ArticlePDF Available

Linkage maps from multiple genetic crosses and loci linked to growth-related virulent phenotype in Plasmodium yoelii

Authors:

Abstract and Figures

Plasmodium yoelii is an excellent model for studying malaria pathogenesis that is often intractable to investigate using human parasites; however, genetic studies of the parasite have been hindered by lack of genome-wide linkage resources. Here, we performed 14 genetic crosses between three pairs of P. yoelii clones/subspecies, isolated 75 independent recombinant progeny from the crosses, and constructed a high-resolution linkage map for this parasite. Microsatellite genotypes from the progeny formed 14 linkage groups belonging to the 14 parasite chromosomes, allowing assignment of sequence contigs to chromosomes. Growth-related virulent phenotypes from 25 progeny of one of the crosses were significantly associated with a major locus on chromosome 13 and with two secondary loci on chromosomes 7 and 10. The chromosome 10 and 13 loci are both linked to day 5 parasitemia, and their effects on parasite growth rate are independent but additive. The locus on chromosome 7 is associated with day 10 parasitemia. The chromosome 13 locus spans ~220 kb of DNA containing 51 predicted genes, including the P. yoelii erythrocyte binding ligand, in which a C741Y substitution in the R6 domain is implicated in the change of growth rate. Similarly, the chromosome 10 locus spans ~234 kb with 71 candidate genes, containing a member of the 235-kDa rhoptry proteins (Py235) that can bind to the erythrocyte surface membrane. Atypical virulent phenotypes among the progeny were also observed. This study provides critical tools and information for genetic investigations of virulence and biology of P. yoelii.
Content may be subject to copyright.
Linkage maps from multiple genetic crosses and
loci linked to growth-related virulent phenotype
in Plasmodium yoelii
Jian Li
a,b,1
, Sittiporn Pattaradilokrat
b,1
, Feng Zhu
a,1
, Hongying Jiang
b
, Shengfa Liu
a
, Lingxian Hong
a
, Yong Fu
c
, Lily Koo
d
,
Wenyue Xu
c
, Weiqing Pan
e
, Jane M. Carlton
f
, Osamu Kaneko
g
, Richard Carter
h
, John C. Wootton
i
, and Xin-zhuan Su
b,2
a
State Key Laboratory of Stress Cell Biology, School of Life Sciences, Xiamen University, Xiamen, Fujian 361005, Peoples Republic of China;
b
Laboratory of
Malaria and Vector Research, and
d
Research Technologies Branch, National Institute of Allergy and Infectious Diseases, National Institutes of Health,
Bethesda, MD 20892;
c
Department of Pathogenic Biology, Third Military Medical University, Chongqing 400038, Peoples Republic of China;
e
Department of
Pathogen Biology, Second Military Medical University, Shanghai 200433, Peoples Republic of China;
f
Department of Medical Parasitology, Langone Medical
Center, New York University, New York, NY 10010;
g
Department of Protozoology, Institute of Tropical Medicine and the Global Center of Excellence Program,
Nagasaki University, Nagasaki 852-8523, Japan;
h
Division of Biological Sciences, Institute of Cell, Animal and Population Biology, Ashworth Laboratories,
University of Edinburgh, Edinburgh EH9 3JT, United Kingdom; and
i
Computational Biology Branch, National Center for Biotechnology Information, National
Library of Medicine, National Institutes of Health, Bethesda, MD 20894
Edited by Thomas E. Wellems, National Institutes of Health, Bethesda, MD, and approved May 27, 2011 (received for review February 9, 2011)
Plasmodium yoelii is an excellent model for studying malaria path-
ogenesis that is often intractable to investigate using human para-
sites; however, genetic studies of the parasite have been hindered
by lack of genome-wide linkage resources. Here, we performed 14
genetic crosses between three pairs of P. yoelii clones/subspecies,
isolated 75 independent recombinant progeny from the crosses,
and constructed a high-resolution linkage map for this parasite.
Microsatellite genotypes from the progeny formed 14 linkage
groups belonging to the 14 parasite chromosomes, allowing assign-
ment of sequence contigs to chromosomes. Growth-related virulent
phenotypes from 25 progeny of one of the crosses were signi-
cantly associated with a major locus on chromosome 13 and with
two secondary loci on chromosomes 7 and 10. The chromosome 10
and 13 loci are both linked to day 5 parasitemia, and their effects on
parasite growth rate are independent but additive. The locus on
chromosome 7 is associated with day 10 parasitemia. The chromo-
some 13 locus spans 220 kb of DNA containing 51 predicted genes,
including the P. yoelii erythrocyte binding ligand, in which a C741Y
substitution in the R6 domain is implicated in the change of growth
rate. Similarly, the chromosome 10 locus spans 234 kb with 71
candidate genes, containing a member of the 235-kDa rhoptry pro-
teins (Py235) that can bind to the erythrocyte surface membrane.
Atypical virulent phenotypes among the progeny were also ob-
served. This study provides critical tools and information for genetic
investigations of virulence and biology of P. yoelii.
genetic mapping
|
inheritance
|
crossover
|
rodent
The rodent malaria parasite Plasmodium yoelii is an important
model for studying malaria biology and pathogenesis. Because
a malaria disease phenotype represents the outcome of the host-
parasite interaction, the use of inbred mice to control host genetic
background variation is critical for studying the inuence of par-
asite virulent factors on a disease phenotype. Many genetically
distinct (or similar) strains of P. yoelii and subspecies exhibiting
a wide range of variations in growth rate and pathogenicity in their
rodent hosts are available, which can be explored for studying
disease and/or growth phenotypes. Compared with Plasmodium
falciparum and Plasmodium chabaudi chabaudi (13), however,
genetic studies in P. yoelii have been limited (2, 4, 5), partly be-
cause of the lack of genetic markers and well-characterized phe-
notypes. Recently, hundreds of polymorphic microsatellite (MS)
markers have been developed from the P. yoelii genome (6), set-
ting the stage for development of genome-wide genetic maps for
this parasite. Additionally, a strategy called linkage group selec-
tion (LGS) was developed to map the determinants affecting se-
lectable rodent malaria traits (2, 7). Indeed, a C713R substitution
in the gene encoding the P. y. yoelii erythrocyte binding ligand
(PyEBL) was recently linked to parasite growth rate and virulence
using the LGS technique, although other determinants are likely
to play a role (8, 9). [Note: There are three subspecies of P. yoelii
(P. yoelii yoelii, P. yoelii nigeriensis, and P. yoelii killicki) and two
subspecies of P. chabaudi (P. chabaudi chabaudi and P. chabaudi
adami). P. yoelii is used here to refer generally to P. yoelii lines and
subspecies; subspecies and lines will be specied in the text when
necessary.] For mapping genes affecting complex traits or pheno-
types that cannot be selected, however, evaluation of phenotypes
from individual progeny of genetic crosses is necessary. De-
velopment of a genetic map and collection of genetic cross progeny
with differences in disease phenotypes will provide important tools
for studying such malaria disease phenotypes in detail.
Although the P. y. yoelii genome was the rst rodent malaria
parasite genome sequenced, the assembly is still fragmented be-
cause of the currently low coverage and lack of a genetic map to
guide the assembly (10). Recently, a rodent malaria syntenic map
was constructed based on genomic sequences from three rodent
malaria parasites (P. y. yoelii,P. c. chabaudi,andPlasmodium
berghei) and sequence synteny to the genomic sequence of P. fal-
ciparum (11); however, thousands of sequence gaps still exist, and
many contigs are yet to be assigned to their proper chromosomes.
Increasing sequence coverage may close additional gaps, but de-
velopment of physical and genetic maps will be necessary for
assigning all the contigs to chromosomal positions and for as-
sembling the chromosomes completely.
Here, we have performed 14 individual genetic crosses using six
parasite lines/subspecies, cloned 75 independent recombinant
progeny from the crosses, genotyped 82 recombinant progeny from
genetic crosses of four parental pairs (including 7 progeny from a
previous P. y. yoelii YM ×P. y. yoelii A/C cross) (12) with hundreds
of MS markers, and developed a high-resolution linkage map. We
also identied three genetic loci, including the gene encoding
PyEBL, linked to quantitative growth-related virulent pheno-
Author contributions: X.-z.S. designed research; J.L., S.P., F.Z., Y.F., and L.K. performed
research; S.L., L.H., W.X., W.P., and O.K. contributed new reagents/analytic tools; J.L., S.P.,
H.J., L.K., J.M.C., R.C., J.C.W., and X.-z.S. analyzed data; and J.L., S.P., J.M.C., O.K., R.C.,
J.C.W., and X.-z.S. wrote the paper.
The authors declare no conict of interest.
This article is a PNAS Direct Submission.
Freely available online through the PNAS open access option.
1
J.L., S.P., and F.Z. contributed equally to this work.
2
To whom correspondence should be addressed. E-mail: xsu@niaid.nih.gov.
See Author Summary on page 12575.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1102261108/-/DCSupplemental.
E374E382
|
PNAS
|
August 2, 2011
|
vol. 108
|
no. 31 www.pnas.org/cgi/doi/10.1073/pnas.1102261108
types (GRVPs) using trait data from individual nonselected
progeny clones.
Results
Frequencies of Clonal Infection and Independent Recombinant
Progeny. We performed 14 independent genetic crosses using the
six P. yoelii lines or subspecies (P. y. yoelii 17XNL ×P. y. nigeriensis
N67, P. y. yoelii BY265 ×P. y. nigeriensis NSM, and P. y. yoelii YM ×
P. y. yoelii 33X). For simplicity, these parasite strains are re-
ferred to as 17XNL, N67, BY265, NSM, YM, 33X, and A/C (for
P. y. yoelii A/C), respectively. Mice (n= 2,326) were injected with
0.61.5 infected red blood cells (iRBCs) obtained from mice that
were infected with sporozoites derived from mosquitoes fed on
blood samples containing both parental parasites. Of the 2,326
mice, 889 (38.2%) had parasites in their blood 79 d postinjection
(Table 1). DNA samples from the infected mice were genotyped
with 45 MS markers, 18 for the BY265 ×NSM cross, 20 for the
YM ×33X cross, and 28 for the 17XNL ×N67 cross, with some
markers typed in more than one cross. For example, Py2699 was
used to type progeny from the 17XNL ×N67, YM ×33X, and
BY265 ×NSM crosses to verify clonal infections and to identify
recombinant progeny (Datasets S1 and S2). Because the parasite
has a haploid genome in the mouse host, the progeny that carry
only one of the parental alleles at all MS loci typed for each cross
are considered clonal. Among the 889 infected mice, 488 (54.9%)
were found to have clonal infections, including 75 (8.4%) in-
dependent recombinant progeny with unique genotypes and 248
(27.9%) having parasites with parental genotypes (Table 1). We
also genotyped 7 progeny from the YM ×A/C cross performed at
the University of Edinburgh, bringing the total number of prog-
eny typed with MSs to 82.
MS Polymorphism and Genetic Variations Between Parasite Strains.
