and diet on quantitative trait loci for HDL cholesterol levels in an SM/J by NZB/BlNJ intercross
population. J Lipid Res, 45, 881-888, 2004.
857
858
859
860
861
862
863
864
865
866
867
80. Kun JF, Mordmuller B, Lell B, Lehman LG, Luckner D, et al., Polymorphism in promoter region of
inducible nitric oxide synthase gene and protection against malaria. Lancet, 351, 265-266, 1998.
81. Kun JF, Klabunde J, Lell B, Luckner D, Alpers M, et al. , Association of the ICAM-1Kilifi
mutation with protection against severe malaria in Lambarene, Gabon (In Process Citation). Am
J Trop Med Hyg, 61, 776-779, 1999.
82. Kun JF, Mordmuller B, Perkins DJ, May J, Mercereau-Puijalon O, et al., Nitric oxide synthase 2
(G-954C) increased nitric oxid production, and protection against malaria. J Infect Dis, 184,
330-336, 2001.
83. Lamb TJ, Brown DE, Potocnik AJ, Langhorne, J, Insights into the immunopathogenesis of malaria
using
mouse
models.
Expert Rev Mol Med, 8, 1-22, 2006.
868
84. Landau I, Killick-Kendrick R, Rodent plasmodia of the Republique Centrafricaine: The sporogony
31
869
870
and tissue stages of Plasmodium chabaudi and P. berghei yoelii. Trans R Soc Trop Med Hyg,
60, 633-649, 1966.
871
872
873
85. Landau I, Michel JC, Adam JP, Boulard Y, The life-cycle of Plasmodium vinckei lentum n.subsp.
in the laboratory; comments on the nomenclature of the murine malaria parasites. Ann Trop
Med Parasitol, 64, 315-323, 1970.
874
875
86. Landau I, & Boulard, Y, Life cycles and morphology. In: R. Killick-Kendrick W, Peters, eds.
Rodent Malaria. London: Academic Press; 53-84, 1978.
876
877
878
87. Lander ES, Schork NJ, Genetic dissection of complex traits. Science, 265, 2037-2048, 1994.
88. Langhorne J, Meding SJ, Eichmann K, Gillard SS, The response of CD4+ T cells to Plasmodium
chabaudi chabaudi. Immunol Rev, 112, 71-94, 1989.
879
880
881
882
883
884
885
89. Lell B, May J, Schmidt-Ott RJ, Lehman LG, Luckner D, et al., The role of red blood cell
polymorphisms in resistance and susceptibility to malaria. Clin Infect Dis, 28, 794-799, 1999.
90. Levesque MC, Hobbs MR, Anstey NM, Vaughn TN, Chancellor JA, et al., Nitric oxide synthase
type 2 promoter polymorphisms, nitric oxide production, and disease severity in tanzanian
children with malaria (In Process Citation). J Infect Dis, 180, 1994-2002, 1999.
91. Lin E, Pappenfuss T, Tan RB, Senyschyn D, Bahlo M, Speed TP, Foote SJ, Mapping of the
Plasmodium chabaudi resistance locus char2. Infect Immun, 74, 5814-5819, 2006.
886
887
888
92. Longley R, Smith C, Fortin A, Berghout J, McMorran B, Burgio G, Foote S, Gros P. Host
resistance to malaria: using mouse models to explore the host response. Mamm Genome, 22(12), 32-42, 2011.
889
890
891
93. Lou J, Lucas R, Grau GE, Pathogenesis of cerebral malaria: recent experimental data and possible
applications
for
humans.
Clin Microbiol Rev, 14, 810-820, 2001.
892
893
894
895
896
897
898
94. Lucas R, Lou J, Morel DR, Ricou B, Suter, PM, Grau GE, TNF receptors in the microvascular
pathology of acute respiratory distress syndrome and cerebral malaria. J Leukoc Biol, 61,
551-558, 1997.
95. Luoni G, Verra F, Arca B, Sirima BS, Troye-Blomberg M, et al., Antimalarial antibody levels and
IL4 polymorphism in the Fulani of West Africa. Genes Immun, 2, 411-414, 2001.
