Skip to main content

Monitoring PfMDR1 transport in Plasmodium falciparum

Abstract

Background

The Plasmodium falciparum multidrug resistance 1 transporter, PfMDR1, contains five amino acid polymorphisms that are suggested to be involved in altered drug transport from the parasite’s cytosol into the digestive vacuole (DV). Transport of a substrate into another intracellular compartment influences drug availability at its site of action, therefore making the parasite more susceptible or resistant to a drug. Fluo-4 is a known fluorescent substrate that can be used as a molecular tool to investigate transport dynamics of PfMDR1 in many parasite strains.

Methods

Six P. falciparum strains with varying PfMDR1 mutations were loaded with Fluo-4 AM. Accumulation of the fluorophore in the DV was measured using confocal microscopy. The role of a key amino acid mutation was verified using selected parasite clones with point mutations at PfMDR1 amino acid position 1042. Equal expression of PfMDR1 was confirmed by Western blot.

Results

Fluo-4 was transported by PfMDR1 into the DV of most drug-sensitive and -resistant parasites. Asparagine at PfMDR1 amino acid position 1042 was crucial for Fluo-4 transport, while the N1042D substitution abolished Fluo-4 transport. Competition studies of Fluo-4 with chloroquine, quinine and mefloquine were performed on parasites harbouring asparagine at position 1042. A distinct Fluo-4 transport inhibition pattern for each tested anti-malarial drug was observed in parasite strains of different genetic background.

Conclusion

This study demonstrates that Fluo-4 can be used to investigate PfMDR1 transport dynamics in both drug-sensitive and -resistant parasites. Furthermore, direct evidence of altered Fluo-4 transport in PfMDR1 is linked to a single amino acid mutation in the substrate binding pocket. This system offers a great tool to investigate the role of substrate transport by PfMDR1 and the mutations necessary to support transport, which would lead to new insights for the development of novel anti-malarial drugs.

Background

Emerging resistance to commonly used anti-malarial drugs is a major setback in the fight against malaria worldwide [1]. Understanding the molecular mechanisms behind drug resistance is of high importance in ongoing efforts to control this disease.

Early on, researchers found a correlation between anti-malarial resistance and the Plasmodium falciparum multidrug resistance 1 transporter (PfMDR1) [2, 3]. PfMDR1 is a P-glycoprotein homologue (Pgh1) and belongs to the ATP binding cassette (ABC) transporter superfamily. It is a 162 kDa protein with two nucleotide binding domains (NBD) and twelve transmembrane domains (TMDs), with a putative substrate binding pocket in TMD11 [4]. The transporter is located in the membrane of the digestive vacuole (DV) [5], transporting substrates from the parasite’s cytoplasm into the DV [6].

Two factors have been suggested to play a role in altered drug susceptibility to specific anti-malarial drugs: PfMDR1 polymorphisms and pfmdr1 gene duplications. Resistance has been associated with one or more variations at five amino acid positions in the PfMDR1 transporter, with wild-type PfMDR1 containing the amino acids N86, Y184, S1034, N1042, D1246. Increased pfmdr1 copy numbers have been linked to mefloquine (MQ), lumefantrine (LF), halofantrine (HF), quinine (QN) and artemisinin (AS) resistance [7–9], while PfMDR1 amino acid mutations S1034C, N1042D, D1246Y were found to enhance parasite susceptibility to MQ, HF and AS, independent of gene copy number [2, 10].

Additional evidence for altered drug transport in wild-type versus mutant PfMDR1 was shown through expression of different pfmdr1 variants in Xenopus laevis oocytes. Wild-type PfMDR1 transported QN and chloroquine (CQ) but not HF, while mutant PfMDR1 transported HF but not QN or CQ [11]. Furthermore, residue 184 altered transport kinetics independent of drug binding specificity [12].

Computational models of PfMDR1 describe a substrate-binding pocket that includes the amino acid residues 1034 and 1042 [4, 13]. The binding of several anti-malarial drugs was investigated using docking simulations within the PfMDR1 substrate-binding pocket. Among the tested drugs, MQ was the only candidate whose ability to form an H-bond with residue 1042 was completely abolished through the N1042D substitution [13].

Apart from anti-malarial drug transport, it was shown that several PfMDR1 drug-resistant variants could transport the fluorescent substrate Fluo-4 into the DV of the parasite [6]. In contrast, Fluo-4 transport was abolished in drug-sensitive HB3 parasites harbouring the PfMDR1 variant N86, F184, S1034, D1042, D1246. In a follow-up paper, the pump rate of PfMDR1 (F/Y86, Y184, S1034, N1042, D1246) was determined for Dd2 parasites using live cell imaging of intact infected erythrocytes [14].

In this study, drug-sensitive and -resistant P. falciparum strains of different genetic background and varying PfMDR1 polymorphisms were used to investigate the crucial mutations required for Fluo-4 transport. In addition, P. falciparum clones harbouring a single mutation at position 1042 were analysed. Aspartic acid at position 1042 was found to abolish Fluo-4 transport into the DV. This could be restored by replacing aspartic acid at position 1042 with asparagine. Furthermore, the N1042D substitution resulted in increased sensitivity to MQ. Using Fluo-4 as a competitive substrate offers a powerful tool to investigate the role of PfMDR1 in transport of currently used anti-malarial drugs for both drug-sensitive and -resistant parasites.