The 82 independent recombinant progeny identied from the
initial MS typing were further analyzed with additional genome-
wide MS markers. To determine which MSs are polymorphic
between a particular parental pair of the crosses, we rst geno-
typed DNA samples from the parental parasites [N67, BY265,
and 17XNL were genotyped previously (6)] with 591 MS markers
(Dataset S1). A panel of polymorphic MS markers for each cross
was selected for typing progeny from the crosses after comparing
PCR product sizes between the parents. Eventually, 485, 499, 339,
and 182 MS markers were shown to be polymorphic between the
parental pairs of BY265 ×NSM, 17XNL ×N67, YM ×33X, and
YM ×A/C crosses, respectively, and were used to type the
progeny of the crosses. Four hundred seventy-seven MS markers
produced genotypes from the 32 progeny of the BY265 ×NSM
cross, 486 MS markers produced genotypes from the 25 progeny
of the 17XNL ×N67 cross, 330 MS markers had genotypes from
the 18 progeny of the YM ×33X cross, and 178 MS markers had
genotypes from the 7 progeny of the YM ×A/C cross, generating
33,939 MS genotypes with a genotype-calling rate of 96.0%
(Dataset S2 and Table S1).
More than 82% of the 591 MSs were polymorphic between the
parental pairs of BY265 ×NSM and 17XNL ×N67 (Table S1), in-
dicating highly diverse genomes of the P. yoelii strains or subspecies
from different geographic origins. In contrast, 70% and 43% of
the MS markers that were shown to be polymorphic among seven
isolates previously (6) were monomorphic between the parents of
the YM ×A/C and YM ×33X crosses, respectively. Parasites YM,
17XL, and 17XNL emerged as very similar, with fewer than 25
polymorphic MS markers, which reects their common origin from
17X and is consistent with a previous study based on amplied
fragment length polymorphism (AFLP) (13) (Figs. S1 and S2).
These results showed close relationships and potentially shared
chromosomes or chromosomal segments between the 17X, YM, A/
C, and 33X. Because A/C is a progeny clone from a genetic cross
between P. y. yoelii 17XA (17XA) that was derived from the isolates
17X and 33X, which originated from the same locality in the
Central African Republic as parasite 17X (Fig. S2), it is not sur-
prising to see a close genetic relationship or shared chromosome
segments between17X/YM and 33X.
Comparative Genetic Linkage Maps and Estimates of Genetic
Distances. We rst constructed three individual genetic maps
from the three crosses (excluding YM ×A/C, which has only 7
progeny) and estimated the genetic distances from the crosses
using 37 markers that were physically mapped to specic chro-
mosomes previously (11, 1416) (Fig. S3 and Table S2). These
markers therefore anchored specic linkage groups to their re-
spective chromosomes and were useful in resolving some linkage
groups into independent chromosomes. Although the genome-
wide distances from the 17XNL ×N67 and BY265 ×NSM crosses
were similar [431.2 and 473.7 centimorgan (cM), respectively], the
total genetic distance from the YM ×33X cross was approximately
double those from the other two crosses (807.0 cM) (Table 2).
Despite these differences in estimates of genetic distance, all three
crosses gave relatively even marker intervals (in cM) on all linkage
groups, and the marker orders on each linkage group of the three
crosses were essentially the same. We therefore combined the
genotypes from the 82 progeny of the four crosses to construct
a composite linkage map (Fig. 1 and Fig. S3). The resulting map
totaled 579.2 cM, with markers quite evenly distributed across
each of 14 linkage groups (no interval was >10 cM) (Fig. 1 and
Table 2). Using the estimated genome size of 23 Mb (10), we
obtained an average genome-wide unit recombination rate of 39.7
kb/cM (25.2 cM/Mb) for P. yoelii.
Shared Chromosomal Segments and Lack of Signicant Segregation
Bias. Analysis of the marker inheritance patterns also revealed that
some entire chromosomes and long chromosomal segments were
shared among the YM, A/C, and 33X parasites, explaining the
smaller numbers of polymorphic MS markers in the crosses of
these parasites. Chromosomes 2, 3, 5, 7, and 10 [our linkage
groups are numbered to match the chromosome numbers in the
syntenic map of Kooij et al. (11)] and parts of many other chro-
mosomes were monomorphic between YM and A/C (Table 2 and
Table 1. Genetic crosses of P. yoelii performed and numbers of recombinant progeny obtained during this study
Crosses
Crosses
performed
Mice
injected
Mice
infected
% mice
infected
Clonal
progeny
Parental
clones
Recom
progeny
% recom
progeny No. IRP % IRP
17XNL ×N67 4 501 145 28.9 122 84 38 26.2 25 17.2
YM ×33X 3 209 77 36.8 58 10 48 62.3 18 23.4
BY265 ×NSM 7 1,616 667 41.3 308 154 154 23.1 32 4.8
Total 14 2,326 889 38.6 488 248 240 27 75 8.4
Clonal progeny, numbers of progeny that are clonal after being typed with MS markers; Mice infected, numbers of mice infected with parasites; % Mice
infected, percentage of mice infected with parasites; Mice injected, numbers of mice injected with diluted blood/parasites; No. IRP, numbers of independent
recombinant progeny; % IRP, percentage of independent recombinant progeny from total infected mice; Recom progeny, numbers of recombinant progeny;
% Recom progeny, percentage of recombinant progeny from total infected mice.
Li et al. PNAS
|
August 2, 2011
|
vol. 108
|
no. 31
|
E375
MICROBIOLOGY PNAS PLUS
Dataset S2); similarly, markers from chromosomes 2 and 10 and
parts of chromosomes 3, 5, 6, 7, and 12 were monomorphic in the
YM ×33X cross. Because YM, which derives from 17X, and 33X
were from two independent wild isolates of the same location
(Fig. S2B), the results suggested that either 17X or 33X was a
progeny of a cross involving an ancestor of the other parent that
occurred in the wild before they were isolated.
We also examined the possibility of segregation bias by plot-
ting the ratios of parental genotypes along each of the 14 chro-
mosomes (Fig. 2A). No statistically signicant distortion was
Table 2. Informative MS markers, crossover counts, and genetic distances from crosses of different parental combinations
YM ×A/C BY265 ×NSM 17XNL ×N67 YM ×33X All crosses
(7 progeny) (32 progeny) (25 progeny) (18 progeny) (82 progeny)
Chr MS CO G. Dis, cM MS CO G. Dis, cM MS CO G. Dis, cM MS CO G. Dis, cM MS CO G. Dis, cM
16131 6 20.9 31 5 21.8 27 3 18.0 34 15 21.1
20011 7 23.0 12 1 4.2 1 0 12 8 14.5
30015 7 23.2 17 3 13.1 4 1 5.9 17 11 19.2
4126 24 8 27.0 26 9 38.7 20 9 58.8 27 32 42.6
51028 5 17.4 25 1 5.0 10 2 12.0 30 8 15.7
6114 20 7 26.8 23 3 13.2 16 13 88.0 28 27 38.3
71024 3 10.4 24 9 38.0 16 11 70.1 25 23 33.4
8304 38 12 40.2 36 13 56.6 35 20 123.6 41 49 64.9
9232 39 15 50.5 39 11 47.5 38 9 55.1 47 37 54.8
10 1 0 31 10 33.4 30 7 30.6 1 0 33 17 31.7
11 8 4 48 17 61.1 49 12 60.1 41 17 106.0 52 50 70.0
12 12 0 39 12 43.8 40 5 22.3 11 0 44 17 40.7
13 50 4 56 16 54.9 59 12 52.7 56 27 180.7 67 59 77.3
14 12 6 49 12 41.1 51 6 27.4 37 13 88.8 53 37 55.0
UA 11 24 24 17 29
Total 178 31 477 137 473.7 486 97 431.2 330 125 807.0 539 390 579.2
Chr, chromosome; CO, crossover counts; G. Dis, genetic distances in cM; MS, numbers of polymorphic microsatellites; UA, unassigned.
12
1.8
1.8
1.8
3.6
1.8
1.8
1.8
Py652
Py378
Py171
Py1184
Py2407
Py1308
Py2673
Py575
3
1.8
1.8
1.8
3.7
1.8
1.8
3.6
2.8
Py2428
Py1582
Py82
Py1476
Py845'
Py1176
Py1865
Py1428
Py2221
4
1.2
3.6
3.7
5.5
1.3
1.3
1.3
9.4
3.9
1.4
5.0
5.0
Py345
Py2246
Py141
Py1578
Py2390
Py1758
Py529
Py642
Py1077
Py2160
Py994
Py1404
Py2181
5
1.4
3.0
2.1
5.8
3.5
Py1214
Py1406
Py1490
Py2484
Py2133
Py1062
67
3.7
1.8
1.4
2.1
2.1
4.2
5.6
4.2
1.4
1.4
4.2
1.4
Py174
Py516
Py1377
Py742'
Py1277
Py1938
Py1238
Py2961
Py1798
Py156
Py1325'
Py943
Py1957
8
1.4
4.1
1.4
2.5
3.8
1.2
1.2
1.2
1.3
1.2
1.2
1.3
1.5
6.2
1.3
2.6
2.5
2.5
3.8
2.5
2.6
4.4
Py187
Py924
Py759
Py95
Py658
Py984
Py348
Py59
Py734
Py917
Py498
Py455
Py1983
Py2768
Py2578
Py2370
Py1687
Py1448
Py1441
Py1703
Py1156'
Py1156
Py1638
Py1341
Py1250
Py2669
Py2032
Py2032'
2.1
5.9
1.9
1.2
1.9
9.4
2.5
2.5
1.3
2.5
3.0
5.8
5.8
2.8
2.8
1.3
1.5
5.8
1.4
5.7
1.5
1.4
2.8
1.5
1.5
1.4
2.8
1.5
1.4
Py1432
Py194'
Py1318
Py109
Py928
Py1213
Py649
Py1418
Py1425
Py549
Py595
Py555
Py1653
Py2083
Py607
Py1257
Py1821
Py563
Py1249
Py643
Py2355
Py2361
Py289
Py1527
Py2033
11
1.8
1.9
3.6
5.6
1.8
1.9
1.0
6.5
1.8
3.9
1.9
Py2721
Py1839
Py332'‘
Py189
Py280
Py56
Py2588
Py317
Py1829
Py2266
Py453
Py1235
10
9
8.6
2.8
1.4
2.7
1.4
4.2
1.5
1.4
1.3
1.2
1.2
2.5
1.3
2.6
3.8
1.4
3.9
1.3
1.2
5.1
1.3
1.2
1.2
Py1076
Py1885
Py1063
Py1924
Py765
Py191
Py1751
Py1041
Py1659
Py2785
Py1776
Py1033
Py762
Py1003
Py1632
Py181
Py548
Py148
Py1481
Py806
Py982
Py1761
Py719
Py2857
12 13
1.5
1.3
3.8
2.5
2.5
1.3
2.6
1.3
1.2
6.6
1.2
2.5
1.3
1.3
1.2
6.7
1.3
2.5
7.9
5.2
3.8
3.9
1.2
1.2
1.2
1.3
1.2
2.5
1.2
2.5
1.3
Py193
Py523
Py1091
Py1959
Py2423
Py1970
Py213
Py2096
Py935
Py1281
Py2123
Py2513
Py373
Py952
Py2609
Py1084
Py799
Py2443
Py346
Py2486
Py372
Py400
Py1555
Py876
Py753
Py2546
Py1401
Py809
Py1803
Py35
P
y
2092
Py2826
14
1.4
2.8
2.9
2.8
1.4
1.2
4.2
1.5
1.6
1.4
Py1710
Py1618
Py1494
Py170
Py277
Py864
Py619
Py2426
Py1355
Py1516
Py451
1.8
1.3
1.2
1.4
3.8
9.4
1.3
3.7
1.6
1.3
1.4
1.7
1.7
6.8
Py1539
Py265
Py1121
Py2311
Py871
Py1685
Py2034
Py1315
Py1457
Py2699
Py565
Py740
Py436
Py152
Py1080
1.5
2.4
7.1
2.2
2.2
2.3
6.9
4.6
4.6
2.3
4.6
Py1331
Py2354
Py98
Py1442
Py519
Py283
Py216
Py1753
Py437
Py2008
Py2006
Py2260
1.5
1.5
4.2
1.5
1.4
1.3
4.4
1.4
3.9
8.3
3.2
1.7
2.5
1.2
1.3
2.6
3.8
7.9
1.4
Py114
Py893
Py133'
Py981
Py810
Py810'
Py2195
Py995
Py1718
Py1841
Py1447
Py2962
Py1272
Py1495
Py398
Py2422
Py266
Py815
Py2265
Py517
Fig. 1. A composite P. yoelii genetic map constructed from genetic crosses of four parental combinations. Genetic distances between MS markers on the 14
chromosomes of 82 progeny were calculated using Mapmaker/Exp3.0. The linkage groups corresponding to the 14 chromosomes in the syntenic map (11) are
marked from 1 to 14. The numbers between two ticks on the left side of the vertical lines are genetic distances in centimorgans, and the names of the MS
markers are on the right side of vertical lines. Additional names and clusters of the MS markers can be found in Fig. S3. Note that chromosome numbering in
P. yoelii and other rodent Plasmodium sp. is different from that of P. falciparum. The thick bars on chromosomes 10 and 13 mark the two loci linked to a GRVP.