96. Luty AJ, Kun JF, Kremsner PG, Mannose-binding lectin plasma levels and gene polymorphisms in
Plasmodium falciparum malaria. J Infect Dis, 178, 1221-1224, 1998.
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
97. Mackinnon MJ, Gunawardena DM, Rajakaruna J, Weerasingha S, Mendis KN, et al., Quantifying
genetic and nongenetic contributions to malarial infection in a Sri Lankan population. PNAS,
97, 12661-12666, 2000.
98. Mackinnon MJ, Mwangi TW, Snow RW, Marsh K, Williams TN, Heritability of malaria in Africa.
PLoS Med 2: e340, 2005.
99. Machleder D, Ivandic B, Welch C, Castellani L, Reue K, Lusis AJ, Complex genetic control of
HDL levels in mice in response to an atherogenic diet. Coordinate regulation of HDL levels and
bile acid metabolism. J Clin Invest, 99, 1406-1419, 1997.
100. Marquis JF, Gros P. Genetic analysis of resistance to infections in mice: A/J meets C57BL/6J.
Curr Top Microbiol Immunol. 321, 27-57, 2008.
101. Marsh DG, Neely JD, Breazeale DR, Ghosh B, Freidhoff LR, Ehrlich-Kautzky E, Schou C,
Krishnaswamy G, Beaty TH, Linkage analysis of IL4 and other chromosome 5q31.1 markers
and total serum immunoglobulin E concentrations. Science, 264, 1152-1156, 1994.
32
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
102. Marsh K, Snow RW, Malaria transmission and morbidity. Parassitologia, 41, 241-246, 1999.
103. Marquet S. Abel L. Hillaire D, Dessein H, Kalil J, Feingold J, Weissenbach J, Dessein AJ. Genetic
localization of a locus controlling the intensity of infection by Schistosoma mansoni on
chromosome 5q31-q33. Nat Genet, 14, 181-184, 1996.
104. McGuire W, Hill AV, Allsopp CE, Greenwood BM, Kwiatkowski D, Variation in the TNF-alpha
promoter region associated with susceptibility to cerebral malaria. Nature, 371, 508-510, 1994.
105. McGuire W, Knight JC, Hill AV, Allsopp CE, Greenwood BM, et al., Severe malarial anemia and
cerebral malaria are associated with different tumor necrosis factor promoter alleles. J Infect
Dis, 179, 287-290, 1999.
106. Min-Oo G, Fortin A, Tam MF, Nantel A, Stevenson, MM. Gros P, Pyruvate kinase deficiency in
mice
protects
against
malaria.
Nat Genet, 35, 357-362, 2003.
107. Mockenhaupt FP, Ehrhardt S, Cramer JP, Otchwemah RN, Anemana SD, et al., Hemoglobin C
and resistance to severe malaria in Ghanaian children. J Infect Dis, 190, 1006-1009, 2004.
108. Mockenhaupt FP, Ehrhardt S, Gellert S, Otchwemah RN, Dietz E, et al., Alpha(+)-thalassemia
protects African children from severe malaria. Blood, 104, 2003-2006., 2004.
109. Mockenhaupt FP, Cramer JP, Hamann L, Stegemann MS, Eckert J, et al., Toll-like receptor (TLR)
polymorphisms in African children: Common TLR-4 variants predispose to severe malaria.
PNAS, 103, 177-182, 2006.
110. Modiano D, Petrarca V, Sirima BS, Nebie I, Diallo D, et al., Different response to Plasmodium
falciparum malaria in west African sympatric ethnic groups. PNAS, 93, 13206-13211, 1996.
111. Modiano D, Chiucchiuini A, Petrarca V, Sirima BS, Luoni G, et al. , Humoral response to
Plasmodium falciparum Pf155/ring-infected erythrocyte surface antigen and Pf332 in three
sympatric ethnic groups of Burkina Faso. Am J Trop Med Hyg, 58, 220-224, 1998.
112. Modiano D, Luoni G, Sirima BS, Simpore J, Verra F, et al., Haemoglobin C protects against
clinical Plasmodium falciparum malaria. Nature, 414, 305-308, 2001.
113. Molyneux ME, Taylor TE, Wirima JJ, Borgstein A, Clinical features and prognostic indicators in
paediatric cerebral malaria: a study of 131 comatose Malawian children. Q J Med, 71, 441-459,
1989.
