ArticlePDF Available

Mutations in the Plasmodium falciparum chloroquine resistance transporter, PfCRT, enlarge the parasite’s food vacuole and alter drug sensitivities

Authors:

Abstract and Figures

Mutations in the Plasmodium falciparum chloroquine resistance transporter, PfCRT, are the major determinant of chloroquine resistance in this lethal human malaria parasite. Here, we describe P. falciparum lines subjected to selection by amantadine or blasticidin that carry PfCRT mutations (C101F or L272F), causing the development of enlarged food vacuoles. These parasites also have increased sensitivity to chloroquine and some other quinoline antimalarials, but exhibit no or minimal change in sensitivity to artemisinins, when compared with parental strains. A transgenic parasite line expressing the L272F variant of PfCRT confirmed this increased chloroquine sensitivity and enlarged food vacuole phenotype. Furthermore, the introduction of the C101F or L272F mutation into a chloroquine-resistant variant of PfCRT reduced the ability of this protein to transport chloroquine by approximately 93 and 82%, respectively, when expressed in Xenopus oocytes. These data provide, at least in part, a mechanistic explanation for the increased sensitivity of the mutant parasite lines to chloroquine. Taken together, these findings provide new insights into PfCRT function and PfCRT-mediated drug resistance, as well as the food vacuole, which is an important target of many antimalarial drugs.
Introduction of PfCRT L272F into Dd2. ( a ) Schematic of zinc-finger nuclease (ZFN)-mediated generation of vacuole-enlarged parasites. Dd2 parasites were first enriched for the episomal pcrt Dd2 L272F - h dhfr or pcrt Dd2 -h dhfr donor plasmids (latter not shown). The donor plasmids encoded a cDNA copy of the Dd2 pfcrt allele (dark blue, plasmid), either L272F-mutated (dark blue bump) or wild-type (not shown), followed by a dhfr selection cassette (light grey). Each donor-enriched parasite was then transfected with the pZFN pfcrt -bsd plasmid, expressing the genomic (light blue) pfcrt intron 1-targeting ZFN pair (ZFN L and ZFN R, orange) and the bsd selection cassette (dark grey). ZFN-induced recombination in pfcrt yielded either control Dd2 Dd2 (not shown) or Dd2 Dd2 L272F parasites (dark blue, locus). ( b ) PCR verification of parental, recombinant control, and Dd2 Dd2 L272F parasite clones. Primer (p) positions are shown in panel a . ( c ) Light mi c roscope analysis of representative examples of parental, recombinant control, and experimental parasite clones. Ring morphology for each parasite Dd2, Dd2 Dd2 , Dd2 Dd2 L272F , and GC03 was normal. Progression through the trophozoite and schizont stages showed normal morphological development except for the Dd2 Dd2 L272F clone, which exhibited the characteristic enlarged vacuole and diffuse hemozoin phenotypes seen in 3D7 L272F and FCB C101F Giemsa stained parasites. ( d ) Transmission electron micrographs of Dd2 Dd2 (i) and Dd2 Dd2 L272F parasites (ii) showing similar cytoplasmic appearances except for the enlarged food vacuole (FV) in Dd2 Dd2 L272F . Note neither parasite exhibits knobs. This confirms similar morphological appearances to those of 3D7 and 3D7 L272F , respectively but without knob formation. N—Nucleus. Bars represent 1 μ m.
… 
Content may be subject to copyright.
1
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
www.nature.com/scientificreports
Mutations in the Plasmodium
falciparum chloroquine resistance
transporter, PfCRT, enlarge the
parasite’s food vacuole and alter
drug sensitivities
Serena Pulcini1,*, Henry M. Staines1,*, Andrew H. Lee2, Sarah H. Shak3,
Guillaume Bouyer1,4,5, Catherine M. Moore1, Daniel A. Daley6, Matthew J. Hoke6,
Lindsey M. Altenhofen7, Heather J. Painter7, Jianbing Mu8, David J. P. Ferguson9,
Manuel Llinás7, Rowena E. Martin3, David A. Fidock2,10, Roland A. Cooper6,11 & Sanjeev Krishna1
Mutations in the Plasmodium falciparum chloroquine resistance transporter, PfCRT, are the major
determinant of chloroquine resistance in this lethal human malaria parasite. Here, we describe
P. falciparum lines subjected to selection by amantadine or blasticidin that carry PfCRT mutations
(C101F or L272F), causing the development of enlarged food vacuoles. These parasites also have
increased sensitivity to chloroquine and some other quinoline antimalarials, but exhibit no or
minimal change in sensitivity to artemisinins, when compared with parental strains. A transgenic
parasite line expressing the L272F variant of PfCRT conrmed this increased chloroquine sensitivity
and enlarged food vacuole phenotype. Furthermore, the introduction of the C101F or L272F mutation
into a chloroquine-resistant variant of PfCRT reduced the ability of this protein to transport
chloroquine by approximately 93 and 82%, respectively, when expressed in Xenopus oocytes. These
data provide, at least in part, a mechanistic explanation for the increased sensitivity of the mutant
parasite lines to chloroquine. Taken together, these ndings provide new insights into PfCRT function
and PfCRT-mediated drug resistance, as well as the food vacuole, which is an important target of
many antimalarial drugs.
1Institute for Infection and Immunity, St. George’s, University of London, London SW17 0RE, UK. 2Department of
Microbiology and Immunology, Columbia University Medical Center, New York, NY 10032, USA. 3Research School
of Biology, Australian National University, Canberra, ACT 2601, Australia. 4Sorbonne Universités, UPMC Univ. Paris
06, UMR 8227, Integrative Biology of Marine Models, Comparative Physiology of Erythrocytes, Station Biologique
de Rosco, Rosco, France. 5CNRS, UMR 8227, Integrative Biology of Marine Models, Comparative Physiology of
Erythrocytes, Station Biologique de Rosco, Rosco, France. 6Department of Biological Sciences, Old Dominion
University, Norfolk, VA 23529, USA. 7Department of Biochemistry and Molecular Biology and Center for Malaria
Research, Pennsylvania State University, State College, Pennsylvania 16802, USA. 8Laboratory of Malaria and
Vector Research, National Institute of Allergy and Infectious Diseases, National Institutes of Health, Rockville MD
20852, USA. 9Nueld Department of Clinical Laboratory Sciences, University of Oxford, John Radclie Hospital,
Oxford OX3 9DU, UK. 10Division of Infectious Diseases, Department of Medicine, Columbia University Medical
Center, New York, NY 10032, USA. 11Department of Natural Sciences and Mathematics, Dominican University
of California, San Rafael, CA 94901, USA. *These authors contributed equally to this work. Correspondence and
requests for materials should be addressed to S.K. (email: s.krishna@sgul.ac.uk)
Received: 29 May 2015
Accepted: 14 August 2015
Published: 30 September 2015
OPEN
www.nature.com/scientificreports/
2
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
Chloroquine (CQ) rapidly became one of the most useful antimalarial drugs for rst-line therapy soon
aer the Second World War. Resistance to CQ was rst reported in the late 1950s in Plasmodium fal-
ciparum. It then spread globally and forced the development of alternative regimes, culminating in the
more expensive artemisinin-based combination therapies (ACTs) used today. e locus containing the
P. falciparum chloroquine resistance transporter gene (pfcrt) was initially mapped by classical genetic
studies as being crucial to the development of CQ resistance, with this gene subsequently being identied
and its role conrmed using reverse genetic approaches1–3. CQ resistance is now emerging in P. vivax,
for which it remains the rst-line treatment4.
CQ is a diprotic weak base that accumulates in the parasite’s acidic food vacuole (FV) by diusion
and subsequent trapping by protonation. CQ interferes with the detoxication of heme in the FV, which
leads to parasite death5. Predicted to have 10 transmembrane domains (TMDs), PfCRT is located in the
FV membrane1,6 and, when mutated, increases export of CQ from the FV and its target of heme polym-
erisation7. Single nucleotide polymorphisms (SNPs) in PfCRT in eld isolates correlate with a resistance
phenotype in in vitro assays and are sensitive markers for treatment failure in patients8,9. However, these
molecular markers are not always specic because other variables such as previous exposure to malaria
can inuence treatment response in patients10.
One polymorphism at position 76 (K76T) in the rst TMD of PfCRT seems to be key to CQ resist-
ance. is substitution removes a positive charge from a predicted substrate-binding site in PfCRT, allow-
ing protonated CQ to escape from the FV down its electrochemical gradient11. Other mutations (K76I
and K76N) in this position also arise when P. falciparum is exposed in vitro to lethal concentrations of
CQ, allowing parasites to survive and supporting the critical role of this residue1,6.
e native function of PfCRT is not clear, although it has been postulated to be involved in hemoglo-
bin catabolism, possibly by mediating the transport of hemoglobin-derived peptides/amino acids from
the FV12, a hypothesis consistent with recent heterologous expression and metabolomics studies7,13,14.
PfCRT has also been proposed to function as a chloride channel, a proton pump or a regulator of
proton pumps, a general activator or modulator of transport systems (reviewed in11) or, most recently,
a proton-coupled transporter of a broad range of cationic substrates15. ere are many reasons to eluci-
date the function of PfCRT in parasites, including the suggestion that PfCRT could itself become a new
drug target16,17, or that chemosensitizing agents could be directed against PfCRT to restore the ecacy
of CQ16–18. Furthermore, CQ continues to be used in the treatment of non-falciparum malarias. It may
also regain ecacy against falciparum malaria in areas where usage has been tightly regulated, since the
withdrawal of CQ can result in dramatic decreases in the prevalence of CQ-resistant parasites19.
Here, mutations in pfcrt that alter parasite phenotype give new insights into its native function as a
transporter. e novel and pleiotropic phenotypic characteristics associated with mutated PfCRT include
altered FV morphology and changes in quinoline sensitivities. We also investigated the eect of these
changes on the parasites sensitivity to other antimalarial classes, such as the artemisinins, that some have
considered to act (at least in part) in the FV of the parasite20,21.
Results
New and previously described mutations in pfcrt. SNPs were identied in pfcrt in two dierent P.
falciparum lines (Fig.1a,b). e rst was discovered aer isolating amantadine (AMT)-resistant mutants
of the CQ-resistant parasite strain FCB, following selection with 80 μ M of this antiviral agent. Viable par-
asites were observed in one of four drug-pressured asks at 42 days, whereas none had emerged within
the remaining asks by 60 days. PCR amplication and sequencing of pfcrt in four clonal lines derived
from the AMT-resistant culture detected a single non-synonymous SNP, g302t. is encoded the amino
acid mutation C101F. ese lines were therefore designated FCBC101F. Position 101 is predicted to lie
within the second TMD of PfCRT (Fig.1a). is mutation was earlier observed in a CQ-resistant Dd2
parasite line derived by continuous piperaquine (PPQ) pressure22, although that study did not describe
any changes in parasite morphology.
e second parasite line, derived from the CQ-sensitive strain 3D7, was selected by blasticidin (BSD)
pressure as an inadvertent outcome of transfection experiments on an unrelated gene (that had aimed
to achieve single cross-over homologous recombination with a tagging plasmid under BSD selection)23.
Aer several weeks of selection, pfcrt cDNA transcripts of the daughter parasite line and parental 3D7
were sequenced. A mutation at position c814t in the pfcrt coding sequence, resulting in the amino acid
mutation L272F, was detected in the selected line, designated 3D7L272F, and was absent in its parent. is
substitution is positioned immediately aer the seventh predicted TMD, placing it in the FV compart-
ment (Fig.1a). To our knowledge, this mutation has not been reported previously. No other mutations
in pfcrt were detected in either of the new parasite lines.
Given that 3D7L272F arose in unusual circumstances (BSD is a general inhibitor of protein translation
and is not thought to target the FV), whole-genome sequencing was undertaken to identify further
mutations. is conrmed the presence of the c814t mutation in pfcrt and identied only 2 additional
SNPs. e rst was c5549g in PF3D7_1229100 (the P. falciparum multidrug resistance-associated pro-
tein 2, PfMRP2), resulting in a stop-gain mutation (S1850*) and the loss of 259 amino acids from the
C-terminus. e second was t1032a in PF3D7_1462400 (a conserved protein of unknown function),
resulting in a stop-gain mutation (Y344*) and the loss of 2979 amino acids from the C-terminus.
Truncation of the latter sequence has been observed in other laboratory clones of 3D724. Furthermore,
www.nature.com/scientificreports/
3
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
there was no evidence of integrated copies of the plasmid vector containing the BSD selection marker23,
which had been used during the generation of the 3D7L272F line.
Enlarged FVs of parasites with mutations in pfcrt. A monstrously swollen FV was observed at all
stages that ordinarily display a vacuole in the asexual cycle of both parasite lines FCBC101F and 3D7L272F
(Fig.2). is phenotype was stably maintained in the parasites following repeated rounds of parasite cul-
ture and cryopreservation. e enlarged FVs were already apparent in the early to mid trophozoite stages
of the FCBC101F line, when compared with FVs from FCB parental controls (Fig.2a le and right panels).