E376
|
www.pnas.org/cgi/doi/10.1073/pnas.1102261108 Li et al.
present (Fig. 2B); however, the largest (nonsignicant) deviation
from the expected 1:1 ratio occurred at regions of chromosomes
7 and 13 close to loci associated with growth phenotypes (see
below), which may reect contributions of these loci to parasite
tness. The relative lack of segregation distortion is an advantage
for genetic mapping in P. yoelii and contrasts with the signicantly
skewed inheritance reported for a few chromosomal regions in
genetic crosses of P. falciparum (1, 3) and Toxoplasma gondii (17).
Assignment of Orphan Contigs to Chromosomes. The linkage maps
were compared with the composite synteny chromosome maps of
three rodent malaria parasites (11). Except for four contigs, the
linkage maps placed the contigs with MS markers in the same
order of those in the synteny map (Dataset S2). A contig with MS
Py2653 was assigned to chromosome 7 in the synteny map, but the
inheritance pattern of the MS in the progeny of the BY265×NSM
and 17XNL ×N67 crosses matched those of Py1080 on chromo-
some 6. Similarly, the contig with Py1484 initially placed on
chromosome 7 of the synteny map was assigned to chromosome
12, and the contig with Py517initially assigned to chromosome 13
matched MS Py353on chromosome 14 (Dataset S2). Finally, the
positions of the contigs containing Py1803 and Py35 on chromo-
some 13 were reversed in the synteny map.
Our linkage maps also assigned 28 orphan contigs (159 kb)
that were not assigned to the synteny maps previously to 13 of
the 14 linkage groups (those in gray background in Dataset S2).
These genetic maps and the placement of the contigs in linkage
groups will greatly facilitate the assembly and completion of the
genome sequence.
Mapping Genes Contributing to GRVPs. We next applied the linkage
map to analyze the GRVP in the 17XNL ×N67 cross. Between the
two parasites used as parents in the genetic crosses, N67 grows
faster than 17XNL, showing a strong early burst of rapid growth
associated with greater virulence. The difference in parasitemia
between N67 and 17XNL was the largest at day 5 postinjection of
1×10
5
iRBCs, when N67 parasitemia reached 2060% but the
17XNL parasitemia was 110% (Fig. 3 AC). Accordingly, we
counted day 5 parasitemia in replicate mice injected with the 25
progeny from the 17XNL ×N67 cross (Fig. 3Dand Dataset S3).
The 25 progeny from the 17XNL ×N67 cross produced relatively
consistent growth phenotypes in replicate mice and could be
largely classied into either a fast- or slow-growth phenotype,
representing the phenotypes of N67 or 17XNL, respectively;
however, some progeny clones had unstable phenotypes (e.g., in-
consistently producing either slow or fast growth in replicate
infections of isogenic mice) (Dataset S3). For example, three of
the eight mice infected with progeny G007#3 had a fast-growth
phenotype at day 5, whereas the remaining ve had a slow-growth
phenotype (a third mouse reached 32% parasitemia at day 7).
Quantitative trait loci (QTL) linkage analysis was performed
using the parasitemia data from the 25 progeny of the 17XNL ×
N67 cross. We rst used two distinct conservative statistical phe-
notype-genotype association strategies (nonparametric Wilcoxon
rank statistics and mutual information) that make no assumptions
about pedigree, given that we had high marker density but rela-
tively few progeny and that the parents can also be included to
234 Chromosome
15
68
7910
11 12 13 14
0.2
0
0.4
0.6
0.8
1.0
Allele Ratio
A
234
15
68
7910 11 12 13 14
0100 200
Marker number 400
300
0
1
2
3
1
2
3
Log(1/P)
B
Fig. 2. Plots of parental genotype inheritance among the progeny of
the three genetic crosses. (A) Ratios of alleles from one parent of each of the
crosses were calculated and plotted. Green, ratios of YM alleles over the
total alleles from the progeny of the YM ×33X cross; red, ratios of 17XNL
alleles over the total alleles from the progeny of the 17XNL ×N67 cross; and
black, ratios of BY265 alleles over the total alleles from the progeny of the
BY265 ×NSM cross. (B) Log(1/P) values are plotted upward (gray) for dis-
tortion toward 17XNL and downward (blue) for distortion toward N67.
Dotted lines represent the P<0.01 genome-wide signicance threshold
estimated using a Bonferroni correction based on the number of genetic
intervals in the genome map. The largest (nonsignicant) deviation from the
expected 1:1 ratio occurred at regions of chromosomes 7 and 13 close to the
loci associated with the GRVP (Fig. 3).
A
010 15 20 25
5
40
0
80
Parasitemia
Day post infection
B
40
0
80
010 15 20 25
5
Parasitemia
Day post infection
0
20
40
60
Parasitemia
Progeny
D
0
5
10
15
05
10 15
20
Daypostinfection
Ratio parasitmia
(N67/17XNL)
C
40
20
0
60
80
0
51015
20
E
Day post infection
%Pa
rasitemia
N67
N67C
Fig. 3. Measurements of growth rate (parasitemia) of the parents of the 17XNL ×N67 cross. Mean parasitemia and SEs of the 17XNL parasite (A), mean
parasitemia and SEs of the N67 parasite (B), ratios of parasitemia N67/17XNL at days 216 postinfection (C), and day 5 parasitemia from the parents and
progeny of the 17XNL ×N67 cross (D) are shown. In D,therst two bars from the left are the parents 17XNL and N67, respectively. SEs were from at least four
mice. (E) Mean parasitemia with SEs from mice infected with N67 and N67C parasites. A total of 15 and 5 C57BL/6 mice were injected with 1 ×10
5
N67C and
N67 parasites, respectively. All the mice infected with N67C died at day 7, whereas the majority of the mice infected with N67 were still alive at day 17 when
experiments were stopped.
Li et al. PNAS
|
August 2, 2011
|
vol. 108
|
no. 31
|
E377
MICROBIOLOGY PNAS PLUS
increase the sample size using these methods (Materials and
Methods). The result identied a major peak on chromosome 13
containing the PyEBL gene (genome-wide P<1×10
5
)and
a minor peak on chromosome 10 (P<1×10
2
) (Fig. 4A). As
a distinct phenotype, day 10 parasitemia was analyzed and found to
be weakly associated with a locus on chromosome 7 (P= 0.05 after
1,000 permutations) (Fig. 4B). We also used R/qtl (18) to identify
loci linked to the GRVP; again, the chromosome 13 and chro-
mosome 10 loci were the two major peaks with logarithm of odds
scores of 3.4 and 3.2, respectively (Fig. 4C). There was also a peak
on chromosome 7, but it was not statistically signicant. The
chromosome 13 locus spans a DNA segment of 220 kb (between
MS markers Py2123 and Py2609) and contains 28 P. yoelii contigs
and 51 predicted genes, including the gene encoding PyEBL
(Table S3), and the chromosome 10 locus covers 234 kb (between
MS markers Py280 and Py1597) containing 28 P. yoelii contigs and
71 predicted genes (Table S4), including a gene encoding a mem-
ber of the 235-kDa rhoptry proteins that can bind to the erythro-
cyte surface membrane (19).
We also investigated the possibility of interactions among
genome-wide markers, particularly between the two loci linked
to day 5 parasitemia in the 17XNL ×N67 cross. Using standard
interval mapping (the expectation and maximization method) in
R/qtl, we found no statistically signicant epistasis among ge-
nome-wide markers (Fig. 5A), including no interaction between
the loci on chromosome 10 and chromosome 13 (P= 0.31);
instead, the three notable two-locus effects were all additive,
namely, chromosome 10/chromosome 13 (P= 0.005), chromo-
some 7/chromosome 10 (P= 0.012), and chromosome 13/chro-
mosome 13 (P= 0.011) (Fig. 5A). The two interacting loci on
chromosome 13 are 18 cM apart and are inversely additive, with
one containing pyebl and one at the beginning of the chromo-
some that has a minimum QTL peak (Fig. 4C). The chromosome
10 and chromosome 13 loci are estimated to explain 23% and
25% of the phenotype variance (totaling 70% if combined;
P<0.001), respectively. From the mutual information method,
the percentages explained by the chromosome 10 and chromo-
some 13 loci to the association with the day 5 parasitemia are
23% and 37%, respectively. Given the relatively broad bands of
uncertainty on estimates of relative contributions and the small
sample size of 25 progeny, these values from R/qtl and mutual
information can be considered to be in good agreement. Indeed,
the progeny carrying N67 alleles at both chromosome 13 and
chromosome 10 loci had the highest level of parasitemia at day 5
compared with those carrying one or both of the 17XNL alleles
(Fig. 5B).