114. Mombo LE, Ntoumi F, Bisseye C, Ossari S, Lu CY, et al., Human genetic polymorphisms and
asymptomatic Plasmodium falciparum malaria in Gabonese schoolchildren. Am J Trop Med
Hyg, 68, 186-190, 2003.
115. Morahan G, Boutlis CS, Huang D, Pain A, Saunders JR, et al., A promoter polymorphism in the
gene encoding interleukin-12 p40 (IL12B) is associated with mortality from cerebral malaria
and with reduced nitric oxide production. Genes Immun, 3, 414-418, 2002.
116. Moreno A, Ferrer E, Arahuetes S, Eguiluz C, Van Rooijen N, Benito A,. The course of infections
and pathology in immunomodulated NOD/LtSz-SCID mice inoculated with Plasmodium
falciparum laboratory lines and clinical isolates. Int J Parasitol, 36, 361-369, 2006.
117. Mouse Genome Sequencing Consortium. Initial sequencing and comparative analysis of the
mouse genome. Nature, 420, 520-562, 2002.
954
955
956
957
958
959
960
118. Nagayasu E, Nagakura K, Akaki M, Tamiya G, Makino S, Nakano Y, Kimura M, Aikawa M,
Association of a determinant on mouse chromosome 18 with experimental severe
Plasmodium berghei malaria. Infect Immun, 70, 512-516, 2002.
119. Narum DL, Ogun SA, Batchelor AH, Holder AA Passive immunization with a multicomponent
vaccine against conserved domains of apical membrane antigen 1 and 235-kilodalton rhoptry
proteins protects mice against Plasmodium yoelii blood-stage challenge infection. Infect
Immun, 74, 5529-5536, 2006.
33
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
120. Neill AL, Hunt NH Pathology of fatal and resolving Plasmodium berghei cerebral malaria in mice.
Parasitology, 105, 165-175, 1992.
121. Neill AL, Chan-Ling T, Hunt NH Comparisons between microvascular changes in cerebral and
non-cerebral malaria in mice, using the retinal whole-mount technique. Parasitology, 107,
477-487, 1998.
122. Ng SB, Buckingham KJ, Lee C, Bigham AW, Tabor HK, Dent KM, Huff CD, Shannon PT, Jabs
EW, Nickerson DA, Shendure J and Bamshad MJ, Exome sequencing identifies the cause of a
mendelian disorder. Nat Genet. 42(1):30-35, 2010.
123. Ntoumi F, Flori L, Mayengue PI, Matondo Maya DW, Issifou S, et al., Influence of carriage of
hemoglobin AS and the Fc gamma receptor IIa-R131 allele on levels of immunoglobulin G2
antibodies to Plasmodium falciparum merozoite antigens in Gabonese children. J Infect Dis,
192, 1975-1980, 2005.
124. Ohashi J, Naka I, Patarapotikul J, Hananantachai H, Looareesuwan S, et al., Significant
association of longer forms of CCTTT Microsatellite repeat in the inducible nitric oxide
synthase promoter with severe malaria in Thailand. J Infect Dis, 186, 578-581, 2002.
125. Ohashi J, Naka I, Patarapotikul J, Hananantachai H, Looareesuwan S, et al., A single-nucleotide
substitution from C to T at position -1055 in the IL-13 promoter is associated with protection
from severe malaria in Thailand. Genes Immun, 4, 528-531, 2003.
126. Ohno T, Ishih A, Kohara Y, Yonekawa H, Terada M, Nishimura m, Chromosomal mapping of the
host resistance locus to rodent malaria (Plasmodium yoelii) infection in mice. Immunogenetics,
53, 736-740, 2001.
982
983
984
985
986
987
988
989
990
991
992
127. Ohno T, Nishimura M, Detection of a new cerebral malaria susceptibility locus, using CBA mice.
Immunogenetics, 56, 675-678, 2004.
128. Omer FM, Riley EM, Transforming growth factor beta production is inversely correlated with
severity of murine malaria infection. J Exp Med, 188, 39-48, 1998.