Methods

Parasite strains and culture conditions

Three CQ-sensitive (CQS) (3D7, D10, HB3) and three CQ-resistant (CQR) (Dd2, FCR3, FCB) P. falciparum strains, as well as stably transfected P. falciparum clones, derived from the parental lines GC03 and 3BA6, were used in this study. The two strains GC03 and 3BA6 are progeny of the HB3 × Dd2 genetic cross [15] and harbour the PfMDR1 variant from HB3 but differ in their PfCRT phenotype and genotype (Table 1). PfMDR1 mutants derived from GC03 and 3BA6 were produced by Sidhu and colleagues through partial pfmdr1 gene replacement that substituted aspartic acid at position 1042 with asparagine while leaving amino acids at residues 1034 and 1246 unchanged. The resulting clones were: SNDGC03 (S1034, N1042, D1246 in a GC03 genetic background), SND3BA6 (S1034, N1042, D1246 in a 3BA6 genetic background), as well as the recombinant controls SDDGC03 and SDD3BA6 [10]. All strains were cultured continuously, as described by Trager and Jensen [16], with modifications. Briefly, parasites were propagated at 5% haematocrit in culture medium containing RPMI 1640 (Life Technologies, Burlington, ON, Canada) supplemented with 25 mM HEPES, 2 mM l-glutamine, gentamicin (20 µg/ml) (Life Technologies, Burlington, ON, Canada), 100 µM hypoxanthine (Sigma-Aldrich, Oakville, ON, Canada), 0.5% AlbuMAX I (Life Technologies, Burlington, ON, Canada). Parasites were maintained at 37°C with an atmosphere of 5% CO2, 3% O2 and 92% N2. A+ red blood cells were obtained from the Interstate Blood Bank (Memphis, TN, USA). Giemsa-stained blood smears were prepared daily to monitor parasite growth. For synchronization, parasites were treated with 5% d-sorbitol (BioShop Canada, Burlington, ON, Canada) for 10 min at 37°C; sorbitol was removed and parasites were washed once before putting them back into culture. To obtain highly synchronous parasite cultures, sorbitol treatment was repeated after 6–8 h.

Table 1 Main PfCRT and PfMDR1 mutations of Plasmodium falciparum parasites used in this study

DNA isolation and sequence analysis

The full-length sequence of pfmdr1 and partial sequence of pfcrt was verified for all strains. Parasite strains were grown to ≥5% parasitaemia and DNA was extracted using the QIAamp DNA Blood Mini Kit (Qiagen, Toronto, ON, Canada) according to manufacturer’s instructions. DNA was amplified in overlapping PCR fractions using HotStarTaq DNA polymerase (Qiagen, Toronto, ON, Canada). To account for the AT-rich nucleotide content in the P. falciparum genome, dNTPs (Invitrogen Canada, Burlington, ON, Canada) were mixed at 75% AT and 25% GC. For PCR optimization, 2 mM MgCl2, 300 µM dNTPs and 300 nM primers were used for the reaction. For each reaction mix, 20 ng genomic DNA was used. PCR reactions consisted of an initial activation step of 94°C for 3 min, followed by 35 cycles of 94°C for 60 s, 49–61°C (adjusted for each primer pair) for 30 s, and 72°C for 1 min. Primers used for pfmdr1 gene sequencing are described in [14]. Primers used for sequencing the pfcrt gene region containing the main mutation sites were: pfcrt_F: 5′-GGAGGTTCTTGTCTTGGTAAATG, pfcrt_R: 5′-TGGTAGGTGGAATAGATTCTCTTATAAA. Samples were sent for sequencing to Genome Quebec, Canada and analysed using the BioEdit software [17].

Protein expression

To determine PfMDR1 protein levels, 20 µg whole cell protein lysates were loaded in each lane of an 8% acrylamide gel containing SDS. The proteins were transferred onto a PVDF membrane, which was then blocked O/N at 4°C with 5% milk (w/v) and 0.05% Tween-20 (ACP Chemicals, St-Leonard, QC, Canada) in phosphate buffered saline (PBS). The membrane was further incubated with the appropriate dilution of primary anti-PfMDR1 (kindly provided by Prof Cowman, Walter and Eliza Hall Institute, VIC, Australia) or anti-PfHSP70 (GenWay Biotech, San Diego, CA, USA) antibody (1:2,000) in PBS-T + 5% milk for 1 h (RT), then washed and incubated with HRP-conjugated anti-rabbit IgG secondary antibody (Abcam, Toronto, ON, Canada) (1:20,000) for 1 h (RT). Immunoreactive bands were detected with ImmunStar WesternC Chemiluminescent Kit (Bio-Rad Laboratories, Mississauga, ON, Canada) using a myECL Imager (Thermo Scientific, Burlington, ON, Canada). To confirm equal protein loading, chemiluminescence intensities of PfMDR1 were calculated for each parasite clone relative to the respective PfHSP70 chemiluminescence using ImageJ 1.47q (National Institutes of Health, USA).