Unique C741Y Substitution in PyEBL Is Likely Associated with the
GRVP in the N67 Background. The loci on chromosomes 7 and 10
have not been previously reported and require further in-
vestigation to identify the gene(s) playing a role in parasite
growth; however, the locus on chromosome 13 contains the pyebl
gene, which was previously linked to a GRVP both by LGS in the
YM ×33X cross and by genetic manipulation implicating a single
C713R substitution (position 713 in the YM sequence) in the
PyEBL R6 domain as the crucial determinant (8, 9). Our mapping
of the primary locus to chromosome 13 strongly suggests that
PyEBL again plays a role in the GRVP in N67. We therefore
sequenced the N67 PyEBL gene to determine whether this
C713R substitution is also present in the N67 parasite. We found
39 amino acid substitutions and two indels between 17XNL and
N67 (Fig. S4); however, the C713R substitution seen in YM does
not exist in N67. Instead, a C741Y substitution (with the other
changes) was present in the R6 domain (Fig. S4). Interestingly,
a parasite submitted to MR4 (http://www.mr4.org/) under the
name of P. y. yoelii 33X(Pr3) [for simplicity, 33X(Pr3) will be
used] has an identical PyEBL sequence and a very similar geno-
mic background to that of N67, except for the C741Y substitution
in N67 (Fig. S4 and Table S5). Among 21 MS markers typed, only
a single MS had different alleles between N67 and 33X(Pr3),
whereas all the 21 MSs had different alleles among 33X and 33X
(Pr3)/N67 (Table S5). The results suggest that N67 and 33X(Pr3)
are closely related or isogenic and that the 33X(Pr3) in our hands
was not the original parasite derived from 33X. We therefore
designated this parasite N67C for having a 741C in PyEBL. Al-
though N67C grows slightly slower than N67 in the early infection,
it is more virulent than N67 because all the N67C-infected mice
died at day 7, whereas mice infected with N67 died at approxi-
mately day 15 after a decline in parasitemia on day 7 (Fig. 3 A
and E). The difference in the GRVP between N67 and N67C is
0
1
2
3
1
23 465
8
79
10
11 13
12 14
C
LOD score
0100 400
300
200
0
0.5
1.0
1.5
Log(1/P)
Chromosome/marker position
0.3
0.2
MI
0.1
123 4567 8 91011 12 13 14
0.0
B
A
Fig. 4. Genetic loci linked to the GRVP. (A) Plots of mutual information scores and MS markers across the 14 P. yoelii chromosomes using day 5 parasitemia
from the 25 progeny and the parents of the 17XNL ×N67 cross. MI, mutual information. (B) Plots of log(1/P) and MS markers using day 10 parasitemia. (C) Plot
of logarithm of odds scores using R/qtl (Materials and Methods) after natural logarithm transformation of the day 5 parasitemia data so that the dataset is
normally distributed. LOD, logarithm of odds. Multiple calculation methods, including maximum likelihood and extended HaleyKnott regression, were
evaluated; all produced the essentially the same results. The horizontal dashed lines indicate signicant levels at P= 0.001 (A), P=0.05(B), and P=0.05(C).
E378
|
www.pnas.org/cgi/doi/10.1073/pnas.1102261108 Li et al.
therefore likely caused by the C741Y substitution. Thus, two
different single mutational substitutions associated with high day
5 parasitemia and with destabilizing effects focused on the R6
domain appear to have originated independently in the distinct
PyEBL sequence backgrounds of 17XNL/YM and N67/N67C.
Discussion
Genetic crosses have been performed with several malaria parasite
species, including P. falciparum,P. c. chabaudi, and P. y. yoelii (3, 8,
2023). The process of cloning, genotyping, and identifying
recombinant progeny has been the most time-consuming pro-
cedure in developing a linkage map and mapping malaria traits.
The numbers of progeny isolated from Plasmodium crosses have
been relatively small; for example, the P. falciparum Dd2 ×HB3
and GB4 ×7G8 crosses have 35 and 33 progeny, respectively (3,
22, 24). Because cloning rodent malaria parasites requires inject-
ing a single parasite into a large number of mice, only 7 progeny
from the YM ×A/C cross were previously obtained (8). In this
study, we obtained 488 cloned parasites from 2,326 mice and
identied 75 independent recombinant progeny from multiple
genetic crosses after genotyping the parasites with hundreds of MS
markers. This study describes the rst P. yoelii linkage map, which
represents a signicant advance in malaria parasite genetics.
Performing multiple genetic crosses using different parents
allowed comparison of inheritance bias and recombination fre-
quency among crosses. Our study revealed a somewhat greater
genome-wide genetic distance (centimorgans) in the YM ×33X
cross than in the other crosses, possibly reecting fewer progeny
from this cross or some unknown genetic factors among P. yoelii
strains in control of meiotic recombination frequencies. Never-
theless, in all these crosses, the recombination rates were sub-
stantially less than those previously reported for P. falciparum
and P. c. chabaudi. We found relatively small inheritance bias, in
contrast to previous ndings with P. falciparum crosses (1, 3).
Inheritance patterns of chromosomes 7 and 13 were marginally
skewed, possibly reecting alleles that may promote parasite
survival (e.g., a chromosome 13 locus strongly linked to para-
sitemia at day 5, a chromosome 7 locus weakly linked to para-
sitemia at day 10).
The linkage map provides a solid framework for future im-
provement of the P. yoelii genome assembly. The P. yoelii genome
is currently estimated at 23 Mb but is highly fragmented in 5,687
contigs (10). Although the syntenic map based on sequences from
three rodent parasites and the genome sequence of P. falciparum
had assigned 2,400 P. y. yoelii contigs (total length of 15.2 Mb)
to putative chromosomes, there are still many contigs that cannot
be assigned (10, 11). We were able to assign 28 orphan contigs
(159 kb) to 13 chromosomes. Additional orphan contigs can
now be assigned to chromosomes or linkage groups if poly-
morphic markers from the contigs can be typed on the progeny of
the genetic crosses. Our genetic map has 14 linkage groups, in
accordance with the 14 nuclear chromosomes of all Plasmodium
species investigated, and the marker order in these linkage groups
matched well with those in the chromosomal synteny maps. The
linkage maps developed in this study provide essential tools for
complete genome assembly of P. yoelii.
The same chromosome 13 locus containing the gene encoding
PyEBL was identied by both the LGS (8) and our classic QTL
mapping based on the phenotypes of individual nonselected
progeny clones. Our study also identied a unique mutation
and a potential different mechanism affecting parasite GRVPs.
These mutations appear to have originated independently in the
diverged PyEBL sequence backgrounds of P. y. yoelii 17X and P. y.
nigeriensis N67, which differ by 39 amino acid substitutions and
two indels. In theory, other genes in the chromosome 13 locus
cannot be excluded as candidates contributing to the GRVP;
however, the differences in the PyEBL are likely to be the primary
determinant for the differences in the GRVP between N67 and
17XNL based on the two previous studies (8, 9). Furthermore,
N67 and N67C have an almost identical genomic background and
the same PyEBL sequence, except for the substitution at C741Y.
The two different single mutational substitutions associated with
higher early parasitemia, C713R in YM (8, 9) and C741Y in N67
identied here, are predicted to have similar disruptive effects on
the R6 domain. Inspection of the homologous EBA-175 R6/KIX
domain crystal structure (25) suggests that both the C713R and
the C741Y substitutions break two different strongly conserved
disulde bonds (Fig. S4)and act indirectly, partly through sol-
vation effects on the primary hydrophobic core and a hydrophobic
groove in the dimer interface. The disruption of the disulde
bonds appears to correlate with higher early peak parasitemia
before day 7. N67 (with a 741Y) grows slightly faster than N67C
(741C) at days 25, and YM/17XL (with a 713R) also grows faster
than 17XNL (713C). The C741Y substitution may partly explain
the reduced virulence of N67 compared with N67C, but the drop
in parasitemia after day 7 in N67 could be caused by other un-
C1317XNL
C1017XNL C13N67
C10N67
C1317XNL
C1017XNL
C13N67
C13N67
0
10
20
30
40
50
60
% mean parasitemia
10 915
B
P<0.01
P<0.01
P=0.02
915
Chromosome
Chromosome
12
3456 7 8910 11 1213 14
1
2
3
4
5
67
8
9
10
11
12
13
14
A
0
1
2
3
0
2
4
6
LODfLODi
Fig. 5. Interaction and additive effect between the loci on the parasite
chromosomes. (A) Loci with additive effects but no signicant interactions
genome-wide were detected. The upper left triangle displays interactions
among pairwise loci. The lower right triangle displays additive effects. LODf,
logarithm of odds additive effect; LODi, logarithm of odds interaction. Sig-
nals of additive effect between markers on chromosomes 1013, chromo-
somes 710, and two loci on chromosome 13 (red arrows) were detected.
The color scale bar shows LOD scores, with those on the left and right cor-
responding to LODi and LODf values, respectively. The two arrows indicate
cutoff LOD scores from 1,000 permutations (LODi = 4.5; LODf = 5.9; P<0.05).
(B) Relationship of parasitemia and different combinations of PyEBL alleles.
Progeny carrying N67 PyEBL alleles at the mapped loci also have signicantly
higher parasitemia than those carrying 17XNL alleles (unpaired ttest). The
numbers within each bar are the numbers of progeny in each group.
Li et al. PNAS
|
August 2, 2011
|
vol. 108
|
no. 31
|
E379
MICROBIOLOGY PNAS PLUS
known factors too, or it could represent a separate phenotype.
The mechanism of how the change in amino acid in a merozoite
protein affecting parasite early growth remains unknown.
The presence of loci other than the chromosome 13 locus was
implicated in GRVPs in the previous studies (8, 9); however, the
exact locations of the additional determinants were unknown.
Our study now points to two loci on chromosomes 10 and 7 that
are linked to day 5 and day 10 parasitemia, respectively. In-
terestingly, a copy of the Py235 rhoptry protein (PY06636) that
has been implicated in RBC invasion is present in the chromo-
some 10 locus (19, 26). Members of the Py235 rhoptry proteins
have been suggested to be potential factors that may be re-
sponsible for the difference in virulence (27). Moreover, our
results show independent and additive effects for the QTL on
chromosomes 13 and 10, with no evidence for signicant epistasis
genome-wide. Three additive effects were found, although the
additive effect from the two loci on chromosome 13 could be
artifact because of no obvious QTL peak at the locus on the
chromosome end and the small sample size. Further studies are
necessary to identify or conrm genes in the other two loci that
contribute to the GRVP. Interestingly, genetic loci contributing
to subtle but quantiable strain-specic differences in prolifer-
ation rates of P. falciparum parasite Dd2 (faster growth) and HB3
during in vitro cultivation, which were attributed to a decreased
cycle time and increased merozoite production and invasion rates,
have been mapped recently, including a major genetic effect on
chromosome 12 containing 165 candidate genes (28).
Recombinant progeny with atypical growth phenotypes
were obtained after genetic recombination. Display of either fast
or slow growth in a single progeny was observed, which was likely
controlled through gene expression regulation or combinational
effects of multiple genes, or possibly through subtle uncontrolled
environmental effects inuencing the mice. Progeny with an in-
termediate growth rate or growth rate higher than that of the
parents were also obtained. These transgressive cases indicate
that clones that are more virulent could be generated in the wild
through genetic recombination in sexual crosses of parasites as well
as by de novo mutations. Indeed, recombinant progeny more vir-
ulent than parental clones have been reported in T. gondii, sug-
gesting that sexual recombination can be a powerful force driving
the natural evolution of virulence (29). Alternatively, deleterious
mutations can accumulate in clonally maintained organisms, and
recombinationmay effectively remove these deleterious mutations
from some progeny, restoring tness and resulting in progeny with
phenotypes greater than those observed in the parental lines.
Genetic mapping of rodent malaria traits has been largely based
on LGS, which does not require cloning and typing individual
progeny (7). In this procedure, a phenotype-specic selection
pressure is applied to the uncloned progeny of a genetic cross
between two parents with different relevant phenotypes. Selected
and unselected progeny are analyzed using genome-wide quanti-
tative genetic markers. Under selection, the sensitive allele of the
target gene will be removed, leading to a reduced allele frequency
or selection valleyat the locus carrying the resistantgene (7).