129. Omi K, Ohashi J, Patarapotikul J, Hananantachai H, Naka I, et al., Fcgamma receptor IIA and IIIB
polymorphisms are associated with susceptibility to cerebral malaria. Parasitol Int, 51, 361-366,
2002.
130. Omi K, Ohashi J, Patarapotikul J, Hananantachai H, Naka I, et al., CD36 polymorphism is
associated with protection from cerebral malaria. Am J Hum Genet, 72, 364-374, 2003.
131. Padua RA Plasmodium chabaudi: Genetics on resistance to chloroquine. Exp Parasitol, 52, 419426, 1981.
993
994
995
996
997
998
999
1000
1001
1002
1003
132. Pain A, Urban BC, Kai O, Casals-Pascual C, Shafi J, et al., A non-sense mutation in Cd36 gene is
associated with protection from severe malaria. Lancet, 357, 1502-1503, 2001.
133. Patel SS, King CL, Mgone CS, Kazura JW, Zimmerman PA, Glycophorin C (Gerbich antigen
blood group) and band 3 polymorphisms in two malaria holoendemic regions of Papua New
Guinea. Am J Hematol, 75, 1-5, 2004.
134. Patel SS, Mehlotra RK, Kastens W, Mgone CS, Kazura JW, et al., The association of the
glycophorin C exon 3 deletion with ovalocytosis and malaria susceptibility in the Wosera,
Papua New Guinea. Blood, 98, 3489-3491, 2001.
135. Paterson AH, Molecular dissection of quantitative traits: progress and prospects. Genome
Research, 5, 321-333, 1995.
136. Phenome Project at Jackson Laboratory. URL : http://www.informatics.jax.org/
1004
1005
1006
137. Pongponratn E, Riganti M, Punpoowong B, Aikawa M, Microvascular sequestration of parasitized
erythrocytes in human falciparum malaria: a pathological study. Am J Trop Med Hyg, 44, 168175, 1991.
34
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
1026
1027
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1038
1039
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
138. Pongponratn E, Turner GD, Day NP, Phu NH, Simpson JA, et al., An ultrastructural study of the
brain in fatal Plasmodium falciparum malaria. Am J Trop Med Hyg, 69, 345-359, 2003.
139. Polder TW, Jerusalem CR, Eling WM, Morphological characteristics of intracerebral arterioles in
clinical (Plasmodium falciparum) and experimental (Plasmodium berghei) cerebral malaria. J
Neurol Sci, 101, 35-46, 1991.
140. Quaye IK, Ekuban FA, Goka BQ, Adabayeri V, Kurtzhals JA, et al., Haptoglobin 1-1 is associated
with susceptibility to severe Plasmodium falciparum malaria. Trans R Soc Trop Med Hyg, 94,
216-219, 2000.
141. Ranque S, Safeukui I, Poudiougou B, Traore A, Keita M, et al., Familial aggregation of cerebral
malaria and severe malarial anemia. J Infect Dis, 191, 799-804, 2005,
142. Rihet P, Abel L, Traore Y, Traore-Leroux T, Aucan C, et al., Human malaria: segregation analysis
of blood infection levels in a suburban area and a rural area in Burkina Faso. Genet Epidemiol,
15, 435-450, 1998.
143. Rihet P, Traore Y, Abel L, Aucan C, Traore-Leroux T, et al., Malaria in humans: Plasmodium
falciparum blood infection levels are linked to chromosome 5q31-q33. Am J Hum Genet, 63,
498-505, 1998.
144. Rihet P, Flori L, Tall F, Traore AS, Fumoux F, Hemoglobin C is associated with reduced
Plasmodium falciparum parasitemia and low risk of mild malaria attack. Hum Mol Genet, 13,
1-6, 2004.
145. 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, 69, 129-136, 2001.
146. Roberts DJ, Craig AG, Berendt AR, Pinches R, Nash G, et al., Rapid switching to multiple
antigenic and adhesive phenotypes in malaria. Nature, 357, 689-692, 1992.
147. Rogier C, Commenges D, Trape JF, Evidence for an age-dependent pyrogenic threshold of
Plasmodium falciparum parasitemia in highly endemic populations. Am J Trop Med Hyg, 54,
613-619, 1996.