In more mature FCBC101F parasites, the FVs were strikingly clear in appearance, with hemozoin crystals
apparently marginalized to the FV periphery and opposite the developing nuclei, although live imaging
suggests that the hemozoin is distributed normally (Fig. 2b). e immature ring stages of development
were indistinguishable from those of the parental strain. Similar ndings were evident in the parasite
line 3D7L272F when compared with 3D7 (Fig.2c le and right panels). Measurement of the area of the
FV was also undertaken and expressed as a ratio of the parasites area to correct for parasite age (Fig.3a).
is conrmed that FCBC101F and 3D7L272F parasites have a relative FV/parasite area that is approximately
twice that of FCB and 3D7, respectively (p < 0.0001). Neither FCBC101F nor 3D7L272F parasites appeared
to be enlarged within their host red blood cells (RBCs).
e 3D7L272F line was selected for a more detailed characterization. e appearances of parasites
examined with transmission electron microscopy (TEM) were consistent with observations made with
light microscopy (Fig.3b,c), with few dierences evident between parental strains and daughter parasite
lines except for the size of the FV. Specic to this line, TEM also revealed that “knobs, electron dense
protrusions of the RBC membrane caused by parasite infection, which are important determinants of
cytoadherence25 and which are oen lost from infected RBCs during long term parasite culture26, were
Figure 1. PfCRT mutations. (a) Schematic representation of PfCRT and positions of previously identied
polymorphisms8,71 from eld isolates (green circles) and from drug-pressured laboratory lines (purple
circles). e critical CQ resistance mutation site (K76) is shaded red, and the two residues at which
mutations are described in this study are shaded in orange (C101) and blue (L272). (b) PfCRT haplotypes
included this study.
www.nature.com/scientificreports/
4
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
displayed approximately 7.5-fold more on the host surface of 3D7L272F-infected RBCs than 3D7-infected
RBCs. is is unlikely to be directly related to the mutation in pfcrt and may be due to sub-population
selection.
Since BSD pressure has been shown to alter infected RBC permeability27–29, electrophysiological
transport studies were also undertaken to compare 3D7 and 3D7L272F-infected RBCs, although no dier-
ences were observed (Supplementary Fig. S1).
Figure 2. Representative morphology of parasite lines FCBC101F and 3D7L272F. (a) Appearance of
enlarged FVs in xed FCBC101F parasites (le panel), when compared with parental FCB parasites of similar
developmental stages (right panel). (b) Images of live FCBC101F and FCB trophozoite-stage parasites, using
bright-eld and dark-eld microscopy (le and right panels, respectively). (c) Appearance of enlarged FVs
in xed 3D7L272F parasites (le panel), when compared with parental 3D7 parasites of similar developmental
stages (right panel). e diameter of a RBC is ~7 μ m.
www.nature.com/scientificreports/
5
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
In vitro sensitivity to antimalarials. Both cell lines with mutations in pfcrt displayed altered suscep-
tibility to antimalarials when compared with the parental strains (Table 1). Using a 72 h in vitro growth
inhibition assay that yields IC50 values, FCBC101F parasites were found to be 83 fold less susceptible to
AMT (used in its selection). FCBC101F showed a 5–6 fold increase in sensitivity to CQ, yet interestingly
still retained the characteristic verapamil (VP)-reversibility of CQ-resistant parasites30. Furthermore,
compared with FCB, FCBC101F was signicantly (p < 0.01) more sensitive to quinolines (quinine (QN),
quinidine (QD) and monodesethyl amodiaquine (MDAQ)) but not the arylmethanol, meoquine (MQ).
ere was a small (29%) increase in sensitivity to artemisinin (ART; p <0.01). e FCBC101F line became
approximately 2-fold more resistant to PPQ relative to controls (p < 0.05).
In similar experiments, 3D7L272F parasites, assayed over 48 h in vitro, were ~2.5 fold more sensitive to
CQ than 3D7, with respective IC50 values of 6.1 and 15 nM (p < 0.01). e 3D7L272F parasites were also
slightly more sensitive to QN than 3D7 parasites. VP sensitivity was not examined because unlike FCB,
the 3D7 line is already CQ-sensitive. e increased sensitivity to CQ therefore indicates that 3D7L272F is
Figure 3. Morphological comparisons of parasites with mutations in pfcrt compared with their parental
controls. (a) Parasites were synchronized by sorbitol lysis and areas of FVs and parasites (approximately
38 h post-invasion) were measured and expressed as ratios (AFV/AParasite). 3D7, FCB and Dd2Dd2 (open
bars; n = 72, 31 and 42, respectively) and 3D7L272F, FCBC101F and Dd2Dd2 L272F (closed bars; n = 84, 58 and
88, respectively) parasites were analyzed and signicant enlargement of the FV was conrmed in 3D7L272F,
FCBC101F and Dd2Dd2 L272F parasites relative to 3D7, FCB and Dd2Dd2, respectively (*p < 0.0001: two-
tailed, unpaired, Student’s t-test). (b,c) Transmission electron micrographs of 3D7 and 3D7L272F parasites,
respectively, showing the food vacuole (FV) and nucleus (N). Note the enlarged electron lucent FV in
3D7L272F (suggesting changes in the process of hemoglobin degradation and formation of hemozoin crystals).
RBCs infected with 3D7L272F displayed approximately 7.5-fold more knobs (arrowheads) on the host surface
than 3D7-infected RBCs, although this is likely due to sub-population selection rather than a direct link to
the mutation in pfcrt. (insert) Detail of a knob. Bars represent 1 μ m (b,c) and 100 nm (insert).
Drug
Mean ± SEM IC50 values for individual parasite strains/lines
3D7 3D7L272F FCB FCBC101F FCB + VP FCBC101F + VP
CQ 15 ± 1.8 6.1 ± 1.5* 187 ± 7.1 34 ± 1.8* 47 ± 1.2 14 ± 1.0
QN 176 ± 17 108 ± 13* 333 ± 5.8 220 ± 23* 161 ± 33 223 ± 31
QD 167 ± 13 62 ± 3.1* 46 ± 1.9 46 ± 2.9
MQ 87 ± 29 64 ± 17 13 ± 1.1 14 ± 0.9 8.8 ± 1.0 18 ± 2.0
AQ 25 ± 2.3 27 ± 2.1 — —
MDAQ 14 ± 1.5 15 ± 1.5 52 ± 4.1 18 ± 1.0* 16 ± 1.0 9.1 ± 0.7
PPQ 18 ± 4.1 22 ± 5.3 12 ± 0.6 27 ± 1.8* 13 ± 0.4 25 ± 3.1
ART 3.5 ± 1.2 4.1 ± 1.6 13 ± 0.9 9.1 ± 0.7** 9.9 ± 0.9 9.1 ± 0.5
BSD (μ M) 1.5 ± 0.6 47 ± 6.0* — —
AMT (μ M) 5.6 ± 0.6 465 ± 54* 11 ± 0.5 687 ± 32
Table 1. In vitro sensitivity of 3D7, 3D7L272F, FCB and FCBC101F (in presence or absence of verapamil,
VP) to antimalarial drugs. CQ, chloroquine; QN, quinine; QD, quinidine; MQ, meoquine; AQ,
amodiaquine; MDAQ, monodesethyl amodiaquine; PPQ, piperaquine; ART, artemisinin; BSD, blasticidin;
AMT, amantadine; VP, verapamil (used at 0.8 μ M). IC50 values are listed in nM, except where indicated,
and are show as the mean ± SEM. n = 3 independent assays (each performed as a single replicate for FCB
parasites and in quintuplicate for 3D7 parasites). Signicantly dierent mean IC50 values relative to controls
(F-test; *p < 0.05, **p < 0.01).
www.nature.com/scientificreports/
6
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
a ‘CQ-hypersensitive’ parasite line. e mean IC50 values for MQ, MDAQ, PPQ and ART were similar
between the 3D7L272F and 3D7 parasites (Table1).
Transfection studies. To conrm the phenotype observed in 3D7L272F, we engineered the L272F
mutation in pfcrt using zinc-nger nuclease mediated allelic replacement31 in the Dd2 line and compared
results with congenic controls. Figure 4a,b illustrate this strategy and provide conrmation of integra-
tion. As observed in 3D7L272F, signicant FV distension (~2 fold as measured by vacuolar area relative
to parasite area; Fig. 3a) was generated by introduction of this single amino acid change (Fig. 4c,d).
However, a signicant increase in BSD resistance was not observed between the Dd2Dd2 L272F line and its
congenic control, Dd2Dd2 (Table2), which suggests that the PfCRT L272F mutation was not primarily
responsible for the BSD resistance found in 3D7L272F parasites. e parental strain Dd2 and the congenic
control Dd2Dd2 were both CQ-resistant. However, Dd2Dd2 L272F was considerably more susceptible to CQ
and monodesethyl chloroquine (MDCQ) than the Dd2Dd2 line, although the IC50 values of the L272F
Figure 4. Introduction of PfCRT L272F into Dd2. (a) Schematic of zinc-nger nuclease (ZFN)-mediated
generation of vacuole-enlarged parasites. Dd2 parasites were rst enriched for the episomal pcrtDd2 L272F-
hdhfr or pcrtDd2-hdhfr donor plasmids (latter not shown). e donor plasmids encoded a cDNA copy of
the Dd2 pfcrt allele (dark blue, plasmid), either L272F-mutated (dark blue bump) or wild-type (not shown),
followed by a dhfr selection cassette (light grey). Each donor-enriched parasite was then transfected with
the pZFNpfcrt-bsd plasmid, expressing the genomic (light blue) pfcrt intron 1-targeting ZFN pair (ZFN L
and ZFN R, orange) and the bsd selection cassette (dark grey). ZFN-induced recombination in pfcrt yielded
either control Dd2Dd2 (not shown) or Dd2Dd2 L272F parasites (dark blue, locus). (b) PCR verication of
parental, recombinant control, and Dd2Dd2 L272F parasite clones. Primer (p) positions are shown in panel a.
(c) Light microscope analysis of representative examples of parental, recombinant control, and experimental
parasite clones. Ring morphology for each parasite Dd2, Dd2Dd2, Dd2Dd2 L272F, and GC03 was normal.
Progression through the trophozoite and schizont stages showed normal morphological development
except for the Dd2Dd2 L272F clone, which exhibited the characteristic enlarged vacuole and diuse hemozoin
phenotypes seen in 3D7L272F and FCBC101F Giemsa stained parasites. (d) Transmission electron micrographs
of Dd2Dd2 (i) and Dd2Dd2 L272F parasites (ii) showing similar cytoplasmic appearances except for the enlarged
food vacuole (FV) in Dd2Dd2 L272F. Note neither parasite exhibits knobs. is conrms similar morphological
appearances to those of 3D7 and 3D7L272F, respectively but without knob formation. N—Nucleus. Bars
represent 1 μ m.
www.nature.com/scientificreports/
7
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
mutant remained higher than the fully CQ-sensitive reference line GC03 (Table2). ART sensitivity, as
measured in these IC50 assays, was unaltered across parasites. ere were no dierences in whole-cell
electrophysiological properties between the Dd2Dd2 and Dd2Dd2 L272F parasite lines (Supplementary Fig.
S1) and the RBCs infected with Dd2Dd2 L272F parasites remained knobless (Fig.4d), suggesting that the
increased expression of knobs in 3D7L272F-infected RBCs was not related to the L272F mutation in pfcrt.
Measurements of CQ transport via the C101F and L272F variants of PfCRT. e Xenopus
oocyte system for the heterologous expression of PfCRT7 was employed to investigate the eect of the
C101F and L272F mutations on the ability of PfCRT to mediate CQ transport. e L272F and C101F
mutations were introduced into the Dd2 haplotype of PfCRT (PfCRTDd2, from the CQ-resistant strain
Dd2; Fig.1b) and L272F was also introduced into PfCRT3D7 (from the CQ-sensitive strain 3D7; Fig.1b).
e resulting variants (L272F PfCRTDd2, L272F PfCRT3D7, and C101F PfCRTDd2), as well as PfCRTDd2 and
PfCRT3D7, were expressed in oocytes. Localization of each of the PfCRT variants to the oocyte plasma
membrane was conrmed by immunouorescence assay (Supplementary Fig. S2a) and a semiquantita-
tive western blot analysis32 indicated that the dierent PfCRT proteins were present at similar levels in
the oocyte membrane (Supplementary Fig. S2b). e ability of the PfCRT variants to mediate [3H]CQ
transport was measured in an acidic medium (pH 5.5), in which the majority of CQ is protonated. e
extent to which oocytes expressing PfCRTDd2 accumulate [3H]CQ varies considerably between batches
of oocytes from dierent frogs, with the PfCRTDd2-expressing oocytes accumulating between 8 and 45
times more [3H]CQ than the control (non-injected and PfCRT3D7-expressing) oocytes. Hence, within
each experiment uptake was expressed relative to that obtained for oocytes expressing PfCRTDd2 (in
the absence of inhibitors). Non-injected oocytes and oocytes expressing PfCRT3D7 have previously been
shown to take up CQ to similar (low) levels via simple diusion of the neutral species of the drug7,32;
this represents the ‘background’ level of CQ accumulation in oocytes, which in this study was estimated
by measuring CQ uptake into PfCRT3D7-expressing oocytes (see Supplementary Fig. S3).