LGS has been successfully used to map loci affecting drug
resistance, immunity, and difference in growth rate/virulence (2,
7, 8, 30), which has established LGS as a convenient approach
for mapping selectable malaria traits. Cloning and evaluating in-
dividual progeny from genetic crosses now enables a powerful
genetic mapping approach that can be applied to screen non-
selectable parasite phenotypes or complex traits that are not
amenable to LGS methods. The progeny obtained in this study not
only provide powerful and durable resources for genetic studies
but also allow investigation of inheritance bias and recombina-
tion parameters, such as hotspots. Moreover, the markers ordered
on the 14 chromosomes now provide a chromosomal framework
for improved assembly of the parasite genome sequence. As with
studies that have been done using P. falciparum genetic crosses (22,
3135), the progeny and genotypes we describe here can be used to
map multiple segregating genetic determinants in this mouse
model, which, in turn, will provide important information for
studying human malaria.
Materials and Methods
Parasites, DNA Sequencing, and MS Typing. The origins of the P. yoelii lines
and the relationships of the cloned lines are summarized in Fig. S2 and Table
S6. Parasites 17XNL, N67, BY265, N67C [under P. y. yoelii 33X(Pr3)], and NSM
were obtained from MR4 (http://www.mr4.org/) or were described pre-
viously (36). YM, 33X, A/C, and P. y. yoelii 17X(A) were obtained from frozen
stocks of the University of Edinburgh (13). YM is a lethal parasite derived
from the nonlethal uncloned isolate 17X following removal of a stabilate
from liquid nitrogen storage in the early 1970s (12, 37); it also has the same
genome and GRVP as 17XL (13). NSM is a meoquine-resistant parasite se-
lected from P. yoelii NS, a parasite line emerged from an isolate of P. berghei
from Katanga, Belgian Congo, in the early 1970s; however, our genotyping
showed that the N67, NSM, and P. yoelii NS in our hands had essentially the
same genome (Dataset S1) (36). A/C is a progeny from a cross of 33X and
17XA, a pyrimethamine-resistant line that is genetically distinct from clones
17XNL, 17XL, and YM but originated from the same isolate 17X (12, 13) (Fig.
S2). Mosquitoes were from a colony of Anopheles stephensi maintained at
the Laboratory of Malaria and Vector Research, National Institutes of Health
(NIH), and a colony of A. stephensi maintained at the Third Military Medical
University of China. Female inbred strain C57BL/6 and outbred CD-1 and
Kunming mice, aged 68 wk, were used in the experiments. All animal
procedures were performed in accordance with animal study protocol LMVR
11E approved by the National Institute of Allergy and Infectious Diseases
Animal Care and Use Committee (NIH) and with the approved protocols of
the Third Military Medical University or Xiamen University in China.
DNA samples were extracted from 20 to 100 μL of heparinized tail blood
from P. yoelii-infected mice using a High Pure PCR Template preparation kit
(Roche). For DNA sequencing of pyebl, oligonucleotide primers 5-CCTCC-
TGTTGCATAGTAGTATTGAT-3 and 5-TTTGATGAACCAAATGCATAGA-3, corre-
sponding to positions 1,4071,431 and 4,2324,211 of the P. y. yoelii 17XNL
contig MALPY01471 (GenBank accession no. AABL01001466) (10), were syn-
thesized to amplify a full-length coding region. PCR reactions wereperformed
in a 50-μL volume consisting of 100 ng of genomic DNA in 1×Ultra-high delity
Accuzyme mix (Bioline). PCR conditions were as follows: one cycle at 94 °C for
5 min; followed by 35 cycles at 94 °C for 30 s, 55 °C for 1 min, and 68 °C for
4 min; and a nal extension at 68 °C for 5 min.The products were treated with
shrimp alkaline phosphatase and exonuclease I before sequencing (38). Se-
quencing reactions were performed in triplicate from different amplication
tubes using an ABI Prism BigDye Terminator ready-to-use reaction kit (Applied
Biosystems). Sequencing primers were described previously (8).
MSs used in this study have been described previously (6, 36). Briey, PCR
products without any labeling procedures were separated using capillary
electrophoresis performed in a QIAxcel machine (QIAGEN) according to the
manufacturers instructions. Genotypes from genetic cross progeny were
scored by matching the PCR product sizes with those from the parents. The
genetic distances between the parasites were calculated using methods
described (36).
P. yoelii Genetic Crosses. Three combinations of genetically distinct clones of
P. yoelii were chosen to produce genetic crosses. Two genetic crosses of (i)
N67 (lethal) ×17XNL (nonlethal) and (ii) YM (lethal) ×33X (nonlethal) were
performed at the NIH laboratory, whereas the genetic cross between BY265
and NSM was conducted at the Third Military Medical University and Xiamen
University of China. The genetic cross between YM and A/C was performed
in David Wallikers laboratory at the University of Edinburgh, and seven
recombinant clones have been cryopreserved since 1976 (12). Because we
could only handle a limited number of mice at a time, we cloned progeny
from multiple individual crosses of each parental pair. The crosses were also
performed with different phenotypes in mind. For example, we were in-
terested in differences in GRVPs in the N67 ×17XNL and YM ×33X crosses
and drug resistance in the BY265 ×NSM cross (NSM is more resistant to
meoquine and chloroquine).
The experimental procedures for the production of genetic crosses in
P. yoelii have been described previously (8, 39). Briey, inbred C57BL/6 or
Kunming outbred (BY265×NSM cross) female mice were coinfected with two
parasites (parents) to be crossed. Parasitemia (percent of parasitized eryth-
rocytes from at least 1,000 cells) from the mice was monitored daily by mi-
croscopic examination of Giemsa-stained thin blood smears. On day 4
postinfection, each infected mouse with male and female gametocytes (av-
E380
|
www.pnas.org/cgi/doi/10.1073/pnas.1102261108 Li et al.
erage parasitemia 1030%) was anesthetized and fed to 50100 female
mosquitoes per infected mouse. The mouse was then euthanized while under
anesthesia. Seventeen days after feeding, sporozoites were harvested from
infected mosquitoes, and 1 ×10
5
to 5 ×10
5
sporozoites were injected i.p. into
female CD-1 mice to obtain blood-stage parasites representing uncloned
progenyof the genetic cross. When blood-stage parasites were microscop-
ically detectable (0.11% parasitemia), blood samples were drawn from the
infected mice and diluted to 0.61.5 iRBCs per 100 μL (inoculum size) before
injection i.v. into recipient mice. Five to nine days after injection, mice were
monitored for the presence of blood-stage malaria parasites. A small volume
(2050 μL) of mouse tail blood was withdrawn and used for parasite genomic
DNA extraction and MS genotyping. For screening of recombinant progeny,
DNA from cross progeny was typed with a panel of MS markers distributed
on chromosomes 114 (Dataset S1). Recombinant progeny were considered
clonal when single MS alleles were found in the MS loci. To improve the
recovery of independent progeny, the parasites were cloned as early as
possible to prevent overgrowthof some fast-growing parasites. Moreover,
the parental input ratio was adjusted appropriately for each particular cross
based on the different growth rates of each pair of parents.
Development of Genetic Linkage Maps. We constructed a genetic linkage map
from the MS genotypes and segregation patterns in the progeny of each
individual cross using Mapmaker/Exp 3.0 (40) to order the markers and es-
timate genetic distances between them. We used 37 MS markers within the
contigs that had been physically assigned to chromosomes previously (11,
1416) as anchors for assigning linkage groups to chromosomes (Dataset S1
and Table S2).
GRVP and Linkage Analysis. To map the determinant(s) that affect the GRVP
between N67 and 17XNL, we evaluated the growth rates of the progeny from
the cross between 17XNL and N67 in female C57BL/6 mice (48 mice per single
parasite clone). Each mouse was injected i.v. with an inoculum containing
1×10
5
iRBCs. Parasitemias were monitored daily, as measured by microscopic
examination of Giemsa-stained thin tail blood smears, and recorded in Excel.
QTL Analysis and Statistics. For genome-wide scans of the day 5 parasitemia,
the association between categorical trait value vectors and genotype vectors
was evaluated using Shannons mutual information (41) with the genome-
wide P<0.001 threshold (Fig. 4A, dashed horizontal line) estimated using
a noncentral χ
2
test conrmed by simulation (41). The simulation estimated
the null distribution of mutual information under nonassociation using
1,000 permutations and the method described by Nelson and OBrien (42).
The day 10 parasitemia was treated as a continuous trait, and associated
QTLs were mapped using a ttest statistic and 1,000 permutations. We also
ran scans based on a nonparametric Wilcoxon test for all phenotypes to
provide independent support for the associations, which gave essentially the
same results.
QTL mapping was also conducted using the R/qtl library in R2.12.2 software
(18). Day 5 parasitemia from the 25 progeny of the N67 and 17XNL cross was
natural-log transformed to obtain a normally distributed dataset. One
thousand permutations with a 5% threshold (P= 0.05) were performed
using standard interval mapping (the expectation and maximization
method). A single QTL genome scan was done rst; a 2D scan and, nally,
two QTL genome scans were then performed because of two similarly sig-
nicant loci on chromosome 10 and chromosome 13. In principle, the 2D and
two QTL scans could identify additional QTL and two-locus epistatic inter-
actions if they are present.
Statistical Analysis of Inheritance Bias. We measured the statistical signi-
cance of deviations from the expected 1:1 segregation ratio. For each marker,
we computed the Kendall τ-rank association statistic between the observed
parents and progeny genotype vector and each of the two vectors, simu-
lating complete bias (i.e., with all progeny assigned as one parental type or
the other). The S-plus function cor.test was used with the method kendall,
where probabilities of dependence were based on the normal approxima-
tion as described by Prokhorov (43).
ACKNOWLEDGMENTS. We thank Drs. Karl W. Broman and Na Li for advice
on QTL analysis and National Institute of Allergy and Infectious Diseases
intramural editor Brenda Rae Marshall for assistance. This work was
supported by grants from the National Basic Research Program of China, 973
Program (Grant 2007CB513103), the Science Planning Program of Fujian
Province (Grant 2010J1008), and the 111 Project of Education of China
(Grant B06016) as well as by the Intramural Research Program of the Division
of Intramural Research, National Institute of Allergy and Infectious Diseases,
National Institutes of Health. J.C.W. was supported by the Intramural
Program of the National Center for Biotechnology Information, National
Library of Medicine, National Institutes of Health.
1. Su X, et al. (1999) A genetic map and recombination parameters of the human
malaria parasite Plasmodium falciparum. Science 286:13511353.
2. Martinelli A, et al. (2005) A genetic approach to the de novo identication of targets
of strain-specic immunity in malaria parasites. Proc Natl Acad Sci USA 102:814819.
3. Hayton K, et al. (2008) Erythrocyte binding protein PfRH5 polymorphisms determine
species-specic pathways of Plasmodium falciparum invasion. Cell Host Microbe 4:
4051.
4. Su X, Hayton K, Wellems TE (2007) Genetic linkage and association analyses for trait
mapping in Plasmodium falciparum. Nat Rev Genet 8:497506.
5. Martinelli A, et al. (2005) An AFLP-based genetic linkage map of Plasmodium
chabaudi chabaudi. Malar J 4:11.
6. Li J, et al. (2009) Hundreds of microsatellites for genotyping Plasmodium yoelii
parasites. Mol Biochem Parasitol 166:153158.
7. Culleton R, Martinelli A, Hunt P, Carter R (2005) Linkage group selection: Rapid gene
discovery in malaria parasites. Genome Res 15:9297.