148. Rudin W, Eugster HP, Bordmann G, Bonato J, Muller M, Yamage M, & Ryffel B, Resistance to
cerebral malaria in tumor necrosis factor-alpha/beta-deficient mice is associated with a
reduction of intercellular adhesion molecule-1 up-regulation and T helper type 1 response.
Am J Pathol, 150, 257-266, 1997.
149. Ruwende C, Khoo SC, Snow RW, Yates SN, Kwiatkowski D, et al. , Natural selection of hemiand heterozygotes for G6PD deficiency in Africa by resistance to severe malaria. Nature, 376,
246-249, 1995.
150. Sabeti PC, Reich DE, Higgins JM, Levine HZ, Richter DJ, et al. (2002) Detecting recent positive
selection in the human genome from haplotype structure. Nature, 419, 832-837, 2002.
151. Sabeti P, Usen S, Farhadian S, Jallow M, Doherty T, et al., CD40L association with protection
from severe malaria. Genes Immun, 3, 286-291, 2002.
152. Sjoberg K, Lepers JP, Raharimalala L, Larsson A, Olerup O, et al. Genetic regulation of human
anti-malarial antibodies in twins. PNAS, 89, 2101-2104, 1992.
153. Singh, J, Ramakrishnan SP, Prakash S & Bhatnagar VN,. Studies on Plasmodium berghei. Vincke
and Lips, 1948. XX. A physiological change observed in a sulphadiazine resistant strain. Indian
J Malariol, 8, 301-307, 1954.
1050
1051
1052
1053
154. Snounou G, Gruner AC, Muller-Graf CD, Mazier D, Renia L, The Plasmodium sporozoite
survives RTS, S vaccination. Trends Parasitol, 21, 456-461, 2005.
155. Stevenson M, Lemieux S, Skamene E, Genetic control of resistance to murine malaria. J Cell
Biochem, 24, 91-102, 1984.
35
1054
1055
1056
1057
1058
1059
1060
1061
1062
156. Stevenson MM, Lyanga JJ, Skamene E. Murine malaria: genetic control of resistance to
Plasmodium chabaudi. Infect Immun, 38, 80-88, 1982.
157. Stirnadel HA, Beck HP, Alpers MP, Smith TA Heritability and segregation analysis of immune
responses to specific malaria antigens in Papua New Guinea. Genet Epidemiol, 17, 16-34, 1999.
158. Stirnadel HA, Beck HP, Alpers MP, Smith TA. Genetic analysis of IgG subclass responses against
RESA and MSP2 of Plasmodium falciparum in adults in Papua New Guinea. Epidemiol Infect,
124, 153-162, 2000.
159. Taylor-Robinson, A.W. Regulation of immunity to malaria: valuable lessons learned from murine
models. Parasitol Today, 11, 334-342, 1995.
1063
1064
1065
1066
1067
1068
1069
1070
160. Terrenato L, Shrestha S, Dixit KA, Luzzatto L, Modiano G, et al. Decreased malaria morbidity in
the Tharu people compared to sympatric populations in Nepal. Ann Trop Med Parasitol, 82, 111, 1988.
161. Thompson GR. Malaria and stress in relation to heamoglobin S and C. Br Med J, 2, 976-978,
1963.
162. Threadgill D.W., Hunt K.W., Williams R.W. Genetic dissection of complex and
quantitative traits: from fantasy to reality via a community effort. Mammalian Genome,
16, 344-355, 2002.
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
163. Tian JH, Miller LH, Kaslow DC, Ahlers J, Good MF, et al. Genetic regulation of protective
immune response in congenic strains of mice vaccinated with a subunit malaria vaccine. J
Immunol, 157, 1176-1183, 2002.
164. Tishkoff SA, Varkonyi R, Cahinhinan N, Abbes S, Argyropoulos G, et al. Haplotype diversity and
linkage disequilibrium at human G6PD: recent origin of alleles that confer malarial resistance.
Science, 293, 455-462, 2001.
165. Ubalee R, Suzuki F, Kikuchi M, Tasanor O, Wattanagoon Y, et al. Strong association of a tumor
necrosis factor-alpha promoter allele with cerebral malaria in Myanmar. Tissue Antigens, 58,
407-410, 2001.