In the data presented in Fig.5a,b, oocytes expressing PfCRTDd2 showed an 11 to 40-fold (mean and
SEM of 21 ± 3; n = 9 separate experiments) increase in CQ uptake relative to the PfCRT3D7-expressing
control. e component of CQ accumulation attributable to diusion (i.e. the uptake of CQ measured
in PfCRT3D7-expressing oocytes) was subtracted to obtain the PfCRT-mediated component of CQ trans-
port. Supplementary Figure S3 shows the total level of CQ accumulation in each oocyte and treatment
type. e introduction of L272F or C101F into PfCRTDd2 substantially reduced the proteins ability to
transport CQ (by ~82% and ~93%, respectively; p < 0.001, ANOVA) whereas the introduction of L272F
into PfCRT3D7 was without eect (p > 0.05). e addition of the CQ resistance-reverser VP (250 μ M)
reduced PfCRTDd2-mediated CQ transport by ~93% (p < 0.001) and also dramatically decreased CQ
uptake via L272F PfCRTDd2 and C101F PfCRTDd2 (by ~84% and ~92%, respectively; p < 0.01), such that
the accumulation of CQ in the latter two treatments was not signicantly dierent from that measured
in the PfCRT3D7-expressing controls (p > 0.05).
To investigate how BSD pressure might have produced the 3D7L272F mutant, interactions between
the PfCRT variants and BSD were assessed by measuring the uptake of [3H]CQ in the presence of
unlabeled BSD (100 or 500 μ M; Fig. 5b). e addition of BSD reduced CQ transport via PfCRTDd2 by
~39% (100 μ M; p < 0.001) and ~56% (500 μ M; p < 0.001) and, to a lesser degree, decreased CQ uptake
via L272F PfCRTDd2 (by ~22% (p > 0.05) and ~49% (p < 0.01), respectively). Neither concentration of
BSD reduced the C101F PfCRTDd2-mediated transport of CQ (p > 0.05), nor was the accumulation of
CQ in the PfCRT3D7-expressing controls aected (p > 0.05). Note that the micromolar concentrations of
the compounds used here to inhibit PfCRT are physiologically relevant given that when present in the
Drug
Mean ± SEM IC50 values for individual parasite strains/lines
Dd2 Dd2Dd2 Dd2Dd2 L272F GC03
CQ 97 ± 6.8 88 ± 6.8 20 ± 1.8* 13 ± 2.3
MDCQ 497 ± 43 440 ± 33 121 ± 9.0* 26 ± 3.0
MDAQ 45 ± 5.8 35 ± 3.6 22 ± 4.0** 18 ± 1.1
PPQ 32 ± 3.1 32 ± 5.8 39 ± 1.5 23 ± 5.3
ART 18 ± 1.1 18 ± 4.6 13 ± 1.9 15 ± 4.7
BSD 456 ± 46 631 ± 35*** 708 ± 54 456 ± 59
Table 2. In vitro sensitivity of Dd2, Dd2Dd2, Dd2Dd2 L272F and GC03 to antimalarial drugs. CQ,
chloroquine; MDCQ, monodesethyl chloroquine; MDAQ, monodesethyl amodiaquine; PPQ, piperaquine;
ART, artemisinin; BSD, blasticidin. IC50 values are listed in nM and are shown as the mean ± SEM. n = 3
independent assays (each performed in duplicate). Signicantly dierent mean IC50 values between Dd2Dd2
and Dd2Dd2 L272F (F-test; *p < 0.0001, **p = 0.07) and between Dd2 and Dd2Dd2 (F-test; ***p = 0.038).
www.nature.com/scientificreports/
8
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
extracellular solution at nanomolar levels, these protonatable drugs are expected to accumulate within
the parasite’s FV via weak-base trapping to micromolar or millimolar concentrations.
Discussion
Mannaberg stained parasites with Romanowsky’s dyes and published detailed studies on the eects of
QN against P. falciparum, which described the emergence of a ‘dropsical distension’ (enlarged FV) in
mature parasites33. Here, we describe a similar peculiar phenotype of P. falciparum parasites that is visible
without the application of antimalarial drugs. is phenotype is comparable between two parasite lines
that have mutations in pfcrt in dierent positions (amino acids 101 and 272) and that have been selected
by two chemically unrelated compounds (AMT and BSD). ese mutations conrm that pfcrt encodes a
function that is critical to maintaining FV volume. In support of this function, mutations in PfCRT that
cause CQ resistance have been reported to increase FV volume34. However, the parasite lines described
in this present study have clearly enlarged FVs but with PfCRT mutations that render the parasites more
CQ sensitive than their control strains (be that either CQ-sensitive 3D7 or CQ-resistant FCB), suggesting
an alternative mechanism of FV volume regulation is induced.
An enlarged FV is also oen observed in the presence of protease inhibitors, such as E64 or leupep-
tin35. Interference with the digestion of hemoglobin leads to a buildup of darkly staining FVs in electron
micrographs and, eventually, to parasite death. e parasites described here have enlarged FVs but these
are electron lucent (Figs2 and 4), suggesting that the digestion of hemoglobin is relatively unaected
(further supported by the presence of visible hemozin within the FVs). e simplest explanation for these
observations is that the C101F and L272F mutations interfere with the transport of the natural substrates
of PfCRT out of the FV. e resulting increase in FV osmotic pressure would lead to water ingress and
produce the unusual swelling observed in the FV of the FCBC101F, 3D7L272F, and Dd2Dd2 L272F parasites.
Figure6 presents a schematic model of this process. ese morphological changes are associated with
other phenotypic changes (which are discussed below). e natural substrate(s) of PfCRT are yet to be
identied. Studies performed with other PfCRT expression systems have reported that the protein might
function as a chloride channel, a proton pump, an activator of Na+/H+ exchangers and non-specic
cation channels or, most recently, a transporter of cationic amino acids as well as a very broad range of
other cations15. However, in many of these studies the insertion of PfCRT into the foreign membrane
required its fusion to other proteins/polypeptides, and in the most recent study the additions to PfCRT
were at both the N- and C- termini, almost doubled its size, and included a protein of undetermined
function15. Moreover, in this and the previous studies, little or no interaction could be detected between
PfCRTDd2 and known inhibitors of this protein (e.g. VP). Of signicant note, the transport kinetics for
the proposed natural substrates did not dier signicantly between PfCRTDd2 and PfCRT3D7—despite
Figure 5. CQ transport activity of the C101F and L272F variants of PfCRT in Xenopus oocytes. (a,b)
e uptake of [3H]CQ into oocytes expressing PfCRT was measured in the absence (closed bars) or presence
of 250 μ M VP (light grey bars; a), 100 μ M BSD (dark grey bars; b), or 500 μ M BSD (open bars; b). Within
each experiment, measurements were made from 10 oocytes per treatment and uptake was expressed
relative to that measured in the PfCRTDd2-expressing oocytes under control conditions. e normalized
data obtained from 4–5 separate experiments (each using oocytes from dierent frogs) were then averaged
and are shown + SEM. Both panels show PfCRT-mediated CQ uptake, calculated by subtracting CQ uptake
measured in PfCRT3D7-expressing oocytes (i.e. the component of CQ accumulation attributable to diusion;
see Supplementary Fig. S3) from that measured in oocytes expressing a variant of PfCRT. In the control
treatments, the rates of CQ uptake (pmol/oocyte/h; n = 9 ± SEM) in oocytes expressing PfCRTDd2 and
PfCRT3D7 were 23.6 ± 2.3 and 1.3 ± 0.2, respectively. ‘ns’ denotes no signicant dierence (p > 0.05) in CQ
accumulation between oocytes expressing a PfCRT variant (in the presence or absence of VP or BSD) and
that measured in the PfCRT3D7-expressing oocytes under control conditions.
www.nature.com/scientificreports/
9
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
multiple lines of evidence indicating that PfCRTDd2 imparts a substantial tness cost13,36–38. Furthermore,
the recent nding that much higher levels of acidic amino acids and/or short acidic peptides accumulate
within CQ-resistant parasites than in CQ-sensitive strains7,13,14 is not readily reconciled with PfCRT
functioning as a chloride channel, a proton pump, or a non-specic cation channel/transporter. ese,
plus other inconsistencies in the data, suggest that PfCRT does not function correctly when fused to
other proteins and that the natural function of PfCRT remains to be resolved.
AMT is an antiviral agent with moderate antimalarial activity that is more potent against CQ-resistant
parasites than against CQ-sensitive strains39. AMT is likely to accumulate in the FV via weak-base
trapping40 and is a low-anity inhibitor of the PfCRTDd2-mediated transport of CQ in the oocyte sys-
tem7. While the antiplasmodial target of AMT remains unclear, AMT resistance has been linked previ-
ously to novel PfCRT mutations (S163R, I356V and V369F; Fig.1) selected in parasites harboring CQ
resistance-associated alleles of pfcrt; these mutations were linked with the loss of CQ resistance in the
AMT-resistant mutants41,42. Here, a dierent single mutation (C101F) in the CQ-resistant FCB strain was
likewise associated with a gain of AMT resistance and a reduction in CQ resistance. is mutation was
identied previously in a PPQ-pressured parasite line that appeared to have acquired an unstable PPQ
resistance phenotype via multiple genetic changes22. One of two PPQ-revertant lines derived during that
study was ~2-fold more resistant to PPQ than the parental Dd2 strain, which along with a reduction in
CQ resistance, is consistent with the data reported here for FCBC101F.
It has been suggested that the S163R mutation reintroduces a positive charge into the PfCRT binding
pocket/translocation pore, thereby compensating for the loss of the positively-charged lysine residue
from position 7611 and resulting in a dramatic reduction in the ability of the protein to transport pro-
tonated CQ7. e S163R mutation also abolishes the CQ resistance-reversing eect of VP42. e C101F
and V369F mutations both entail the introduction of a bulky hydrophobic residue, rather than one car-
rying a positive charge, and it is interesting to note that VP still exerted a resistance-reversing eect in
the FCBC101F parasites (Table1)—even though they were considerably less resistant than the FCB strain
Figure 6. Hypothetical schematic model of the eects of PfCRT mutations. e FV is acidied by
a vacuolar proton pump to create a suitable environment for hemoglobin digestion. e acidic nature
of the FV also leads to near complete diprotonation of CQ, which diuses across the FV membrane
in an uncharged form (CQ) and accumulates as a charged form (either CQH+ or CQH22+, although
predominantly CQH22+). CQH22+ interferes with the polymerization of toxic heme to non-toxic hemozoin,
which leads to parasite death. In normal CQ-sensitive (CQS) parasites, PfCRT, which contains a positive
charge in its pore (K76), exports its natural substrates but little, if any, CQH22+. us, CQH22+ accumulates
in the FV and causes parasite death. In CQ-resistant (CQR) parasites, the positive change in the pore of
PfCRT is lost (K76T) and both its natural substrates and CQH22+ are transported out of the FV. As CQH22+
cannot accumulate in the FV, the parasites become resistant to the drug. In 3D7L272F parasites (where the
parent strain is already CQS), the mutation may reduce residual transport of CQH22+ out of the FV even
further or completely, leading to a greater FV accumulation of CQH22+ and CQ-hypersensitivity or some
other mechanism may be responsible for this phenomenon. e mutation also leads to a reduction in the
export of natural substrates, resulting in a build-up of these substrates. is causes water to enter the FV by
the process of osmosis, leading to swelling. In FCBC101F and Dd2Dd2 L272F parasites (where the parent strains
are CQR), the mutations reduce the export of CQH22+ back towards levels measured in CQS lines and
also reduce natural substrate export, leading to normal CQ sensitivity and FV swelling, respectively. Note
mutations in PfCRT (orange graphic) are denoted by red transmembrane or loop regions, depending on the
location of the amino acid change (see Fig.1a).
www.nature.com/scientificreports/
10
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
to CQ. ese observations are consistent with our direct measurements of CQ transport via C101F
PfCRTDd2 (Fig.5), which conrmed that this protein possesses a relatively low level of CQ transport
activity that can be inhibited by VP. Likewise, our nding that the introduction of L272F into PfCRTDd2
causes a dramatic (but not complete) reduction in the protein’s capacity for CQ transport (Fig.5) corre-
lates well with the low level of CQ resistance exhibited by the Dd2Dd2 L272F line. e phenylalanine resi-
dues are likely to be proximate to the binding site and/or translocation pore of PfCRT (Fig.1) where their
bulky side chains may act to signicantly hinder the transport of certain drugs out of the FV, including
CQ, QN and QD (based on the growth assay data presented in Tables1 and 2). Another mutation that
has arisen under the AMT pressure of a CQ-resistant strain, and which also resulted in both the intro-
duction of a phenylalanine residue (V369F) and a signicant reduction in CQ resistance, did not cause
the FV to swell41. Hence, the enlarged FV phenotype described here appears to manifest only when the
bulky phenylalanine side chain is inserted at specic positions within PfCRT.