8. Pattaradilokrat S, Culleton RL, Cheesman SJ, Carter R (2009) Gene encoding
erythrocyte binding ligand linked to blood stage multiplication rate phenotype in
Plasmodium yoelii yoelii. Proc Natl Acad Sci USA 106:71617166.
9. Otsuki H, et al. (2009) Single amino acid substitution in Plasmodium yoelii erythrocyte
ligand determines its localization and controls parasite virulence. Proc Natl Acad Sci
USA 106:71677172.
10. Carlton JM, et al. (2002) Genome sequence and comparative analysis of the model
rodent malaria parasite Plasmodium yoelii yoelii. Nature 419:512519.
11. Kooij TW, et al. (2005) A Plasmodium whole-genome synteny map: Indels and synteny
breakpoints as foci for species-specic genes. PLoS Pathog 1:e44.
12. Walliker D, Sanderson A, Yoeli M, Hargreaves BJ (1976) A genetic investigation of
virulence in a rodent malaria parasite. Parasitology 72:183194.
13. Pattaradilokrat S, Cheesman SJ, Carter R (2008) Congenicity and genetic polymor-
phism in cloned lines derived from a single isolate of a rodent malaria parasite. Mol
Biochem Parasitol 157:244247.
14. Uzureau P, Barale JC, Janse CJ, Waters AP, Breton CB (2004) Gene targeting demon-
strates that the Plasmodium berghei subtilisin PbSUB2 is essential for red cell invasion
and reveals spontaneous genetic recombination events. Cell Microbiol 6:6578.
15. Thompson J, et al. (2004) PTRAMP; a conserved Plasmodium thrombospondin-related
apical merozoite protein. Mol Biochem Parasitol 134:225232.
16. Janse CJ, Carlton JM, Walliker D, Waters AP (1994) Conserved location of genes on
polymorphic chromosomes of four species of malaria parasites. Mol Biochem Parasitol
68:285296.
17. Khan A, et al. (2005) Composite genome map and recombination parameters derived
from three archetypal lineages of Toxoplasma gondii. Nucleic Acids Res 33:29802992.
18. Broman KW, Wu H, Sen S, Churchill GA (2003) R/qtl: QTL mapping in experimental
crosses. Bioinformatics 19:889890.
19. Preiser PR, Jarra W, Capiod T, Snounou G (1999) A rhoptry-protein-associated
mechanism of clonal phenotypic variation in rodent malaria. Nature 398:618622.
20. Walliker D, Carter R, Morgan S (1971) Genetic recombination in malaria parasites.
Nature 232:561562.
21. Walliker D, et al. (1987) Genetic analysis of the human malaria parasite Plasmodium
falciparum. Science 236:16611666.
22. Wellems TE, et al. (1990) Chloroquine resistance not linked to mdr-like genes in
a Plasmodium falciparum cross. Nature 345:253255.
23. Grech K, et al. (2002) Numerous, robust genetic markers for Plasmodium chabaudi by
the method of amplied fragment length polymorphism. Mol Biochem Parasitol 123:
95104.
24. Su X, Kirkman LA, Fujioka H, Wellems TE (1997) Complex polymorphisms in an
approximately 330 kDa protein are linked to chloroquine-resistant P. falciparum in
Southeast Asia and Africa. Cell 91:593603.
25. Withers-Martinez C, et al. (2008) Malarial EBA-175 region VI crystallographic structure
reveals a KIX-like binding interface. J Mol Biol 375:773781.
26. Khan SM, Jarra W, Bayele H, Preiser PR (2001) Distribution and characterisation of the
235 kDa rhoptry multigene family within the genomes of virulent and avirulent lines
of Plasmodium yoelii. Mol Biochem Parasitol 114:197208.
27. Culleton R, Kaneko O (2010) Erythrocyte binding ligands in malaria parasites:
Intracellular trafcking and parasite virulence. Acta Trop 114:131137.
28. Reilly Ayala HB, Wacker MA, Siwo G, Ferdig MT (2010) Quantitative trait loci mapping
reveals candidate pathways regulating cell cycle duration in Plasmodium falciparum.
BMC Genomics 11:577.
29. Grigg ME, Bonnefoy S, Hehl AB, Suzuki Y, Boothroyd JC (2001) Success and virulence
in Toxoplasma as the result of sexual recombination between two distinct ancestries.
Science 294:161165.
30. Pattaradilokrat S, Cheesman SJ, Carter R (2007) Linkage group selection: Towards
identifying genes controlling strain specic protective immunity in malaria. PLoS ONE
2:e857.
31. Vaidya AB, et al. (1995) A genetic locus on Plasmodium falciparum chromosome 12
linked to a defect in mosquito-infectivity and male gametogenesis. Mol Biochem
Parasitol 69:6571.
Li et al. PNAS
|
August 2, 2011
|
vol. 108
|
no. 31
|
E381
MICROBIOLOGY PNAS PLUS
32. Ferdig MT, et al. (2004) Dissecting the loci of low-level quinine resistance in malaria
parasites. Mol Microbiol 52:985997.
33. Hayton K, Su XZ (2008) Drug resistance and genetic mapping in Plasmodium
falciparum. Curr Genet 54:223239.
34. Gonzales JM, et al. (2008) Regulatory hotspots in the malaria parasite genome dictate
transcriptional variation. PLoS Biol 6:e238.
35. Yuan J, et al. (2009) Genetic mapping of targets mediating differential chemical
phenotypes in Plasmodium falciparum. Nat Chem Biol 5:765771.
36. Li J, et al. (2007) Typing Plasmodium yoelii microsatellites using a simple and
affordable uorescent labeling method. Mol Biochem Parasitol 155:94102.
37. Yoeli M, Hargreaves B, Carter R, Walliker D (1975) Sudden increase in virulence in
a strain of Plasmodium berghei yoelii. Ann Trop Med Parasitol 69:173178.
38. Mu J, et al. (2007) Genome-wide variation and identication of vaccine targets in the
Plasmodium falciparum genome. Nat Genet 39:126130.
39. Pattaradilokrat S, Li J, Su XZ (2011) Protocol for production of a genetic cross of the
rodent malaria parasites. J Vis Exp, 10.3791/2365.
40. Lander E, et al. (1987) MAPMAKER: An interactive computer package for constructing
primary genetic linkage maps of experimental and natural populations. Genomics 1:
174181.
41. Dawy Z, et al. (2006) Gene mapping and marker clustering using Shannons mutual
information. IEEE/ACM Trans Comput Biol Bioinformatics 3:4756.
42. Nelson GW, OBrien SJ (2006) Using mutual information to measure the impact of
multiple genetic factors on AIDS. J Acquir Immune Dec Syndr 42:347354.
43. Prokhorov AV (2001) Kendall coefcient of rank correlation. Online Encyclopedia of
Mathematics, ed Hazewinkel M (Springer, CWI, Amsterdam).
E382
|
www.pnas.org/cgi/doi/10.1073/pnas.1102261108 Li et al.
... After performing eight independent crossing experiments, we obtained 35 uncloned progeny pools, with 25 selected with MFQ (20 mg/kg). DNA samples from the 25 MFQ treated and 10 nontreated progeny pools were genotyped with 190 polymorphic microsatellites (MS) 23 on 14 parasite chromosomes. All progeny from the untreated group carried BY265 alleles (Supplementary Data 1), indicating faster growth for BY265 without drug pressure. ...
... Detailed information for key materials and reagents, including parasite lines, antibodies and chemicals, used in this study is provided in Supplementary Data 4. P. yoelii strains NSM and BY265 have been described previously 23 . NSM is an uncloned parasite that exhibits lowlevel MFQ resistance that was selected from P. yoelii NS, a parasite line that emerged from an isolate of P. berghei from Katanga, Belgian Congo, in the early 1970s 57,58 . ...
... All mosquitoes were raised at 24°C and 75% humidity under a 12:12 light-dark illumination cycle and fad with 5% sucrose solution. The experimental procedures for producing genetic crosses in rodent malaria parasites have been described in detail previously 23,59 . Briefly, outbred ICR mice were co-infected with P. yoelii strains NSR and BY265 in a desired parental ratio according to the relative ability of the individual strain or line to produce gametocytes and oocysts 60 . ...
Article
Full-text available
Mutations in a Plasmodium de-ubiquitinase UBP1 have been linked to antimalarial drug resistance. However, the UBP1-mediated drug-resistant mechanism remains unknown. Through drug selection, genetic mapping, allelic exchange, and functional characterization, here we show that simultaneous mutations of two amino acids (I1560N and P2874T) in the Plasmodium yoelii UBP1 can mediate high-level resistance to mefloquine, lumefantrine, and piperaquine. Mechanistically, the double mutations are shown to impair UBP1 cytoplasmic aggregation and de-ubiquitinating activity, leading to increased ubiquitination levels and altered protein localization, from the parasite digestive vacuole to the plasma membrane, of the P. yoelii multidrug resistance transporter 1 (MDR1). The MDR1 on the plasma membrane enhances the efflux of substrates/drugs out of the parasite cytoplasm to confer multidrug resistance, which can be reversed by inhibition of MDR1 transport. This study reveals a previously unknown drug-resistant mechanism mediated by UBP1 through altered MDR1 localization and substrate transport direction in a mouse model, providing a new malaria treatment strategy.
... Our findings suggest that the loss of one of these three disulfide bonds in PyEBL region 6 alters the protein structure and as a consequence, PyEBL is not trafficked to the microneme, leading to enhanced parasite infectivity. Two groups have independently conducted linkage group selection analysis to identify genes that determine differences in the parasite multiplication rate of different P. yoelii strains [25,26]. Both groups identified the strongest association of the pyebl gene locus on chromosome 13 with the virulence of the parasites. ...
... In the first study [25], the lethal P. yoelii YM strain was crossed with the non-lethal 33XC strain and the substitution of the second Cys to Arg of PyEBL region 6 in the YM strain was identified as a determinant, which is identical to the substitution found in the 17XL strain [17]. The second study conducted a genetic cross between the lethal P. yoelii nigeriensis N67 and the non-lethal P. yoelii 17XNL strain and identified that the 6th Cys of PyEBL region 6 was substituted to Tyr in the N67 strain [26]. The single amino acid substitutions associated with higher virulence are predicted to have similar disruptive effects on PyEBL region 6, specifically, C726R in 17XL [17], C713R in YM [25], and C741Y in N67 [26]. ...
... The second study conducted a genetic cross between the lethal P. yoelii nigeriensis N67 and the non-lethal P. yoelii 17XNL strain and identified that the 6th Cys of PyEBL region 6 was substituted to Tyr in the N67 strain [26]. The single amino acid substitutions associated with higher virulence are predicted to have similar disruptive effects on PyEBL region 6, specifically, C726R in 17XL [17], C713R in YM [25], and C741Y in N67 [26]. These results agree with our observations that the disruption of a single disulfide bond in PyEBL region 6 dramatically changes the virulence of the P. yoelii parasites. ...