166. Valdar W, Flint J, Mott R. Simulating the collaborative cross: power of quantitative trait
loci detection and mapping resolution in large sets of recombinant inbred strains of
mice. Genetics, 172, 1783-1797, 2006.
1083
1084
167. Vidal M.S., Malo D., Vogan K., Skamene E. & Gros P. Natural resistance to infection with
intracellular parasites: isolation of candidate for Bcg. Cell, 73, 469-485, 1993.
1085
1086
1087
1088
1089
168.
Voelkerding KV. et al. Next-generation sequencing: from basic research to diagnostics. Clin
Chem. 55: 641-58, 2009.
169. Walley AJ, Aucan C, Kwiatkowski D, Hill AV. Interleukin-1 gene cluster polymorphisms and
susceptibility to clinical malaria in a Gambian case-control study. Eur J Hum Genet, 12, 132138, 2004.
1090
1091
170. Walliker, D., Carter, R. & Sanderson, A. Genetic studies on Plasmodium: Recombination between
enzyme markers. Parasitology, 70, 19-24, 1975.
1092
1093
1094
1095
1096
171. Wang X, Paigen B. Quantitative trait loci and candidate genes regulating HDL cholesterol: a
murine chromosome map. Arterioscler Thromb Vasc Biol, 22, 1390-1401, 2002.
172. Wang X, Ishimori N, Korstanje R, Rollins J, Paigen B. Identifying novel genes for
atherosclerosis through mouse-human comparative genetics. Am J Hum Genet, 77, 1-15,
2005.
1097
1098
1099
173. Wang M., Lemon W.J., Liu G., Wang Y., Iraqi F.A., Malkinson A.M. & You M. Fine mapping
and identification of candidate pulmonary adenoma susceptibility 1 genes using advanced
intercross lines. Cancer Res, 63, 3317-3324, 2003.
36
1100
1101
174. Warhurst D.C. & Killick-Kendrick R. Spontaneous resistance to chloroquine in a strain of rodent
malaria (Plasmodium berghei yoelii). Nature, 213, 1048-1049, 1967.
1102
1103
1104
1105
1106
1107
1108
1109
1110
1111
1112
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
1123
175. Wattavidanage J, Carter R, Perera KL, Munasingha A, Bandara S, et al. TNFalpha*2 marks high
risk of severe disease during Plasmodium falciparum malaria and other infections in Sri
Lankans. Clin Exp Immunol, 115, 350-355, 1999.
176. Weiss WR, Good MF, Hollingdale MR, Miller LH, Berzofsky JA Genetic control of immunity to
Plasmodium yoelii sporozoites. J Immunol, 143, 4263-4266, 1989.
177. Wheeler D.L., Barrett T., Benson D.A., Bryant S.H., Canese K., Chetvernin V., Church D.M.,
DiCuccio M., Edgar R., Federhen S., Geer L.Y., Helmberg W., Kapustin Y., Kenton D.L.,
Khovayko O., Lipman D.J., Madden T.L., Maglott D.R., Ostell J., Pruitt K.D., Schuler G.D.,
Schriml L.M., Sequeira E., Sherry S.T., Sirotkin K., Souvorov A., Starchenko G., Suzek T.O.,
Tatusov R., Tatusova T.A., Wagner L. & Yaschenko E. Database resources of the National
Center for Biotechnology Information. Nucleic Acids Res, 34, D173-D180, 2006.
179. Williams TN, Wambua S, Uyoga S, Macharia A, Mwacharo JK, et al. Both heterozygous and
homozygous alpha+ thalassemias protect against severe and fatal Plasmodium falciparum
malaria on the coast of Kenya. Blood, 106, 368-371, 2005.
180. Wilson JN, Rockett K, Jallow M, Pinder M, Sisay-Joof F, et al. Analysis of IL10 haplotypic
associations with severe malaria. Genes Immun, 6, 462-466., 2005.
181. Wiltshire T., Pletcher M.J., Batalov S., Barnes S.W., Tarantino L.M. et al. Genome-wide
single-nucleotide polymorphism analysis defines haplotypes patterns in mouse. PNAS,
100, 3380-3385, 2003.