BSD is used commercially as a fungicide against a rice blast disease and acts by inhibiting protein
translation. In biological research it is used to select transformed cells. BSD resistance has previously been
linked to altered expression of clag3.1 and a decrease in the RBC membrane permeability mediated by the
new permeability pathways, NPP28,29. Neither 3D7L272F nor Dd2Dd2 L272F parasites diered in their electro-
physiological NPP characteristics when assayed by whole-cell patch-clamp methods (Supplementary Fig.
S1). is suggests that one or more clag3.1-independent BSD resistance mechanisms exist. Our results
indicate that, under the conditions of the growth assay, the L272F mutation does not cause a signicant
increase in BSD resistance when introduced in isolation into Dd2 parasites (Table 3). Nevertheless, BSD
was found to inhibit CQ uptake via PfCRTDd2 and the potency of this interaction appeared to decrease
upon the introduction of L272F (the addition of 100 μ M BSD was more eective against PfCRTDd2 than
against L272F PfCRTDd2; Fig.5). A demonstration that BSD interacts with PfCRTDd2, and to a lesser
extent with L272F PfCRTDd2, provides support for the idea that BSD also interacts with, and may be
transported by, PfCRT3D7. BSD contains two protonatable nitrogens with pKa values that are well above 7.
It is therefore likely to be accumulated within the FV via weak-base trapping to high micromolar, or even
millimolar, concentrations when present in the extracellular solution at the concentration (5.4 μ M) under
which the 3D7L272F line arose. Given that BSD inhibits protein translation, which occurs outside of the
FV, it is possible that a PfCRT3D7-mediated eux of BSD from the FV could increase the drug’s access to
its main target and that the L272F mutation diminishes this activity, such that the drug remains seques-
tered within the FV. e nding that L272F PfCRTDd2 does not confer BSD resistance when expressed
in Dd2 parasites suggests that either (1) PfCRTDd2 is already a poor transporter of BSD (noting that the
3D7 and Dd2 haplotypes of PfCRT dier by eight mutations) and a reduction in this meager transport
activity by the introduction of L272F has little eect on the accumulation of BSD within the FV, or (2)
PfCRTDd2 has a very high capacity for BSD transport, such that the presence of L272F causes only a mod-
est reduction in its ability to redistribute BSD from the FV into the cytosol. In any case, it is clear that if
L272F is directly involved in altering the parasite’s susceptibility to BSD, its eect is only evident when
one or more other changes are present. In this regard, it is worth noting that one of the two mutations
identied by whole genome analysis of the 3D7L272F line would result in a truncated PfMRP2 protein.
An understanding of the contribution of this transporter to BSD susceptibility, and its possible interplay
with the BSD transport activity of PfCRT, requires further transfection-based analysis.
A diverse range of PfCRT variants implicated in conferring CQ resistance have been shown to
exhibit CQ transport activity (to varying extents) in the oocyte system7,32. However, CQ transport via
the wild-type form of the protein (found in CQ-sensitive parasites such as 3D7) has not been detected
in this assay. Although it is possible that a very low level of CQ eux is mediated by PfCRT3D7 in situ,
and that the introduction of the L272F mutation abolishes this activity, it is perhaps unlikely that this
would result in the 2-fold dierence in the CQ IC50 observed between the 3D7L272F and 3D7 parasites.
An alternative explanation for the hypersensitivity of the 3D7L272F line to CQ and QN entails view-
ing the eect of the L272F mutation as being equivalent to the eect of an ‘anti-PfCRT’ drug. e
presence of the L272F mutation causes the FV to swell, probably because it signicantly obstructs the
PfCRT-mediated eux of solutes from this compartment. A drug that binds to the substrate-binding
site of PfCRT3D7, thereby blocking or dramatically reducing its normal activity, would achieve a similar
eect. If such an anti-PfCRT drug were applied in combination with CQ or QN, which also exert their
antimalarial eects in the FV, it is possible that an additive, or even synergistic, interaction would be
observed in 3D7L272F parasites. at is, the 3D7L272F parasites have a dis-functional FV and this could
render certain drugs more eective against them; perhaps the altered composition of the FV lumen
alters the solubilities of CQ and QN and/or their anities for heme. It is not immediately apparent
why AQ, MDAQ, and PPQ are not likewise more active against 3D7L272F parasites. However, it is worth
noting that AQ, MDAQ and PPQ are much more lipophilic than CQ and this may explain dierences
in potency43,44. Hence, the antimalarial activities of the latter drugs might be less sensitive to changes in
FV volume and composition. Alternatively, or in addition, it is also possible that extending the growth
assays to 72 or 96 h (from 48 h) would reveal dierences in AQ, MDAQ, and PPQ sensitivity between
the 3D7 and 3D7L272F parasites.
If artemisinins act mainly in the FV, a disputed suggestion45 (along with reports of predominantly
non-FV-localized artemisinin targets, e.g.46,47), then their activities may also dier between the parental
and mutant lines. ere was no signicant change in the ART sensitivity of the 3D7, Dd2 and GC03 lines
www.nature.com/scientificreports/
11
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
and only a < 30% change in the IC50 value obtained for FCB (which displays an IC50 value < 15 nM).
Recent observations made in a P. berghei model using protease knockouts that alter vacuolar morphology
also leave artemisinin sensitivity unaltered48. is is also consistent with the lack of PfCRT expression
in early ring stages49, when artemisinins exert their major antimalarial action in vivo. Similarly, the MF
IC50 value was unaected in the experiments presented here. Our observations therefore relate relatively
large changes in the activities of several aminoquinolines to an enlarged FV phenotype that is caused by
specic mutations in PfCRT.
Our results show for the rst time that mutations at position 272 and 101 in PfCRT can hypersensitize
parasites to CQ and enlarge the FV, thereby extending the function of this key transporter to include
maintenance of FV morphology. We suggest that the introduction of a phenylalanine residue at either of
these positions decreases the proteins ability to transport its physiological substrate(s) (as well as certain
drugs) and that the resulting build-up of the physiological substrate(s) causes the FV to swell. e fact
that these mutations do not reintroduce a positive charge into the predicted binding cavity/translocation
pore of PfCRT, as has been observed in other examples of laboratory parasites that revert to CQ-sensitive
status (e.g.42), indicates that there is more than one type of single mutation—and therefore more than one
mechanism—by which the CQ transport activity of PfCRT can be abrogated. is insight extends our
understanding of the structure-function of PfCRT (e.g.32). Moreover, the nding that single mutations
to the protein can result in gross changes to parasite morphology emphasizes the central role of this
transporter in the physiological processes that occur within the FV and provides a novel insight into
one of the factors constraining the evolution of PfCRT. e observation that BSD binds to, and appears
to exert a selection pressure on, PfCRT further broadens the diversity of chemotypes that are known (or
suspected) to interact with the protein. Our data encourage further studies to dene agents that could
reverse antimalarial drug resistance mediated by PfCRT by inhibiting its function.
Methods
Antimalarials and reagents. CQ, QN, QD, MQ, AQ, PPQ, ART, BSD, AMT and VP were purchased
from Sigma Aldrich Chemical Co. MDAQ was purchased from Santa Cruz Biotechnology, Inc. MDCQ
was a gi from William Ellis (Walter Reed Army Institute of Research, Silver Spring, MD). SYBR Green
I was purchased from Invitrogen Corp. Drug stocks were prepared to 10 mM in DMSO or 70% ethanol
and stored below - 20 °C.
In vitro culture and selection of parasites. P. falciparum 3D7 and 3D7L272F parasites were cul-
tured in human RBC suspensions using RPMI 1640 medium (Sigma-Aldrich; Cat. No. R0883-500ML)
supplemented with 2 mM L-glutamine, 34 mM HEPES, 0.5% (w/v) Albumax I, 0.19 mM hypoxanthine,
and 50 μ g/ml gentamycin and maintained at 37 °C under 5% CO2. For parasite clone 3D7L272F, complete
medium was supplemented with 2.5 μ g/ml blasticidin-S HCl (Invitrogen). Parasite growth was followed
by microscopic examination of Field’s stained thin blood smears. Synchronization of early trophozoite
stages was achieved by incubating infected RBCs in 5% (w/v) sorbitol for 10 to 20 min at room temper-
ature50. Following transfection studies23, parasites with abnormally enlarged FVs, as described in results,
reappeared in culture under BSD pressure aer four weeks. In order to select these parasites, the lim-
iting dilution technique was used, and cloned parasites were identied by microscopy using thin blood
smears51.
P. falciparum FCB and FCBC101F parasites were cultured in AB+ or O+ human RBC suspensions using
RPMI 1640 medium (Mediatech, Inc.) supplemented with 0.5% Albumax I, 29.8 mM sodium bicarbo-
nate, 25 mM HEPES, 0.37 mM hypoxanthine, and 10 μ g/ml gentamicin and maintained at 37 °C under
an atmosphere of 90% N2, 5% CO2, and 5% O2. AMT-resistant P. falciparum was selected by single-step
selection based on an earlier method described for CQ6. Before drug pressure, parasites of the FCB strain
were grown to 5% mixed stage parasitemia at 5% hematocrit in 50 ml of media. is starter culture was
then split equally into 4 asks, with fresh media and RBCs to bring the volume in each ask to 50 ml
and 5% hematocrit. When parasitemia of the 4 asks had returned to 5%, the media was replaced with
fresh media containing 80 μ M AMT. At ~14 fold the IC50 value determined for FCB (Table 2), this
concentration of AMT rapidly kills CQ-resistant parasites. For the rst week aer drug application, cul-
tures were monitored daily by Giemsa-stained thin blood lms. Fresh AMT-containing media changes
were performed daily. At one week, 50% of the RBCs were replaced, and fresh AMT media was added.
Cultures were then maintained every third day with fresh AMT media for the duration of the experiment
and monitored by thin smear for emergent parasites. With every second media change, 50% of the RBCs
were replaced with fresh cells. If no surviving parasites were observed aer 60 days, the experiment was
terminated. Aer 42 days, parasites were recovered from one of the 4 asks, which were then cloned by
limiting dilution52 in drug-free media. e mixed culture and four randomly chosen cloned lines were
cryopreserved prior to DNA sequencing and drug susceptibility testing.
Morphological measurements. For comparison, thin blood lms of cultured parasite samples were
made at various time points following sorbitol synchronization and stained with Field’s stain. Pictures
were taken under the same conditions for the 3D7 (parent) strain and the 3D7L272F line and analyzed
www.nature.com/scientificreports/
12
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
with a Nikon Eclipse TE2000 inverted microscope. Areas were measured using ImageJ 1.44o soware
and the ratio was expressed as AFV/AParasite. No image manipulations were carried out aer recording.
For micrographs of FCB and FCBC101F, thin lms from parasite cultures were stained with 2% (v/v)
modied Giemsa (Karyomax® ; Gibco) for 30 min. Slides were washed for 60 s in owing distilled water,
air-dried and mounted with coverslips. Images were photographed in bright eld, using a Lexica DMI4000
inverted microscope under a 100X objective lens. Images were compiled in Adobe Photoshop CS5.1 and
processed equally with a warming photo lter. Live parasite cultures were placed under coverslips and
photographed under a 100X objective, using a Leica DM750 light microscope equipped with ICC50 HD
digital camera. Images were adjusted for white balance with the Leica Application Suite soware and
cropped in Adobe Photoshop CS5.1.
For experiments with Dd2 parasites, thin blood smears were xed with methanol, stained for 20 min
in 10% (v/v) Giemsa (Invitrogen), washed, and air-dried. Images were taken with an Olympus DP12
digital camera attached to an Olympus CX 41 light microscope with a 100X objective (N.A 1.4x). Images
were cropped and corrected for white balance using Adobe Lightroom 3.
Electron microscopy. Samples of 3D7, 3D7L272, Dd2Dd2 and Dd2Dd2 L272F, synchronized at the mature
trophozoite stage were xed in 4% (v/v) glutaraldehyde in 0.1 M phosphate buer and processed for
routine electron microscopy, as described previously53. Samples were post xed in osmium tetroxide,
treated en bloc with uranyl acetate, dehydrated and embedded in Spurr’s epoxy resin. in sections were
stained with uranyl acetate and lead citrate prior to examination in a JEOL1200EX electron microscope.
In vitro inhibition assays. Sensitivity to CQ and other drugs for 3D7 and 3D7L272F parasites was
determined by measurement of [3H]-hypoxanthine incorporation over 48 h, as described previously54.
Nine serial dilutions plus a control (no drug) were tested in quadruplicates and the experiment was
repeated at least three times for each drug. e assay was performed always in parallel on 3D7 and
3D7L272F parasites.
e in vitro susceptibility of FCB and the FCBC101F line of P. falciparum to antimalarial drugs was
measured in a 72 h, 96 well microplate uorescence assay using SYBR Green I detection as described55,56.
Drugs were serially diluted 2-fold in the microplates, except for AMT, which was diluted 3-fold. VP was
used at a concentration of 0.8 μ M where indicated. Synchronous (immature) ring-stage parasites were
assayed at 0.2% parasitemia and 2% hematocrit. Assays were conducted every 48 h until three independ-
ent replicates were performed. For Dd2 parasites, the same methodology was used except parasites were
also stained with 1.6 μ M Mito Tracker Deep Red.