Article
Full-text available
Plasmodium malaria parasites use erythrocyte-binding-like (EBL) ligands to invade erythrocytes in their vertebrate host. EBLs are released from micronemes, which are secretory organelles located at the merozoite apical end and bind to erythrocyte surface receptors. Because of their essential nature, EBLs have been studied as vaccine candidates, such as the Plasmodium vivax Duffy binding protein. Previously, we showed through using the rodent malaria parasite Plasmodium yoelii that a single amino acid substitution within the EBL C-terminal Cys-rich domain (region 6) caused mislocalization of this molecule and resulted in alteration of the infection course and virulence between the non-lethal 17X and lethal 17XL strains. In the present study, we generated a panel of transgenic P. yoelii lines in which seven of the eight conserved Cys residues in EBL region 6 were independently substituted to Ala residues to observe the consequence of these substitutions with respect to EBL localization, the infection course, and virulence. Five out of seven transgenic lines showed EBL mislocalizations and higher parasitemias. Among them, three showed increased virulence, whereas the other two did not kill the infected mice. The remaining two transgenic lines showed low parasitemias similar to their parental 17X strain, and their EBL localizations did not change. The results indicate the importance of Cys residues in EBL region 6 for EBL localization, parasite infection course, and virulence and suggest an association between EBL localization and the parasite infection course.
... Indeed, a genome-wide search of the P. yoelii genome for simple sequence repeats (SSRs) led to the development of nearly 600 MS makers and a physical map [21]. Further, the MS markers were used to genotype progenies from several genetic crosses of P. yoelii parasites resulting in the development of a genetic linkage map [22]. ...
... Genetic mapping was initially performed using different strains of P. chabaudi to investigate drug resistance and strain-specific immunity [2,12,16,17,35,36]. More recently, genetic crosses of P. yoelii strains have been performed to investigate genes related to parasite growth, virulence, and host-parasite interaction [22,[37][38][39][40]. P. berghei was the first rodent malaria parasite used to develop methods for parasite transfection and genetic modification successfully [41,42], although genetic crosses of P. berghei wild type (WT) and gene deletion mutants were also conducted to dissect gene function in male and female gametocyte development and fertilization [43]. ...
... In another study, seventy-five IRPs were cloned from three P. yoelii genetic crosses [17XNL × N67 (25 progenies); YM × 33X (18 progenies); BY265 × NSM (32 progenies)] [22]. After typing with 486 MS markers and measurements of growth-related virulent phenotypes for the 25 progenies from the 17XNL × N67, a major locus on chromosome 13 and two secondary loci on chromosomes 7 and 10, respectively, were discovered. ...
Article
Full-text available
Genetic mapping has been widely employed to search for genes linked to phenotypes/traits of interest. Because of the ease of maintaining rodent malaria parasites in laboratory mice, many genetic crosses of rodent malaria parasites have been performed to map the parasite genes contributing to malaria parasite development, drug resistance, host immune response, and disease pathogenesis. Drs. Richard Carter, David Walliker, and colleagues at the University of Edinburgh, UK, were the pioneers in developing the systems for genetic mapping of malaria parasite traits, including characterization of genetic markers to follow the inheritance and recombination of parasite chromosomes and performing the first genetic cross using rodent malaria parasites. Additionally, many genetic crosses of inbred mice have been performed to link mouse chromosomal loci to the susceptibility to malaria parasite infections. In this chapter, we review and discuss past and recent advances in genetic marker development, performing genetic crosses, and genetic mapping of both parasite and host genes. Genetic mappings using models of rodent malaria parasites and inbred mice have contributed greatly to our understanding of malaria, including parasite development within their hosts, mechanism of drug resistance, and host-parasite interaction.
... IFN-I has been reported to be critical for stimulating stronger immune responses against N67 infection (3,5,21,22), thus, we detected whether SHIP1 deficiency affects the production of IFN-I and found that N67C-infected Ship1-chimeric mice produced significantly higher levels of IFN-α and IFN-β in the serum than the infected WT mice ( Fig. 1F and G). Meanwhile, the mRNA extracted from BMs of mice on days 1 and 6 post-N67 infection showed higher Ifn-a/b expression in Ship1-chimeric mice than the WT mice ( Fig. S1A and B). ...
... The timing of production of IFNs is critical for antimalarial immune responses (3), and that is why we observed more resistance in Ship1-chimeric mice in comparison to WT mice during malaria infection. During the N67 infection, the production of IFN-I peaks at day 1 after infection, which is consistent with previous publication (3,22), and decreases to a basal line at the late stage of infection. Although the amounts of IFN-α/β are low on day 6, Ship1-chimeric mice still generate high levels of IFN-α/β than the WT mice, which further suggests the inhibition of SHIP1 on the production of IFN-I throughout the course of malaria infection. ...
Article
Full-text available
Stringent control of the type I interferon (IFN-I) signaling is critical for host immune defense against infectious diseases, yet the molecular mechanisms that regulate this pathway remain elusive. Here, we show that Src homology 2 containing inositol phosphatase 1 (SHIP1) suppresses IFN-I signaling by promoting IRF3 degradation during malaria infection. Genetic ablation of Ship1 in mice leads to high levels of IFN-I and confers resistance to Plasmodium yoelii nigeriensis ( P.y. ) N67 infection. Mechanistically, SHIP1 promotes the selective autophagic degradation of IRF3 by enhancing K63-linked ubiquitination of IRF3 at lysine 313, which serves as a recognition signal for NDP52-mediated selective autophagic degradation. In addition, SHIP1 is downregulated by IFN-I-induced miR-155-5p upon P.y . N67 infection and severs as a feedback loop of the signaling crosstalk. This study reveals a regulatory mechanism between IFN-I signaling and autophagy, and verifies SHIP1 can be a potential target for therapeutic intervention against malaria and other infectious diseases. IMPORTANCE Malaria remains a serious disease affecting millions of people worldwide. Malaria parasite infection triggers tightly controlled type I interferon (IFN-I) signaling that plays a critical role in host innate immunity; however, the molecular mechanisms underlying the immune responses are still elusive. Here, we discover a host gene [Src homology 2-containing inositol phosphatase 1 (SHIP1)] that can regulate IFN-I signaling by modulating NDP52-mediated selective autophagic degradation of IRF3 and significantly affect parasitemia and resistance of Plasmodium -infected mice. This study identifies SHIP1 as a potential target for immunotherapies in malaria and highlights the crosstalk between IFN-I signaling and autophagy in preventing related infectious diseases. SHIP1 functions as a negative regulator during malaria infection by targeting IRF3 for autophagic degradation.
... The P. yoelii 17XL, BY265 and NSM parasites used in this study were described previously [35]. Female outbred ICR mice and inbred BALB/c mice, 6-8 weeks old, used to maintain parasites and evaluate parasite growth, respectively, were purchased from Xiamen University Laboratory Animal Center or Shanghai Laboratory Animal Center, CAS (SLACCAS). ...
Article
Full-text available
Background Eukaryotic genes contain introns that are removed by the spliceosomal machinery during mRNA maturation. Introns impose a huge energetic burden on a cell; therefore, they must play an essential role in maintaining genome stability and/or regulating gene expression. Many genes (> 50%) in Plasmodium parasites contain predicted introns, including introns in 5′ and 3′ untranslated regions (UTR). However, the roles of UTR introns in the gene expression of malaria parasites remain unknown. Methods In this study, an episomal dual-luciferase assay was developed to evaluate gene expression driven by promoters with or without a 5′UTR intron from four Plasmodium yoelii genes. To investigate the effect of the 5′UTR intron on endogenous gene expression, the pytctp gene was tagged with 3xHA at the N-terminal of the coding region, and parasites with or without the 5′UTR intron were generated using the CRISPR/Cas9 system. Results We showed that promoters with 5′UTR introns had higher activities in driving gene expression than those without 5′UTR introns. The results were confirmed in recombinant parasites expressing an HA-tagged gene ( pytctp ) driven by promoter with or without 5′UTR intron. The enhancement of gene expression was intron size dependent, but not the DNA sequence, e.g. the longer the intron, the higher levels of expression. Similar results were observed when a promoter from one strain of P. yoelii was introduced into different parasite strains. Finally, the 5′UTR introns were alternatively spliced in different parasite development stages, suggesting an active mechanism employed by the parasites to regulate gene expression in various developmental stages. Conclusions Plasmodium 5′UTR introns enhance gene expression in a size-dependent manner; the presence of alternatively spliced mRNAs in different parasite developmental stages suggests that alternative slicing of 5′UTR introns is one of the key mechanisms in regulating parasite gene expression and differentiation. Graphical Abstract
... Malaria Parasites and Mice. The 17XNL parasite and the procedures for infecting C57BL/6n mice were as described previously (67). All animal procedures were performed following the protocol approved (#LMVR11E) by the Institutional Animal Care and Use Committee at the National Institute of Allergy and Infectious Diseases (NIAID). ...
Article
Full-text available
Plasmodium parasites cause malaria with disease outcomes ranging from mild illness to deadly complications such as severe malarial anemia (SMA), pulmonary edema, acute renal failure, and cerebral malaria. In young children, SMA often requires blood transfusion and is a major cause of hospitalization. Malaria parasite infection leads to the destruction of infected and noninfected erythrocytes as well as dyserythropoiesis; however, the mechanism of dyserythropoiesis accompanied by splenomegaly is not completely understood. Using Plasmodium yoelii yoelii 17XNL as a model, we show that both a defect in erythroblastic island (EBI) macrophages in supporting red blood cell (RBC) maturation and the destruction of reticulocytes/RBCs by the parasites contribute to SMA and splenomegaly. After malaria parasite infection, the destruction of both infected and noninfected RBCs stimulates extramedullary erythropoiesis in mice. The continuous decline of RBCs stimulates active erythropoiesis and drives the expansion of EBIs in the spleen, contributing to splenomegaly. Phagocytosis of malaria parasites by macrophages in the bone marrow and spleen may alter their functional properties and abilities to support erythropoiesis, including reduced expression of the adherence molecule CD169 and inability to support erythroblast differentiation, particularly RBC maturation in vitro and in vivo. Therefore, macrophage dysfunction is a key mechanism contributing to SMA. Mitigating and/or alleviating the inhibition of RBC maturation may provide a treatment strategy for SMA.
... For instance, the SAR supergroup (a clade containing Stramenopiles, Alveolata, and Rhizaria) is a deep branching eukaryote clade encompassing brown macro-algae, as well as numerous micro-eukaryotes such as ascomycetes, ciliates, radiolarians, and dinoflagellates. Within this clade, about ten genetic maps have been developed, mostly in parasites [14][15][16][17][18][19][20][21][22][23]. There is no genetic map available for dinoflagellates, which may be explained by the difficulties of obtaining hundreds of descendants [24] to analyse quantitative traits. ...
Article
Full-text available
Dinoflagellates of the genus Alexandrium are responsible for harmful algal blooms and produce paralytic shellfish toxins (PSTs). Their very large and complex genomes make it challenging to identify the genes responsible for toxin synthesis. A family-based genomic association study was developed to determine the inheritance of toxin production in Alexandrium minutum and identify genomic regions linked to this production. We show that the ability to produce toxins is inheritable in a Mendelian way, while the heritability of the toxin profile is more complex. We developed the first dinoflagellate genetic linkage map. Using this map, several major results were obtained: 1. A genomic region related to the ability to produce toxins was identified. 2. This region does not contain any polymorphic sxt genes, known to be involved in toxin production in cyanobacteria. 3. The sxt genes, known to be present in a single cluster in cyanobacteria, are scattered on different linkage groups in A. minutum . 4. The expression of two sxt genes not assigned to any linkage group, sxtI and sxtG , may be regulated by the genomic region related to the ability to produce toxins. Our results provide new insights into the organization of toxicity-related genes in A. minutum , suggesting a dissociated genetic mechanism for the production of the different analogues and the ability to produce toxins. However, most of the newly identified genes remain unannotated. This study therefore proposes new candidate genes to be further explored to understand how dinoflagellates synthesize their toxins.