1124
1125
1126
1127
1128
182. World Health Organization. Division of Control of Tropical Diseases. Severe and complicated
malaria. Trans R Soc Trop Med Hyg, 84 Suppl 2, 1-65.
183. Wunderlich, F., Mossmann, H., Helwig, M. & Schillinger, G. (1988). Resistance to Plasmodium
chabaudi in B10 mice: influence of the H-2 complex and testosterone. Infect Immun, 56, 24002406, 1990.
1129
1130
1131
1132
184. Yoeli M. Upmanis R.S. & Most H. Drug-resistance transfer among rodent plasmodia. 1.
Acquisition of resistance to pyrimethamine by a drug-sensitive strain of Plasmodium berghei in
the course of its concomitant development with a pyrimethamine-resistant P. vinckei strain.
Parasitology, 69, 173-178, 1969.
178. WHO report world_malaria_report 2012: http://www.who.int/malaria/publications
1133
1134
1135
1136
1137
1138
1139
37
1140
1141
Figure Legends:
1142
1143
Figure 1. Immunological mechanisms directed against asexual blood stages and genes located within
1144
chromosome 5q31-q33 (red) and chromosome 6p21.3 (green).
1145
1146
Figure 2. Mapping of the H2 region and candidate genes Tnf and Lta on mouse chromosome 17 (the
1147
design of mouse chromosome 17 was taken from the mouse genome resources at the Ensembl site).
1148
1149
Figure 3. Location of the 5q31-q33 region on chromosome 5 linked to the control parasitic levels in
1150
human malaria. Candidate genes and their homologous locations in the mouse genome are shown on
1151
the right side of the figure. The figure was drawn with the chromosome design and information based
1152
on the NCBI (177) and Ensembl (22) data.
1153
1154
1155
38
1156
1157
Table 1. Loci reported at least once to show linkage or association with P. falciparum malaria.
Locus
Functiona
Location
Type of study
Phenotype
HbA1
HbA2
HbB
A
16pterp13.3
11p15.5
Case-control
Severe malaria
Mild malaria
Severe malaria
Mild malaria
Antibody levels
A
Case-control
Family-based
(linkage and
association)
References
(7,108,179)
(2,3,4,8,67,107,
112,123,144)
ABO
A
9q34
Case-control
Severe malaria
(43,89)
G6PD
A
Xq28
Case-control
Severe malaria
(56,149)
SCLC4A1
HP
ICAM1
CD36
CR1
PECAM1
NOS2A
A
A
B
B
B
B
C
17q21-q22
16q22
19p13
7q11.2
1q32
17q23
17cen-q11.2
Case-control
Case-control
Case-control
Case-control
Case-control
Case-control
Case-control
(9,54,133,134)
(14,40)
(18,42,81)
(6,132,130)
(20)
(75)
(24,25,69,80,82,1
24)
TNF
C
6p21.3
Case-control
Family-based
Severe malaria
Severe malaria
Severe malaria
Severe malaria
Severe malaria
Severe malaria
Severe malaria
Mild malaria
Severe malaria
Mild malaria
(linkage and
association)
NCR3
MALSb
C
6p21.3
Family-based
(5,46,76,165,
175)
Mild malaria
(36)
Mild malaria
(45,74)
Severe malaria
Severe malaria
Severe malaria
Severe malaria
Antibody levels
Parasitemia
(67)
(67)
(115)
(125)
(95)
(44,53,143)
Severe malaria
Severe malaria
Severe malaria
Severe malaria
(169)
(151)
(15)
(77)
Severe malaria
(29,78)
(linkage and
association)
C
6p21.3
Family-based
(genetic linkage)
HLA-B
HLA-DR
IL12B
IL13
IL4
PFBIb
C
C
C
C
C
C
6p21.3
6p21.3
5q33
5q31
5q31
5q31-q33
Case-control
Case-control
Case-control
Case-control
Case-control
Family-based
(linkage and
association)
IL1B
CD40L
IFNRA1
IFNGR1
C
C
C
C
2q14
Xq26
21q22.1
6q23-q24
Case-control
Case-control
Case-control
Case-control
Family-based
(linkage and
association)
IFN
C
12q14
Case-control
Family-based
(linkage and
39
association)
FCGR2A
C
C
1q31-q32
1q21-q23
Case-control
Case-control
TLR4
MBL2
C
C
9q32-q33
10q11-q21
Case-control
Case-control
IL10
Severe malaria
Severe malaria
Antibody levels
Severe malaria
Severe malaria
Parasitemia
(180)
(35,123,129)
(151)
(23,96,114)
1158
1159
a
1160
and C, respectively.