Genotypic characterization of pfcrt gene. For 3D7 parasites, RNA was extracted from parasites
collected in RNAlater, using QIAGEN RNeasy Mini Kit, and immediately used to retro-transcribe cDNA
(QIAGEN, QuantiTect Rev. Transcription Kit). Mutation in pfcrt was investigated by PCR, as described
previously1. e same primers (Supplementary Table S1), which amplied overlapping products, were
used to sequence the products to cover the entire open reading frame (ORF) of the gene. Amplication
of the gene and its sequencing was performed twice (by Beckman Coulter Genomics). Alignment of
the reported 3D7 gene from PlasmoDB and 3D7 and 3D7L272F sequenced genes was performed using
MacVector soware (version 11.1).
For FCB parasites, 4 clonal lines of FCBC101F were used for pfcrt sequencing. All ORF sequences of
pfcrt were amplied from P. falciparum genomic DNA57. PCR products were sequenced directly using an
ABI 3730xl DNA analyzer (Applied Biosystems).
Whole genome sequencing and variant detection. Genomic DNA was isolated and prepared
from the parental P. falciparum parasite line 3D7 and 3D7L272F. A total of 10 μ g of gDNA from each line
was sheared to obtain a fragment size of ~200–400 bp using an E220 focused-ultrasonicator (Covaris)
with the following settings: 10% duty cycle, intensity 5, 200 cycles per burst, 180 s treatment length.
e resulting sheared gDNA was size selected on a 2% (w/v) low-melting agarose gel and then puried
and concentrated using MinElute purication columns followed by the QIAquick PCR purication kit
(QIAGEN). Barcoded libraries for Illumina TruSeq single-end sequencing were then constructed from
the size-selected, sheared material using NEBNext DNA Library Preparation reagents (New England
Biolabs) by following the standard Illumina (Illumina) library preparation protocol. Finally, barcoded
libraries were size selected using Agencourt AMPure XP magnetic beads (Agencourt Biosciences,
Beckman Coulter) thereby removing any adapter dimers and resulting in a highly enriched barcoded
library of 200–400 bp adapter-ligated fragments. e quality of the nal sequencing libraries was assessed
using an Agilent 2100 Bioanalyzer (Agilent Technologies) run alongside the original size-selected frag-
mented gDNA from the same preparation, and the concentration of each library was quantied using
a Quant-iT dsDNA Broad-Range Assay Kit (Invitrogen). e nal libraries were multiplexed with three
barcoded samples and 20% (v/v) PhiX control DNA (Illumina, Catalog # FC-110-3001) per lane and were
sequenced using an Illumina HiSeq 2500 Rapid Run (150 bp) system (Illumina).
Sequencing outputs were uploaded into Galaxy58, which is hosted locally at the Millennium Science
Complex at Pennsylvania State University. Sequence reads were mapped to the P. falciparum 3D7
reference genome v. 10.0 (http://plasmodb.org/common/downloads/release-10.0/Pfalciparum/) and
www.nature.com/scientificreports/
13
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
pCam-BSD-PfATP6-doubleHA plasmid sequence23, using the Burrows-Wheeler alignment tool59, and
les were converted to allow for further analysis (GATK/BAM-to-SAM)60. Sequence variations were
detected by Freebayes (version 0.9.0.a) using stringent ltering parameters based on quality and read
depth61,62. en SNPe (version 3.3) was applied to annotate and determine statistical signicance of
each variant63. Genome copy number variations were detected based upon local chromosomal read depth
using CNVnator (version 0.3) and annotated with Intansv (version 0.99.3)64. Alignments and variants
were visualized using the Integrative Genomics Viewer65. Unique reads were selected and ltered for
Map Quality > 30.
Plasmid construction and generation of Dd2 recombinant parasites. e donor plasmid pcrtDd2-
hdhfr has been previously reported31. e mutation-encoding plasmid, pcrtDd2 L272F-hdhfr, was generated
by site-directed mutagenesis of pcrtDd2-hdhfr, using primers p3527 + p3528 (Supplementary Table S2).
ZFN-editing transfection methods have been previously described31. Briey, Dd2 parasites were elec-
troporated with either pcrtDd2-hdhfr or pcrtDd2 L272F-hdhfr donor plasmid66. On Day 1 post-electroporation,
they were cultured in the presence of 2.5 nM WR99210 (obtained from Jacobus Pharmaceuticals Inc.).
Once recovered, both pcrtDd2-hdhfr and pcrtDd2 L272F-hdhfr transfected parasites were electroporated a
second time with pZFNcrt-bsd separately. On Day 1 post-electroporation each transfection was cultured
with 2 μ g/ml blasticidin S (Invitrogen) and 2.5 nM WR99210 for six days and followed by addition of
only 2.5 nM WR99210, generating Dd2Dd2and Dd2Dd2 L272F parasites, respectively. Clones were established
from the bulk cultures by limiting dilution52. PCR primers for verication of parental, recombinant con-
trol, and Dd2Dd2 L272F parasite clones are shown in Supplementary Table S2.
Expression of the C101F and L272F variants of PfCRT in X. laevis oocytes and measurements
of CQ transport. Ethical approval of the work performed with the X. laevis frogs was obtained from
the Australian National University Animal Experimentation Ethics Committee (Animal Ethics Protocol
Number A2013/13) in accordance with the Australian Code of Practice for the Care and Use of Animals
for Scientic Purposes. e C101F PfCRTDd2, L272F PfCRTDd2, and L272F PfCRT3D7 coding sequences
were generated via site-directed mutagenesis using an approach described previously32. e mutations
were introduced into codon-harmonized versions of the PfCRTDd2 and PfCRT3D7 coding sequences,
which encode retention motif-free forms of these proteins that are expressed at the plasma membrane
of X. laevis oocytes7. All of the resulting coding sequences were veried by sequencing. e in vitro tran-
scription of cRNA and the harvest and preparation of oocytes were performed as outlined elsewhere67.
e oocytes were microinjected with 20 ng of cRNA and the uptake of [3H]CQ (0.25 μ M; 20 Ci/mmol;
American Radiolabeled Chemicals) was measured 34 days post-injection as detailed previously67. e
measurements were made over 1.5 h at 27.5 °C and in medium that, unless otherwise specied, con-
tained 96 mM NaCl, 2 mM KCl, 2 mM MgCl2, 1.8 mM CaCl2, 10 mM MES, 10 mM Tris·base (pH 5.5),
and 15 μ M unlabeled CQ. In all cases, at least three separate experiments were performed (on oocytes
from dierent frogs), and in each experiment measurements were made from 10 oocytes per treatment.
Immunouorescence and western blot analyses of oocytes expressing PfCRT.
Immunouorescence analyses were performed on oocytes 3 days post-injection using an approach
described elsewhere68. Briey, the oocytes were xed and labeled with rabbit anti-PfCRT antibody (con-
centration of 1:100; Genscript32; ) and Alexa Fluor 488 goat anti-rabbit antibody (concentration of 1:500;
Molecular Probes). e oocytes were embedded in an acrylic resin using the Technovit 7100 plastic
embedding system (Kulzer) as outlined previously69 and images of 4 μ m slices were obtained with a Leica
Microsystems inverted confocal laser microscope. At least two separate experiments were performed (on
oocytes from dierent frogs) for each treatment and slices were examined from a minimum of three
oocytes within a treatment. All of the slices taken from oocytes expressing a PfCRT variant displayed
a uorescent band above the pigment layer (consistent with the localization of PfCRT to the plasma
membrane) that was not present in non-injected oocytes.
e preparation of oocyte membranes and the semi-quantication of PfCRT protein was carried out
using a protocol described in detail elsewhere32. Protein samples prepared from oocyte membranes were
separated on a 4–14% bis-Tris SDS-polyacrylamide gel (Life Technologies) and transferred to nitrocel-
lulose membranes. e membranes were probed with rabbit anti-PfCRT antibody (1:4,000) followed by
horseradish peroxidase-conjugated goat anti-rabbit antibody (1:8,000; Life Technologies). e PfCRT
band for each variant was detected by chemiluminescence (Pierce), quantied using the Image J so-
ware70, and expressed as a percentage of the intensity measured for the PfCRTDd2 band. In all cases, at
least three separate experiments were performed (on oocytes from dierent frogs), and in each experi-
ment measurements were averaged from two independent replicates.
Curve tting and statistical analyses. Mean half-maximal inhibitory concentrations (IC50 val-
ues) were derived by plotting percent growth inhibition against log drug concentration, and tting the
response data to a variable slope, sigmoidal curve-t function for normalized data using Prism 5.0d for
Macintosh (GraphPad Soware). IC50 values represent means ± standard error from 3 independent tests.
IC50 values between mutant and parent lines were tested for statistically signicant dierences using an
www.nature.com/scientificreports/
14
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
F-test that determines whether the two dose response data sets are best described by single or independ-
ent curve ts (p < 0.05). In the case of the oocyte data, statistical comparisons were made using ANOVA
in conjunction with Tukey’s multiple comparisons test. Other data were compared using the Student’s
t-test and Fisher’s exact test as noted.
References
1. Fidoc, D. A. et al. Mutations in the P. falciparum digestive vacuole transmembrane protein PfCT and evidence for their role
in chloroquine resistance. Mol Cell 6, 861–871 (2000).
2. Sidhu, A. B., Verdier-Pinard, D. & Fidoc, D. A. Chloroquine resistance in Plasmodium falciparum malaria parasites conferred
by pfcr t mutations. Science 298, 210–213 (2002).
3. Su, X., irman, L. A., Fujioa, H. & Wellems, T. E. Complex polymorphisms in an approximately 330 Da protein are lined
to chloroquine-resistant P. falciparum in Southeast Asia and Africa. Cell 91, 593–603 (1997).
4. Price, . N. et al. Global extent of chloroquine-resistant Plasmodium vivax: a systematic review and meta-analysis. Lancet Infect
Dis 14, 982–991 (2014).
5. Chou, A. C., Chevli, . & Fitch, C. D. Ferriprotoporphyrin IX fullls the criteria for identication as the chloroquine receptor
of malaria parasites. Biochemistry 19, 1543–1549 (1980).
6. Cooper, . A. et al. Alternative mutations at position 76 of the vacuolar transmembrane protein PfCT are associated with
chloroquine resistance and unique stereospecic quinine and quinidine responses in Plasmodium falciparum. Mol Pharmacol 61,
35–42 (2002).
7. Martin, . E. et al. Chloroquine transport via the malaria parasite's chloroquine resistance transporter. Science 325, 1680–1682
(2009).
8. Ecer, A., Lehane, A. M., Clain, J. & Fidoc, D. A. PfCT and its role in antimalarial drug resistance. Trends Parasitol 28, 504–514
(2012).
9. Picot, S. et al. A systematic review and meta-analysis of evidence for correlation between molecular marers of parasite resistance
and treatment outcome in falciparum malaria. Malar J 8, 89 (2009).
10. Wellems, T. E. & Plowe, C. V. Chloroquine-resistant malaria. J Infect Dis 184, 770–776 (2001).
11. Bray, P. G. et al. Dening the role of PfCT in Plasmodium falciparum chloroquine resistance. Mol Microbiol 56, 323–333 (2005).
12. Martin, . E. & ir, . e malaria parasite's chloroquine resistance transporter is a member of the drug/metabolite transporter
superfamily. Mol Biol Evol 21, 1938–1949 (2004).
13. Lewis, I. A. et al. Metabolic QTL analysis lins chloroquine resistance in Plasmodium falciparum to impaired hemoglobin
catabolism. PLoS Genet 10, e1004085 (2014).
14. Teng, . et al. 1H-NM metabolite proles of dierent strains of Plasmodium falciparum. Biosci ep 34, e00150 (2014).
15. Juge, N. et al. Plasmodium falciparum chloroquine resistance transporter is a H+-coupled polyspecic nutrient and drug exporter.
Proc Natl Acad Sci USA 112, 3356–3361 (2015).
16. Egan, T. J. & uter, D. Dual-functioning antimalarials that inhibit the chloroquine-resistance transporter. Future Microbiol 8,
475–489 (2013).
17. Summers, . L., Nash, M. N. & Martin, . E. now your enemy: understanding the role of PfCT in drug resistance could lead
to new antimalarial tactics. Cell Mol Life Sci 69, 1967–1995 (2012).
18. Burgess, S. J. et al. A chloroquine-lie molecule designed to reverse resistance in Plasmodium falciparum. J Med Chem 49,
5623–5625 (2006).
19. Frosch, A. E., Venatesan, M. & Laufer, M. . Patterns of chloroquine use and resistance in sub-Saharan Africa: a systematic
review of household survey and molecular data. Malar J 10, 116 (2011).
20. del Pilar Crespo, M. et al. Artemisinin and a series of novel endoperoxide antimalarials exert early eects on digestive vacuole
morphology. Antimicrob Agents Chemother 52, 98–109 (2008).
21. Pandey, A. V., Tewani, B. L., Singh, . L. & Chauhan, V. S. Artemisinin, an endoperoxide antimalarial, disrupts the hemoglobin
catabolism and heme detoxication systems in malarial parasite. J Biol Chem 274, 19383–19388 (1999).