Article
Olfactory receptors (Olfr) are G protein–coupled receptors that are normally expressed on olfactory sensory neurons to detect volatile chemicals or odorants. Interestingly, many Olfrs are also expressed in diverse tissues and function in cell–cell recognition, migration, and proliferation as well as immune responses and disease processes. Here, we showed that many Olfr genes were expressed in the mouse spleen, linked to Plasmodium yoelii genetic loci significantly, and/or had genome-wide patterns of LOD scores (GPLSs) similar to those of host Toll-like receptor genes. Expression of specific Olfr genes such as Olfr1386 in HEK293T cells significantly increased luciferase signals driven by IFN-β and NF-κB promoters, with elevated levels of phosphorylated TBK1, IRF3, P38, and JNK. Mice without Olfr1386 were generated using the CRISPR/Cas9 method, and the Olfr1386 −/− mice showed significantly lower IFN-α/β levels and longer survival than wild-type (WT) littermates after infection with P. yoelii YM parasites. Inhibition of G protein signaling and P38 activity could affect cyclic AMP-responsive element promoter-driven luciferase signals and IFN-β mRNA levels in HEK293T cells expressing the Olfr1386 gene, respectively. Screening of malaria parasite metabolites identified nicotinamide adenine dinucleotide (NAD) as a potential ligand for Olfr1386, and NAD could stimulate IFN-β responses and phosphorylation of TBK1 and STAT1/2 in RAW264.7 cells. Additionally, parasite RNA (pRNA) could significantly increase Olfr1386 mRNA levels. This study links multiple Olfrs to host immune response pathways, identifies a candidate ligand for Olfr1386, and demonstrates the important roles of Olfr1386 in regulating type I interferon (IFN-I) responses during malaria parasite infections.
Article
Rodent malaria parasites have been widely used in all aspects of malaria research to study parasite development within rodent and insect hosts, drug resistance, disease pathogenesis, host immune response, and vaccine efficacy. Rodent malaria parasites were isolated from African thicket rats and initially characterized by scientists at the University of Edinburgh, UK, particularly by Drs. Richard Carter, David Walliker, and colleagues. Through their efforts and elegant work, many rodent malaria parasite species, subspecies, and strains are now available. Because of the ease of maintaining these parasites in laboratory mice, genetic crosses can be performed to map the parasite and host genes contributing to parasite growth and disease severity. Recombinant DNA technologies are now available to manipulate the parasite genomes and to study gene functions efficiently. In this chapter, we provide a brief history of the isolation and species identification of rodent malaria parasites. We also discuss some recent studies to further characterize the different developing stages of the parasites including parasite genomes and chromosomes. Although there are differences between rodent and human malaria parasite infections, the knowledge gained from studies of rodent malaria parasites has contributed greatly to our understanding of and the fight against human malaria.
Article
Full-text available
Genetic investigations of malaria require a genome-wide, high-resolution linkage map of Plasmodium falciparum. A genetic cross was used to construct such a map from 901 markers that fall into 14 inferred linkage groups corresponding to the 14 nuclear chromosomes. Meiotic crossover activity in the genome proved high (17 kilobases per centimorgan) and notably uniform over chromosome length. Gene conversion events and spontaneous microsatellite length changes were evident in the inheritance data. The markers, map, and recombination parameters are facilitating genome sequence assembly, localization of determinants for such traits as virulence and drug resistance, and genetic studies of parasite field populations.
Article
Full-text available
Species of malaria parasite that infect rodents have long been used as models for malaria disease research. Here we report the whole-genome shotgun sequence of one species, Plasmodium yoelii yoelii, and comparative studies with the genome of the human malaria parasite Plasmodium falciparum clone 3D7. A synteny map of 2,212 P. y. yoelii contiguous DNA sequences (contigs) aligned to 14 P. falciparum chromosomes reveals marked conservation of gene synteny within the body of each chromosome. Of about 5,300 P. falciparum genes, more than 3,300 P. y. yoelii orthologues of predominantly metabolic function were identified. Over 800 copies of a variant antigen gene located in subtelomeric regions were found. This is the first genome sequence of a model eukaryotic parasite, and it provides insight into the use of such systems in the modelling of Plasmodium biology and disease.
Article
Full-text available
Variation in response to antimalarial drugs and in pathogenicity of malaria parasites is of biologic and medical importance. Linkage mapping has led to successful identification of genes or loci underlying various traits in malaria parasites of rodents and humans. The malaria parasite Plasmodium yoelii is one of many malaria species isolated from wild African rodents and has been adapted to grow in laboratories. This species reproduces many of the biologic characteristics of the human malaria parasites; genetic markers such as microsatellite and amplified fragment length polymorphism (AFLP) markers have also been developed for the parasite. Thus, genetic studies in rodent malaria parasites can be performed to complement research on Plasmodium falciparum. Here, we demonstrate the techniques for producing a genetic cross in P. yoelii that were first pioneered by Drs. David Walliker, Richard Carter, and colleagues at the University of Edinburgh. Genetic crosses in P. yoelii and other rodent malaria parasites are conducted by infecting mice Mus musculus with an inoculum containing gametocytes of two genetically distinct clones that differ in phenotypes of interest and by allowing mosquitoes to feed on the infected mice 4 days after infection. The presence of male and female gametocytes in the mouse blood is microscopically confirmed before feeding. Within 48 hrs after feeding, in the midgut of the mosquito, the haploid gametocytes differentiate into male and female gametes, fertilize, and form a diploid zygote (Fig. 1). During development of a zygote into an ookinete, meiosis appears to occur. If the zygote is derived through cross-fertilization between gametes of the two genetically distinct parasites, genetic exchanges (chromosomal reassortment and cross-overs between the non-sister chromatids of a pair of homologous chromosomes; Fig. 2) may occur, resulting in recombination of genetic material at homologous loci. Each zygote undergoes two successive nuclear divisions, leading to four haploid nuclei. An ookinete further develops into an oocyst. Once the oocyst matures, thousands of sporozoites (the progeny of the cross) are formed and released into mosquito hemoceal. Sporozoites are harvested from the salivary glands and injected into a new murine host, where pre-erythrocytic and erythrocytic stage development takes place. Erythrocytic forms are cloned and classified with regard to the characters distinguishing the parental lines prior to genetic linkage mapping. Control infections of individual parental clones are performed in the same way as the production of a genetic cross.
Article
Full-text available
Elevated parasite biomass in the human red blood cells can lead to increased malaria morbidity. The genes and mechanisms regulating growth and development of Plasmodium falciparum through its erythrocytic cycle are not well understood. We previously showed that strains HB3 and Dd2 diverge in their proliferation rates, and here use quantitative trait loci mapping in 34 progeny from a cross between these parent clones along with integrative bioinformatics to identify genetic loci and candidate genes that control divergences in cell cycle duration. Genetic mapping of cell cycle duration revealed a four-locus genetic model, including a major genetic effect on chromosome 12, which accounts for 75% of the inherited phenotype variation. These QTL span 165 genes, the majority of which have no predicted function based on homology. We present a method to systematically prioritize candidate genes using the extensive sequence and transcriptional information available for the parent lines. Putative functions were assigned to the prioritized genes based on protein interaction networks and expression eQTL from our earlier study. DNA metabolism or antigenic variation functional categories were enriched among our prioritized candidate genes. Genes were then analyzed to determine if they interact with cyclins or other proteins known to be involved in the regulation of cell cycle. We show that the divergent proliferation rate between a drug resistant and drug sensitive parent clone is under genetic regulation and is segregating as a complex trait in 34 progeny. We map a major locus along with additional secondary effects, and use the wealth of genome data to identify key candidate genes. Of particular interest are a nucleosome assembly protein (PFL0185c), a Zinc finger transcription factor (PFL0465c) both on chromosome 12 and a ribosomal protein L7Ae-related on chromosome 4 (PFD0960c).
Article
Full-text available
Studies of gene function and molecular mechanisms in Plasmodium falciparum are hampered by difficulties in characterizing and measuring phenotypic differences between individual parasites. We screened seven parasite lines for differences in responses to 1,279 bioactive chemicals. Hundreds of compounds were active in inhibiting parasite growth; 607 differential chemical phenotypes, defined as pairwise IC(50) differences of fivefold or more between parasite lines, were cataloged. We mapped major determinants for three differential chemical phenotypes between the parents of a genetic cross, and we identified target genes by fine mapping and testing the responses of parasites in which candidate genes were genetically replaced with mutant alleles. Differential sensitivity to dihydroergotamine methanesulfonate (1), a serotonin receptor antagonist, was mapped to a gene encoding the homolog of human P-glycoprotein (PfPgh-1). This study identifies new leads for antimalarial drugs and demonstrates the utility of a high-throughput chemical genomic strategy for studying malaria traits.
Article
The number of chromosomes and the chromosomal location and linkage of more than 50 probes, mainly of genes, have been established in four species of Plasmodium which infect African murine rodents. We expected that the location and linkage of genes would not be conserved between these species of malaria parasites since extensive inter- and intraspecific size differences of the chromosomes existed and large scale internal rearrangements and chromosome translocations in parasites from laboratory lines had been reported. Our study showed that all four species contained 14 chromosomes, ranging in size between 0.5 and 3.5 Mb, which showed extensive size polymorphisms. The location and linkage of the genes on the polymorphic chromosomes, however, was conserved and nearly identical between these species. These results indicate that size polymorphisms of the chromosomes are more likely due to variation in non-coding (subtelomeric, repeat) sequences and show that a high plasticity of internal regions of chromosomes that may exist does not frequently affect chromosomal location and linkage of genes.
Article
The intracellular trafficking of an Erythrocyte Binding Like (EBL) ligand has recently been shown to dramatically affect the multiplication rate and virulence of the rodent malaria parasite Plasmodium yoelii yoelii. In this review, we describe the current understanding of the role of EBL and other erythrocyte binding ligands in erythrocyte invasion, and discuss the mechanisms by which they may control multiplication rates and virulence in malaria parasites.
Article
Genetic crosses have been employed to study various traits of rodent malaria parasites and to locate loci that contribute to drug resistance, immune protection, and disease virulence. Compared with human malaria parasites, genetic crossing of rodent malaria parasites is more easily performed; however, genotyping methods using microsatellites (MSs) or large-scale single nucleotide polymorphisms (SNPs) that have been widely used in typing Plasmodium falciparum are not available for rodent malaria species. Here we report a genome-wide search of the Plasmodium yoelii yoelii (P. yoelii) genome for simple sequence repeats (SSRs) and the identification of nearly 600 polymorphic MS markers for typing the genomes of P. yoelii and Plasmodium berghei. The MS markers are randomly distributed across the 14 physical chromosomes assembled from genome sequences of three rodent malaria species, although some variations in the numbers of MS expected according to chromosome size exist. The majority of the MS markers are AT-rich repeats, similar to those found in the P. falciparum genome. The MS markers provide an important resource for genotyping, lay a foundation for developing linkage maps, and will greatly facilitate genetic studies of P. yoelii.