1161
b
1162
Infection levels and mild Malaria Susceptibility, respectively. Accession numbers are 248310 (PFBI)
1163
and 609148 (MALS).
Loci involved in red blood cell physiology, cytoadherence, and immune responses are encoded A, B,
Loci are referenced in OMIM database. PFBI and MALS mean Plasmodium falciparum Blood
1164
1165
40
1166
1167
Table 2. Susceptibility of inbred mouse strains to malaria after infection with 3 Plasmodium subspecies
1168
(83,136).
Inbred mouse strain
P. chabaudi
P. berghei
P. yoelii
A/HeJ
Susceptible
Susceptible
ND
A/J
Susceptible
Susceptible
ND
AKR/J
Susceptible
Resistant
ND
BALB/cJ
Susceptible
Susceptible
Susceptible
CBA/J
Resistant
Susceptible
ND
C3H/HeJ
Susceptible
Resistant
ND
C57BL/6J
Resistant
Susceptible
Susceptible
C57L/J
Resistant
Susceptible
ND
C58/J
ND
Resistant
ND
DBA/1J
Susceptible
Resistant
ND
DBA/2J
Resistant
Resistant
ND
SJL/J
Susceptible
ND
ND
ST/bJ
ND
Susceptible
ND
1169
1170
ND: not done
1171
41
1172
Table 3. Mouse genetic loci controlling parasitemia (P) and cerebral malaria (CM). aP.c. stands
1173
for Plasmodium chabaudi, bP.c.c. stands for Plasmodium chabaudi chabaudi and cP.b. stands for
1174
Plasmodium berghei.
Phenotype
Locus
Chromosome
P
char1
9
Murine parasite
Candidate
genesa
Trf, Rbp
Gene
identifiedb
?
P. c. adami DS
GypA,
?
P. c.c. ASb
Cspg3
P. c. adami DSa
P. yoelii 17XL
P
char2
8
P.c.c. 54X
P
char3
17
P. c. adami DS
Tnf, Lta
?
P.c.c. 54X
Pklr269A
P
char4
3
P. c.c. AS
Lef1, Cfi
P
char5
5
P.c.c. 54X
Ache, Epo
?
P
char6
5
P.c.c. 54X
Cora1,
?
Ncf1
1175
a
1176
b
P
char7
17
P.c.c. 54X
C3, Il25
?
P
char8
11
P.c.c. 54X
Il4, Csf2
?
CM
unnamed
18
P. b. ANKAc
Csf1r, Cd14
?
CM
cmsc
17
P. b. ANKA
Tnf, Lta
?
CM
berr1
1
P. b. ANKA
Tgfb2
?
CM
berr2
11
P. b. ANKA
-
?
CM
berr3
9
P. b. ANKA
Ccr5, Rbp
?
CM
berr4
4
P. b. ANKA
Tlr4, C8b
?
Genes initially proposed as candidates
Genes
the
variation
of
which
42
alters
the
phenotype
1177
Box 1. Web sites to find mouse genome resources
http://www.ncbi.nlm.nih.gov/genome/guide/mouse/
http://www.ensembl.org/Mus_musculus/
http://genome.ucsc.edu/
http://www.informatics.jax.org/
43
NKp3
+
NK
TNF
IL-18
NKp3
IFN
+
NK
Mo
IRF1
Elimination of the
parasite:
phagocytosis
and/or
Th
Eo
ILA
+
Th
IL-4
IFN
IL-10
HLA-DR
ILTh
Parasitized
erythrocyte
IL-
IL-3
GMIL-5
IL-4 IL-6
IL-5 IL-10
IL- IL-9
+
IgG1
cytophilic IgG IgG2
B FcJRIIA
IgG3
noncytophilic
1
IgG2
IgG4
1
Human Chromosome 5
1