22. Eastman, . T., Dharia, N. V., Winzeler, E. A. & Fidoc, D. A. Piperaquine resistance is associated with a copy number variation
on chromosome 5 in drug-pressured Plasmodium falciparum parasites. Antimicrob Agents Chemother 55, 3908–3916 (2011).
23. Pulcini, S. et al. Expression in yeast lins eld polymorphisms in PfATP6 to in vitro artemisinin resistance and identies new
inhibitor classes. J Infect Dis 208, 468–478 (2013).
24. Bopp, S. E. et al. Mitotic evolution of Plasmodium falciparum shows a stable core genome but recombination in antigen families.
PLoS Genet 9, e1003293 (2013).
25. Berendt, A. ., Ferguson, D. J. & Newbold, C. I. Sequestration in Plasmodium falciparum malaria: sticy cells and sticy problems.
Parasitol Today 6, 247–254 (1990).
26. Langreth, S. G., eese, . T., Motyl, M. . & Trager, W. Plasmodium falciparum: loss of nobs on the infected erythrocyte surface
aer long-term cultivation. Exp Parasitol 48, 213–219 (1979).
27. Hill, D. A. et al. A blasticidin S-resistant Plasmodium falciparum mutant with a defective plasmodial surface anion channel. Proc
Natl Acad Sci USA 104, 1063–1068 (2007).
28. Mira-Martinez, S. et al. Epigenetic switches in clag3 genes mediate blasticidin S resistance in malaria parasites. Cell Microbiol 15,
1913–1923 (2013).
29. Sharma, P. et al. An epigenetic antimalarial resistance mechanism involving parasite genes lined to nutrient uptae. J Biol Chem
288, 19429–19440 (2013).
30. Martin, S. ., Oduola, A. M. & Milhous, W. . eversal of chloroquine resistance in Plasmodium falciparum by verapamil. Science
235, 899–901 (1987).
31. Straimer, J. et al. Site-specic genome editing in Plasmodium falciparum using engineered zinc-nger nucleases. Nat Methods 9,
993–998 (2012).
32. Summers, . L. et al. Diverse mutational pathways converge on saturable chloroquine transport via the malaria parasite's
chloroquine resistance transporter. Proc Natl Acad Sci USA 111, E1759–1767 (2014).
33. Bignami, A., Marchiafava, E. & Mannaberg, J. Two Monographs on Malaria and the Parasites of Malarial Fevers. I. Marchicafaca
and Bignami. II. Mannaberg. e New Sydenham Society (1894).
34. Gligorijevic, B., Bennett, T., McAllister, ., Urbach, J. S. & oepe, P. D. Spinning dis confocal microscopy of live, intraerythrocytic
malarial parasites. 2. Altered vacuolar volume regulation in drug resistant malaria. Biochemistry 45, 12411–12423 (2006).
35. osenthal, P. J., Mcerrow, J. H., Aiawa, M., Nagasawa, H. & Leech, J. H. A malarial cysteine proteinase is necessary for
hemoglobin degradation by Plasmodium falciparum. J Clin Invest 82, 1560–1566 (1988).
36. ublin, J. G. et al. eemergence of chloroquine-sensitive Plasmodium falciparum malaria aer cessation of chloroquine use in
Malawi. J Infect Dis 187, 1870–1875 (2003).
37. Ord, . et al. Seasonal carriage of pfcrt and pfmdr1 alleles in Gambian Plasmodium falciparum imply reduced tness of
chloroquine-resistant parasites. J Infect Dis 196, 1613–1619 (2007).
www.nature.com/scientificreports/
15
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
38. Wang, X. et al. Decreased prevalence of the Plasmodium falciparum chloroquine resistance transporter 76T marer associated
with cessation of chloroquine use against P. falciparum malaria in Hainan, People's epublic of China. Am J Trop Med Hyg 72,
410–414 (2005).
39. Evans, S. G. & Havli, I. Plasmodium falciparum: eects of amantadine, an antiviral, on chloroquine-resistant and -sensitive
parasites in vitro and its inuence on chloroquine activity. Biochem Pharmacol 45, 1168–1170 (1993).
40. ornhuber, J. et al. Lipophilic cationic drugs increase the permeability of lysosomal membranes in a cell culture system. J Cell
Physiol 224, 152–164 (2010).
41. Grin, C. E. et al. Mutation in the Plasmodium falciparum CT protein determines the stereospecic activity of antimalarial
cinchona alaloids. Antimicrob Agents Chemother 56, 5356–5364 (2012).
42. Johnson, D. J. et al. Evidence for a central role for PfCT in conferring Plasmodium falciparum resistance to diverse antimalarial
agents. Mol Cell 15, 867–877 (2004).
43. Bray, P. G., Hawley, S. ., Mungthin, M. & Ward, S. A. Physicochemical properties correlated with drug resistance and the
reversal of drug resistance in Plasmodium falciparum. Mol Pharmacol 50, 1559–1566 (1996).
44. Warhurst, D. C. et al. Activity of piperaquine and other 4-aminoquinoline antiplasmodial drugs against chloroquine-sensitive
and resistant blood-stages of Plasmodium falciparum. ole of beta-haematin inhibition and drug concentration in vacuolar
water- and lipid-phases. Biochem Pharmacol 73, 1910–1926 (2007).
45. Haynes, . ., Cheu, . W., N'Da, D., Coghi, P. & Monti, D. Considerations on the mechanism of action of artemisinin
antimalarials: Part 1 - e 'carbon radical' and 'heme' hypotheses. Infect Disord Drug Targets 13, 217–277 (2013).
46. Ecstein-Ludwig, U. et al. Artemisinins target the SECA of Plasmodium falciparum. Nature 424, 957–961 (2003).
47. Mbengue, A. et al. A molecular mechanism of artemisinin resistance in Plasmodium falciparum malaria. Nature 520, 683–687
(2015).
48. Lin, J. W. et al. eplication of Plasmodium in reticulocytes can occur without hemozoin formation, resulting in chloroquine
resistance. J Exp Med 212, 893–903 (2015).
49. Ehlgen, F., Pham, J. S., de oning-Ward, T., Cowman, A. F. & alph, S. A. Investigation of the Plasmodium falciparum food
vacuole through inducible expression of the chloroquine resistance transporter (PfCT). PLoS One 7, e38781 (2012).
50. Lambros, C. & Vanderberg, J. P. Synchronization of Plasmodium falciparum erythrocytic stages in culture. J Parasitol 65, 418–420
(1979).
51. osario, V. Cloning of naturally occurring mixed infections of malaria parasites. Science 212, 1037–1038 (1981).
52. Goodyer, I. D. & Taraschi, T. F. Plasmodium falciparum: a simple, rapid method for detecting parasite clones in microtiter plates.
Exp Parasitol 86, 158–160 (1997).
53. Ferguson, D. J. et al. Maternal inheritance and stage-specic variation of the apicoplast in Toxoplasma gondii during development
in the intermediate and denitive host. Euaryot Cell 4, 814–826 (2005).
54. Desjardins, . E., Caneld, C. J., Haynes, J. D. & Chulay, J. D. Quantitative assessment of antimalarial activity in vitro by a
semiautomated microdilution technique. Antimicrob Agents Chemother 16, 710–718 (1979).
55. Bennett, T. N. et al. Novel, rapid, and inexpensive cell-based quantication of antimalarial drug ecacy. Antimicrob Agents
Chemother 48, 1807–1810 (2004).
56. Smilstein, M., Sriwilaijaroen, N., elly, J. X., Wilairat, P. & iscoe, M. Simple and inexpensive uorescence-based technique for
high-throughput antimalarial drug screening. Antimicrob Agents Chemother 48, 1803–1806 (2004).
57. Vieira, P. P. et al. pfcrt polymorphism and the spread of chloroquine resistance in Plasmodium falciparum populations across the
Amazon Basin. J Infect Dis 190, 417–424 (2004).
58. Goecs, J., Neruteno, A. & Taylor, J. Galaxy: a comprehensive approach for supporting accessible, reproducible, and transparent
computational research in the life sciences. Genome Biol 11, 86 (2010).
59. Li, H. & Durbin, . Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754–1760
(2009).
60. Li, H. et al. e Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).
61. DePristo, M. A. et al. A framewor for variation discovery and genotyping using next-generation DNA sequencing data. Nat
Genet 43, 491–498 (2011).
62. Garrison, E. & Marth, G. Haplotype-based variant detection from short-read sequencing. arXiv:12073907 [q-bioGN], (2012).
63. Cingolani, P. et al. A program for annotating and predicting the eects of single nucleotide polymorphisms, SnpE: SNPs in the
genome of Drosophila melanogaster strain w1118; iso-2; iso-3. Fly (Austin) 6, 80–92 (2012).
64. Abyzov, A., Urban, A. E., Snyder, M. & Gerstein, M. CNVnator: an approach to discover, genotype, and characterize typical and
atypical CNVs from family and population genome sequencing. Genome es 21, 974–984 (2011).
65. obinson, J. T. et al. Integrative genomics viewer. Nat Biotechnol 29, 24–26 (2011).
66. Fidoc, D. A., Nomura, T. & Wellems, T. E. Cycloguanil and its parent compound proguanil demonstrate distinct activities
against Plasmodium falciparum malaria parasites transformed with human dihydrofolate reductase. Mol Pharmacol 54, 1140–1147
(1998).
67. Bellanca, S. et al. Multiple drugs compete for transport via the Plasmodium falciparum chloroquine resistance transporter at
distinct but interdependent sites. J Biol Chem 289, 36336–36351 (2014).
68. Broer, S. Xenopus laevis Oocytes. Methods Mol Biol 637, 295–310 (2010).
69. Broer, S. et al. Comparison of lactate transport in astroglial cells and monocarboxylate transporter 1 (MCT 1) expressing Xenopus
laevis oocytes. Expression of two dierent monocarboxylate transporters in astroglial cells and neurons. J Biol Chem 272,
30096–30102 (1997).
70. Schneider, C. A., asband, W. S. & Eliceiri, . W. NIH Image to ImageJ: 25 years of image analysis. Nat Methods 9, 671–675
(2012).
71. Baro, N. ., Callaghan, P. S. & oepe, P. D. Function of resistance conferring Plasmodium falciparum chloroquine resistance
transporter isoforms. Biochemistry 52, 4242–4249 (2013).
Acknowledgments
is work was supported by the European Community’s Seventh Framework Programme, FP7/2007-2013
(Marie Curie-funded Initial Training Network InterMal, 215281-2 to SK and NanoMal, 304948 to SK
and HMS), the NIH (R01 AI50234 to DAF and R01 AI071121 to RAC), the Australian National Health
and Medical Research Council (Grant 1007035 and Fellowship 1053082 to REM) and the Burroughs
Wellcome Fund - Investigators in Pathogenesis of Infectious Disease award (1007041.02 to ML).
Author Contributions
Experimentation was undertaken by S.P., A.H.L., S.H.S., G.B., C.M.M., D.A.D., M.J.H., L.M.A., H.J.P., J.M.
and D.J.P.F. and designed by H.M.S., D.J.P.F., M.L., R.E.M., D.A.F., R.A.C. and S.K. e manuscript was
www.nature.com/scientificreports/
16
SCIENTIFIC RepoRts | 5:14552 | DOI: 10.1038/srep14552
prepared by S.P., H.M.S., A.H.L., M.L., R.E.M., D.A.F., R.A.C. and S.K. All authors had the opportunity
to read and approve the manuscript.
Additional Information
Supplementary information accompanies this paper at http://www.nature.com/srep
Competing nancial interests: e authors declare no competing nancial interests.
How to cite this article: Pulcini, S. et al. Mutations in the Plasmodium falciparum chloroquine
resistance transporter, PfCRT, enlarge the parasite's food vacuole and alter drug sensitivities. Sci. Rep.
5, 14552; doi: 10.1038/srep14552 (2015).
is work is licensed under a Creative Commons Attribution 4.0 International License. e
images or other third party material in this article are included in the article’s Creative Com-
mons license, unless indicated otherwise in the credit line; if the material is not included under the
Creative Commons license, users will need to obtain permission from the license holder to reproduce
the material. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/
... Of the transporters present in FVs, chloroquine resistance transporter (CRT) and multidrug resistance protein 1 (MDR1) have been studied in detail because of their demonstrated association with antimalarial resistance (29,30). Various studies have characterized genetic polymorphisms of CRT and MDR1 in the context of drug resistance, parasite fitness, and FV functionality (31,32). It has also been shown that PfCRT transports host-derived peptides of 4 to 11 residues (33). ...
... Altogether, our results suggested that PbAAT1 might transport Hbderived peptides from the FV lumen to the parasite cytosol. A similar swollen FV phenotype has been reported for amantadine-and blasticidin-selected P. falciparum parasites carrying mutations in CRT and for P. falciparum parasites with reduced expression of CRT (31). Numerous factors, including pH, lipid composition, solubility of FV heme, integrity of FV, etc., might affect Hz formation and its morphology (42). ...
Article
Full-text available
The food vacuole plays a central role in the blood stage of parasite development by digesting host hemoglobin acquired from red blood cells and detoxifying the host heme released during hemoglobin digestion into hemozoin. Blood-stage parasites undergo periodic schizont bursts, releasing food vacuoles containing hemozoin. Clinical studies in malaria-infected patients and in vivo animal studies have shown the association of hemozoin with disease pathogenesis and abnormal host immune responses in malaria. Here, we perform a detailed in vivo characterization of putative Plasmodium berghei amino acid transporter 1 localized in the food vacuole to understand its significance in the malaria parasite. We show that the targeted deletion of amino acid transporter 1 in Plasmodium berghei leads to a swollen food vacuole phenotype with the accumulation of host hemoglobin-derived peptides. Plasmodium berghei amino acid transporter 1-knockout parasites produce less hemozoin, and the hemozoin crystals display a thin morphology compared with wild-type parasites. The knockout parasites show reduced sensitivity to chloroquine and amodiaquine by showing recrudescence. More importantly, mice infected with the knockout parasites are protected from cerebral malaria and display reduced neuronal inflammation and cerebral complications. Genetic complementation of the knockout parasites restores the food vacuole morphology with hemozoin levels similar to that of wild-type parasites, causing cerebral malaria in the infected mice. The knockout parasites also show a significant delay in male gametocyte exflagellation. Our findings highlight the significance of amino acid transporter 1 in food vacuole functionality and its association with malaria pathogenesis and gametocyte development. IMPORTANCE Food vacuoles of the malaria parasite are involved in the degradation of red blood cell hemoglobin. The amino acids derived from hemoglobin degradation support parasite growth, and the heme released is detoxified into hemozoin. Antimalarials such as quinolines target hemozoin formation in the food vacuole. Food vacuole transporters transport hemoglobin-derived amino acids and peptides from the food vacuole to the parasite cytosol. Such transporters are also associated with drug resistance. Here, we show that the deletion of amino acid transporter 1 in Plasmodium berghei leads to swollen food vacuoles with the accumulation of hemoglobin-derived peptides. The transporter-deleted parasites generate less hemozoin with thin crystal morphology and show reduced sensitivity to quinolines. Mice infected with transporter-deleted parasites are protected from cerebral malaria. There is also a delay in male gametocyte exflagellation, affecting transmission. Our findings uncover the functional significance of amino acid transporter 1 in the life cycle of the malaria parasite.
... Oxford University achieved the highest number of publications and the highest centrality ranking, with a total of 184 papers published. These papers delve into the evaluation of resistance activity in various regions, as well as the exploration of resistance mechanisms in P. falciparum and P. vivax (Pulcini et al., 2015;Yogavel et al., 2018). The study found that Fidock A. David, affiliated with Columbia University Medical Center in the USA, had published the most papers and had the strongest influence Map illustrating networks of cooperation between countries. ...
Article
Full-text available
Background Malaria has always been a serious infectious disease prevalent in the world. Antimalarial drugs such as chloroquine and artemisinin have been the main compounds used to treat malaria. However, the massive use of this type of drugs accelerates the evolution and spread of malaria parasites, leading to the development of resistance. A large number of related data have been published by researchers in recent years. CiteSpace software has gained popularity among us researchers in recent years, because of its ability to help us obtain the core information we want in a mass of articles. In order to analyze the hotspots and develop trends in this field through visual analysis, this study used CiteSpace software to summarize the available data in the literature to provide insights. Method Relevant literature was collected from the Web of Science Core Collection (WOSCC) from 1 January 2015 to 29 March 2023. CiteSpace software and Microsoft Excel were used to analyze and present the data, respectively. Results A total of 2,561 literatures were retrieved and 2,559 literatures were included in the analysis after the removal of duplicates. An irrefutable witness of the ever-growing interest in the topic of antimalarial drug resistance could be expressed by the exponentially increased number of publications and related citations from 2015 to 2022, and its sustained growth trend by 2023. During the past 7 years, USA, Oxford University, and David A Fidock are the country, institution, and author with the most publications in this field of research, respectively. We focused on the references and keywords from literature and found that the research and development of new drugs is the newest hotspot in this field. A growing number of scientists are devoted to finding new antimalarial drugs. Conclusion This study is the first visual metrological analysis of antimalarial drug resistance, using bibliometric methods. As a baseline information, it is important to analyze research output published globally on antimalarial drug resistance. In order to better understand the current research situation and future research plan agenda, such baseline data are needed accordingly.
... Evidence suggests that these mutations enable the transporter to efflux positively charged PPQ out of its site of action in the DV (17)(18)(19), resulting in parasite survival under high drug concentra tions. PPQ-resistant (PPQ-R) PfCRT mutations have also been linked to alterations in Hb metabolism, DV swelling, and decreased parasite fitness (12,15,16,20,21). ...
Article
Full-text available
Malaria elimination efforts in Southeast Asia have been hindered by multidrug-resistant Plasmodium falciparum. High-grade resistance to piperaquine (PPQ, used in combination with dihydroartemisinin) is associated with PfCRT mutations that arose in strains expressing the PfCRT Dd2 isoform, which mediates resistance to the related 4-aminoquinoline chloroquine (CQ). The PPQ-resistant PfCRT haplotype Dd2 + F145I mediates the highest level resistance but causes a significant growth defect in intra-erythrocytic parasites. Recently, three separate mutations (F131C, I347T and C258W) have been observed on Dd2 + F145I PfCRT either during extended parasite culture or in Southeast Asian isolates no longer subject to PPQ pressure. Competitive growth assays with pfcrt-edited parasites reveal that these compensatory mutations reduce the fitness defect caused by F145I. PPQ survival assays on edited lines show a loss of PPQ resistance in two of the three variants, including the field mutant (C258W). The latter restores CQ resistance. None of these variants alter parasite susceptibility to the first-line partner drug, mefloquine. Utilizing drug transport assays with purified PfCRT isoforms reconstituted into proteoliposomes, we identify differences in mutant PfCRT-mediated transport of PPQ and CQ. Molecular dynamics energy minimization calculations predict that these same mutations cause small but significant conformational changes in PfCRT regions implicated in drug interactions. Metabolomic analyses of isogenic parasite lines reveal differences in hemoglobin-derived peptide accumulation as a hallmark of PfCRT variation. These studies highlight the transient nature of PPQ resistance upon removal of drug pressure and suggest a strategy for employing this drug as part of multiple first-line therapies. IMPORTANCE Our study leverages gene editing techniques in Plasmodium falciparum asexual blood stage parasites to profile novel mutations in mutant PfCRT, an important mediator of piperaquine resistance, which developed in Southeast Asian field isolates or in parasites cultured for long periods of time. We provide evidence that increased parasite fitness of these lines is the primary driver for the emergence of these PfCRT variants. These mutations differentially impact parasite susceptibility to piperaquine and chloroquine, highlighting the multifaceted effects of single point mutations in this transporter. Molecular features of drug resistance and parasite physiology were examined in depth using proteoliposome-based drug uptake studies and peptidomics, respectively. Energy minimization calculations, showing how these novel mutations might impact the PfCRT structure, suggested a small but significant effect on drug interactions. This study reveals the subtle interplay between antimalarial resistance, parasite fitness, PfCRT structure, and intracellular peptide availability in PfCRT-mediated parasite responses to changing drug selective pressures.
... Resistance to the antimalarial drug chloroquine (CQ) first appeared in Plasmodium falciparum in the late 1950s. 13,16 Resistance to CQ in P. falciparum is mediated mainly by PfCRT (Plasmodium falciparum Chloroquine Resistance Transporter) which is a trans membrane protein present on the digestive vacuole membrane of the parasite, the mutant isoforms of this protein tend to efflux the drugs out of the digestive vacuole. A few mutant isoforms of PfMDR1 (Plasmodium falciparum Multidrug-Resistant 1) have been found to contribute chloroquine resistance in recent studies. ...
Article
The increasing availability of drug-resistant Plasmodium falciparum infections is putting a strain on the accessibility of potent, safe and cost-effective anti-malarial treatments, necessitating the development of new anti-malarial drug. Malaria deaths were estimated to be around 409000 in 2019. The present study sets to identify novel antimalarial compounds and the virtual screening study reveals that the designed compounds bind more effectively to Plasmodium falciparum chloroquine resistance transporter (PfCRT) and Plasmodium falciparum Multidrug Resistant1 (PfMDR1) than the known inhibitors. Marvin JS was used to design the chemical structure of the molecules and the molecular docking of 75 designed molecules with PfCRT and PfMDR1 was performed to study the interaction between the small molecule and the proteins. The top docked scoring compounds with the respective proteins were subjected to molecular dynamic simulation to study their interaction stability. The ADME/T (absorption, distribution, metabolism and excretion/toxicity) properties of those molecules were also studied and the majority of the properties was found to be within acceptable ranges. After experimental validation to confirm the findings, the screened molecules could be used as potential anti-malarial drugs.
... However, DAQ's IC 50 is not affected by the presence of verapamil, which suggests that DAQ is not competing with verapamil for efflux mechanisms induced by the parasite's membrane transporters. More experiments are needed to test the hypothesis that DAQ is not a substrate of PfCRT, e.g., (a) using isogenic parasite cell lines expressing either the wild-type PfCRT or the CQ-resistant associated PfCRT isoforms and (b) expressing PfCRT in a heterologous system and measuring the ability of DAQ to inhibit [3H]CQ transport via PfCRT (Pulcini et al., 2015). ...
Article
Malaria is among the tropical diseases that cause the most deaths in Africa. Around 500,000 malaria deaths are reported yearly among African children under the age of five. Chloroquine (CQ) is a low-cost antimalarial used worldwide for the treatment of Plasmodium vivax malaria. Due to resistance mechanisms, CQ is no longer effective against most malaria cases caused by P. falciparum. The World Health Organization recommends artemisinin combination therapies for P. falciparum malaria, but resistance is emerging in Southeast Asia and some parts of Africa. Therefore, new medicines for treating malaria are urgently needed. Previously, our group identified the 4-aminoquinoline DAQ, a CQ analog containing an acetylenic bond in its side chain, which overcomes CQ resistance in K1 P. falciparum strains. In this work, the antiplasmodial profile, drug-like properties, and pharmacokinetics of DAQ were further investigated. DAQ showed no cross-resistance against standard CQ-resistant strains (e.g., Dd2, IPC 4912, RF12) nor against P. falciparum and P. vivax isolates from patients in the Brazilian Amazon. Using drug pressure assays, DAQ showed a low propensity to generate resistance. DAQ showed considerable solubility but low metabolic stability. The main metabolite was identified as a mono N-deethylated derivative (DAQM), which also showed significant inhibitory activity against CQ-resistant P. falciparum strains. Our findings indicated that the presence of a triple bond in CQ-analogues may represent a low-cost opportunity to overcome known mechanisms of resistance in the malaria parasite.
Article
Full-text available
Genes associated with drug resistance of first line drugs for Plasmodium falciparum have been identified and characterized of which three genes most commonly associated with drug resistance are P. falciparum chloroquine resistance transporter gene (PfCRT), P. falciparum multidrug drug resistance gene 1 (PfMDR1), and P. falciparum Kelch protein K13 gene (PfKelch13). Polymorphism in these genes could be used as molecular markers for identifying drug resistant strains. Nucleic acid amplification test (NAAT) along with DNA sequencing is a powerful diagnostic tool that could identify these polymorphisms. However, current NAAT and DNA sequencing technologies require specific instruments which might limit its application in rural areas. More recently, a combination of isothermal amplification and CRISPR detection system showed promising results in detecting mutations at a nucleic acid level. Moreover, the Loop-mediated isothermal amplification (LAMP)-CRISPR systems offer robust and straightforward detection, enabling it to be deployed in rural and remote areas. The aim of this study was to develop a novel diagnostic method, based on LAMP of targeted genes, that would enable the identification of drug-resistant P. falciparum strains. The methods were centered on sequence analysis of P. falciparum genome, LAMP primers design, and CRISPR target prediction. Our designed primers are satisfactory for identifying polymorphism associated with drug resistant in PfCRT, PfMDR1, and PfKelch13. Overall, the developed system is promising to be used as a detection method for P. falciparum treatment-resistant strains. However, optimization and further validation the developed CRISPR-LAMP assay are needed to ensure its accuracy, reliability, and feasibility.
Article
Background: Lumefantrine (LM), piperaquine (PQ), and amodiaquine (AQ) are the essential long-acting partner drugs in the artemisinin-based combination therapies (ACTs) treatment regimens globally. The recent report on the emergence of artemisinin-resistant parasites portends an imminent failure of the partner drug in clearing the high residual parasite densities. Understanding the resistance mechanisms to partner drugs remains critical for tracking resistant parasites. Cysteine desulfurase IscS ( nfs1 ), one of the proteins involved in the iron-sulfur (FeS) biogenesis pathway, has been implicated in mediating malaria parasite drug resistance. Methods: Using the rodent malaria parasites Plasmodium berghei ANKA in mice, we assessed whether the nfs1 gene is associated with LM, PQ, and AQ resistance. We first verified the stability of the LM, PQ, and AQ-resistant parasites in the standard 4-Day Suppressive Test. By means of PCR and sequencing analysis, we probed for single nucleotide polymorphisms (SNPs) in the nfs1 gene. Using qPCR, we then measured the expression of the nfs1 gene in resistant parasites relative to the drug-sensitive parent parasites. Results: Our analyses of nfs1 reveal a non-synonymous Gln142Arg mutation in the LM and PQ-resistant parasites. This mutation was not detected in the AQ-resistant parasites. The mRNA quantification of the nfs1 gene reveals differential expression in both LM and PQ-resistant parasites. Conversely, nfs1 expression remained unchanged in the AQ-resistant parasites. Conclusion: Our data suggest that LM and PQ selection pressure induces nonsynonymous mutation and differential expression of the nfs1 gene in Plasmodium berghei . Collectively, these findings provide a premise for investigating LM and PQ resistance mechanisms in both P. berghei and P. falciparum .
Article
Background: Lumefantrine (LM), piperaquine (PQ), and amodiaquine (AQ) are the essential long-acting partner drugs in the artemisinin-based combination therapies (ACTs) treatment regimens globally. The recent report on the emergence of artemisinin-resistant parasites portends an imminent failure of the partner drug in clearing the high residual parasite densities. Understanding the resistance mechanisms to partner drugs remains critical for tracking resistant parasites. Cysteine desulfurase IscS ( nfs1 ), one of the proteins involved in the iron-sulfur (FeS) biogenesis pathway, has been implicated in mediating malaria parasite drug resistance. Methods: Using the rodent malaria parasites Plasmodium berghei ANKA in mice, we assessed whether the nfs1 gene is associated with LM, PQ, and AQ resistance. We first verified the stability of the LM, PQ, and AQ-resistant parasites in the standard 4-Day Suppressive Test. By means of PCR and sequencing analysis, we probed for single nucleotide polymorphisms (SNPs) in the nfs1 gene. Using qPCR, we then measured the expression of the nfs1 gene in resistant parasites relative to the drug-sensitive parent parasites. Results: Our analyses of nfs1 reveal a non-synonymous Gln142Arg mutation in the LM and PQ-resistant parasites. This mutation was not detected in the AQ-resistant parasites. The mRNA quantification of the nfs1 gene reveals differential expression in both LM and PQ-resistant parasites. Conversely, nfs1 expression remained unchanged in the AQ-resistant parasites. Conclusion: Our data suggest that LM and PQ selection pressure induces nonsynonymous mutation and differential expression of the nfs1 gene in Plasmodium berghei . Collectively, these findings provide a premise for investigating LM and PQ resistance mechanisms in both P. berghei and P. falciparum .
Article
Background: Lumefantrine (LM), piperaquine (PQ), and amodiaquine (AQ), the long-acting components of the artemisinin-based combination therapies (ACTs), are a cornerstone of malaria treatment in Africa. Studies have shown that PQ, AQ, and LM resistance may arise independently of predicted modes of action. Protein kinases have emerged as mediators of drug action and efficacy in malaria parasites; however, the link between top druggable Plasmodium kinases with LM, PQ, and AQ resistance remains unclear. Using LM, PQ, or AQ-resistant Plasmodium berghei parasites, we have evaluated the association of choline kinase (CK), pantothenate kinase 1 (PANK1), diacylglycerol kinase (DAGK), and phosphatidylinositol-4 kinase (PI4Kβ), and calcium-dependent protein kinase 1 (CDPK1) with LM, PQ, and AQ resistance in Plasmodium berghei ANKA. Methods: We used in silico bioinformatics tools to identify ligand-binding motifs, active sites, and sequence conservation across the different parasites. We then used PCR and sequencing analysis to probe for single nucleotide polymorphisms (SNPs) within the predicted functional motifs in the CK, PANK1, DAGK, PI4Kβ, and CDPK1. Using qPCR analysis, we finally measured the mRNA amount of PANK1, DAGK, and PI4Kβ at trophozoites and schizonts stages. Results: We reveal sequence conservation and unique ligand-binding motifs in the CK, PANK1, DAGK, PI4Kβ, and CDPK1 across malaria species. DAGK, PANK1, and PI4Kβ possessed nonsynonymous mutations; surprisingly, the mutations only occurred in the AQr parasites. PANK1 acquired Asn394His, while DAGK contained K270R and K292R mutations. PI4Kβ had Asp366Asn, Ser1367Arg, Tyr1394Asn and Asp1423Asn. We show downregulation of PANK1, DAGK, and PI4Kβ in the trophozoites but upregulation at the schizonts stages in the AQr parasites. Conclusions: The selective acquisition of the mutations and the differential gene expression in AQ-resistant parasites may signify proteins under AQ pressure. The role of the mutations in the resistant parasites and the impact on drug responses require further investigations in malaria parasites.
Thesis
Le paludisme est une maladie infectieuse parasitaire causée par diverses espèces de Plasmodium, P. falciparum étant l'espèce la plus répandue et responsable des cas mortels de la maladie. Les traitements actuels reposent sur des combinaisons thérapeutiques à base d'artémisinine (CTA), couplant un dérivé d'artémisinine (ARTD) à une autre molécule antipaludique. Des parasites résistants aux ARTDs ont émergé en Asie du sud-est ; ceux-ci sont encore non détectés en Afrique. La résistance se traduit par une durée d'élimination des parasites allongée, et est conférée par des mutations non-synonymes localisées sur le domaine Kelch-repeat propeller (KREP) de la protéine P. falciparum K13 (PfK13). Similairement, de multiples mutations non-synonymes sur le transporteur P. falciparum Chloroquine Resistance Transporter (PfCRT) confèrent la résistance à la chloroquine (CQ, ancien traitement) et à la pipéraquine (PPQ, une des molécules partenaires dans les CTAs actuels). Le rôle physiologique de ces protéines, essentielles durant le développement du parasite, demeure mal connu. Ce travail de thèse a donc pour objectif de mieux caractériser PfK13 et PfCRT : i) en prédisant les positions qui seraient impliquées dans des interactions protéine-substrat ; et ii) en étudiant les altérations structurales et physico-chimiques induites par les mutations de résistance. Selon les concepts liés à la théorie de l'évolution moléculaire, les mutations touchant des sites exerçant une fonction critique au sein d'une protéine essentielle sont éliminées par la sélection purificatrice. Ces sites sont donc plus conservés que le reste de la protéine. Par des approches bioinformatiques couplant évolution et structure tertiaire, nous avons mis en évidence des régions extrêmement conservées au sein de PfK13 et PfCRT. En comparant ces résultats avec les données expérimentales de protéines / domaines appartenant aux mêmes familles structurales, nous avons identifié plusieurs sites de PfK13 et PfCRT que nous proposons comme candidats pour des interactions protéine-substrat. À notre surprise, les mutations de résistance aux ARTDs ne sont pas localisées à la surface d'interaction que nous avons prédite sur le domaine KREP de PfK13. Les dynamiques moléculaires que nous avons réalisées sur deux mutations de résistance (C580Y et R539T) ont révélé des déstabilisations structurales locales du domaine KREP. Nous supposons que ces mutations pourraient perturber la stabilité du domaine KREP et ainsi diminuer l'abondance cellulaire de PfK13. Concernant PfCRT, les deux modèles de structure tertiaire, prédits par homologie structurale, montrent que la majorité des mutations de résistance à la CQ et à la PPQ sont localisées au niveau d'une poche probable de liaison que nous avons identifiée au coeur du transporteur. Ces mutations altèrent fortement le potentiel électrostatique à la surface de cette poche, passant d'un potentiel neutre à un potentiel électronégatif. Ce changement de propriété physico-chimique de la poche du transporteur est probablement un déterminant majeur de l'acquisition de la propriété de transport de la CQ di-protonée de la vacuole digestive vers le cytoplasme du parasite. Les sites fonctionnels candidats de PfCRT et PfK13 identifiés doivent maintenant être validés par des approches expérimentales. Nous avons initié des approches biochimiques dites de pull-down différentiels pour le domaine KREP de PfK13. Dans un premier temps, certains domaines PfK13, fusionnés à la glutathion S-transférase (GST), ont été exprimés dans la bactérie Escherichia coli, purifiés, puis incubés avec un lysat parasitaire total afin de caractériser des protéines interagissant avec PfK13. Les mises au point de ces expériences se poursuivent. Dans un second temps, des approches génétiques directement chez le parasite, par des techniques de transfection et d'édition de gènes, seront mises en place pour tester l'importance des sites fonctionnels candidats.
Article
Full-text available
Chloroquine (CQ), a 4-aminoquinoline, accumulates in acidic digestive vacuoles of the malaria parasite, preventing conversion of toxic haematin to beta-haematin. We examine how bis 4-aminoquinoline piperaquine (PQ) and its hydroxy-modification (OH-PQ) retain potency on chloroquine-resistant (CQ-R) Plasmodium falciparum. For CQ, PQ, OH-PQ and 4 and 5, representing halves of PQ, beta-haematin inhibitory activity (BHIA) was assayed, while potency was determined in CQ-sensitive (CQ-S) and CQ-R P. falciparum. From measured pK(a)s and the pH-modulated distribution of base between water and lipid (logD), the vacuolar accumulation ratio (VAR) of charged drug from plasma water (pH 7.4) into vacuolar water (pH 4.8) and lipid accumulation ratio (LAR) were calculated. All agents were active in BHIA. In CQ-S, PQ, OH-PQ and CQ were equally potent while 4 and 5 were 100 times less potent. CQ with two basic centres has a VAR of 143,482, while 4 and 5, with two basic centres of lower pK(a)s have VARs of 1287 and 1966. In contrast PQ and OH-PQ have four basic centres and achieve VARs of 104,378 and 19,874. This confirms the importance of VAR for potency against CQ-S parasites. Contrasting results were seen in CQ-R. 5, PQ and OH-PQ with LARs of 693; 973,492 and 398,118 (compared with 8.25 for CQ) showed similar potency in CQ-S and CQ-R. Importance of LAR for potency against CQ-R parasites probably reflects ability to block efflux by hydrophobic interaction with PfCRT but may relate to beta-haematin inhibition in vacuolar lipid.
Article
Full-text available
Most studies on malaria-parasite digestion of hemoglobin (Hb) have been performed using P. falciparum maintained in mature erythrocytes, in vitro. In this study, we examine Plasmodium Hb degradation in vivo in mice, using the parasite P. berghei, and show that it is possible to create mutant parasites lacking enzymes involved in the initial steps of Hb proteolysis. These mutants only complete development in reticulocytes and mature into both schizonts and gametocytes. Hb degradation is severely impaired and large amounts of undigested Hb remains in the reticulocyte cytoplasm and in vesicles in the parasite. The mutants produce little or no hemozoin (Hz), the detoxification by-product of Hb degradation. Further, they are resistant to chloroquine, an antimalarial drug that interferes with Hz formation, but their sensitivity to artesunate, also thought to be dependent on Hb degradation, is retained. Survival in reticulocytes with reduced or absent Hb digestion may imply a novel mechanism of drug resistance. These findings have implications for drug development against human-malaria parasites, such as P. vivax and P. ovale, which develop inside reticulocytes. © 2015 Lin et al.
Article
The transport of lactate is an essential part of the concept of metabolic coupling between neurons and glia. Lactate transport in primary cultures of astroglial cells was shown to be mediated by a single saturable transport system with a Km value for lactate of 7.7 mM and a Vmax value of 250 nmol/(min x mg of protein). Transport was inhibited by a variety of monocarboxylates and by compounds known to inhibit monocarboxylate transport in other cell types, such as alpha-cyano-4-hydroxycinnamate and p-chloromercurbenzenesulfonate. Using reverse transcriptase-polymerase chain reaction and Northern blotting, the presence of mRNA coding for the monocarboxylate transporter 1 (MCT1) was demonstrated in primary cultures of astroglial cells. In contrast, neuron-rich primary cultures were found to contain the mRNA coding for the monocarboxylate transporter 2 (MCT2). MCT1 was cloned and expressed in Xenopus laevis oocytes. Comparison of lactate transport in MCT1 expressing oocytes with lactate transport in glial cells revealed that MCT1 can account for all characteristics of lactate transport in glial cells. These data provide further molecular support for the existence of a lactate shuttle between astrocytes and neurons.
Article
During oogenesis, Xenopus oocytes accumulate large amounts of storage proteins that—after fertilization—provide the developing embryo with building blocks and energy metabolites. The size of the fully developed oocyte (diameter, 1.2 mm) is largely governed by the stored amounts of egg yolk protein. Thus, the mature oocyte is equipped to initiate protein synthesis, cell growth, and replication after fertilization. The size of the oocyte, the large reserve of storage proteins, and its ability to synthesize protein on demand makes the oocyte an almost ideal single-cell expression system. Some basic physical properties of this expression system are listed in Table 1. Table 1 Physicochemical Properties of Oocytes Property Value Ref. Water-accessible volume 368 ± 21 nL (1) Water permeability (1–4)×10−4 cm/s (2) Surface area 18 mm2 −30 mm2 (3) Membrane potential −30 to −60 mV (3) Buffering capacity 20 mM/pH unit at pH 7.0 (4) Intracellular pH 7.4 ± 0.1 (4) Na+ concentration 4–10 mM (3) K+ concentration 76–120 mM (3) Cl− concentration 24–50 mM (3) Ca2+ concentration <0.3 μM (3)