Growth inhibition assay

Growth inhibition assays were performed as described previously [18], with modifications. Briefly, synchronized ring stage parasites were diluted to a final parasitaemia of 0.5% and a haematocrit of 2%. A total of 100 µl culture medium per well was prepared in a 96-well plate assay, with a drug dilution series of 1:3, ranging from 1 µM to 0.15 nM. Plates were incubated at 37°C, 5% CO2, 3% O2 and 92% N2 for 72 h, then frozen and stored at −80°C. Plates were thawed at room temperature and 100 µl 2Ă— lysis buffer (20 mM Tris pH 7.5, 5 mM EDTA, 0.008% saponin, 0.08% Triton X-100, 0.2 µl SYBR Green I/ml) was added to each well. Plates were incubated in the dark for at least 1 h. Fluorescence intensity was determined using a Synergy H4 plate reader (Fisher Scientific, Nepean, ON, Canada) with 485 nm excitation and 520 nm emission wavelengths. IC50 values were determined by fitting concentration response curves with a custom-made procedure for IGOR Pro 6.2 based on an R script kindly provided by Le Nagard [19, 20].

Live cell imaging

Synchronized trophozoite stage parasites were loaded with 5 µM Fluo-4 AM (Life Technologies, Burlington, ON, Canada) in Ringer’s solution (122.5 mM NaCl, 5.4 mM KCl, 1.2 mM CaCl2, 0.8 mM MgCl2, 11 mM d-glucose, 10 mM HEPES, 1 mM NaH2PO4, pH 7.4) for 50 min at 37°C. Cells were then washed twice with Ringer’s solution and transferred to a microscope chamber. Parasites were kept at 37°C during microscopy using a stage-top incubator (Tokai Hit, Shizuoka-ken, Japan). A series of four images per parasite was taken using a Zeiss LSM710 confocal microscope (Carl Zeiss, Oberkochen, Germany) equipped with a water-corrected objective (C-apochromat 63×/1.20 W Korr M27) and a 488 nm laser (12.5 mW, 2% intensity). The range of emitted fluorescence was measured from 493 to 622 nm. Images were analysed using ImageJ 1.47q (National Institutes of Health, USA).

Results and discussion

PfMDR1 transports Fluo-4 into the digestive vacuole

A genetic linkage between Fluo-4 accumulation in the DV and the PfMDR1 transporter was previously described [6]. Although variations in Fluo-4 accumulation were observed between parasite strains harbouring different PfMDR1 mutations, the amino acid mutation(s) responsible for Fluo-4 transport remained to be determined. To identify PfMDR1 mutation(s) crucial for Fluo-4 transport, several drug-sensitive and -resistant P. falciparum strains of different genetic backgrounds were tested for accumulation of Fluo-4 in the DV. Three CQS and three CQR strains harbouring different PfMDR1 mutations were selected for these experiments (Table 1). While PfMDR1 has been suggested to play a role in CQ resistance, the key genetic indicator for CQS versus CQR parasites is attributed to the amino acid mutation K76T in the P. falciparum chloroquine resistance transporter (PfCRT) [21]. This mutation was taken into consideration in this study and the relevance of PfMDR1 polymorphisms is discussed for strains harbouring either lysine (K76) or threonine (K76T) in PfCRT.

Of the five mutation sites in PfMDR1 known to play a role in drug resistance, the parasite strains used for these experiments contained mutations at amino acid positions 86, 184 and 1042, which are mainly found in African and Asian isolates [7, 22–24]. To determine if mutations at these residues influence transport of Fluo-4 via PfMDR1, accumulation of Fluo-4 in the DV was measured in live parasites using confocal microscopy. All parasite strains, except HB3, showed high accumulation of Fluo-4 in the DV (Figure 1a, b). Fluo-4 accumulation in the DV was impeded by pre-incubation of the parasites with tariquidar (TQ) (Figure 1b), a specific inhibitor of Pgp transporters [25]. Importantly, Fluo-4 accumulation in the DV was shown to be independent of the PfCRT K76T mutation, suggesting that the PfCRT mutation alone has no effect Fluo-4 accumulation.

Figure 1
figure 1

Fluo-4 fluorescence in P. falciparum parasites. Parasites were incubated with 5 µM Fluo-4. a Representative images of P. falciparum-infected erythrocytes. Scale bar 5 µm. b Mean Fluo-4 fluorescence ratio (DV/cytosol) ± SEM. When treated with the Pgp inhibitor tariquidar (TQ), fluorescence ratio was reduced, indicating that Fluo-4 transport occurred exclusively through PfMDR1. Total n ≥ 67 for each strain, done in three independent experiments on different days.

Fluo-4 transport is associated with PfMDR1 N1042

The parasite strain HB3, which does not accumulate Fluo-4 in the DV, harbours two PfMDR1 polymorphisms—Y184F and N1042D—not found in the other tested P. falciparum strains (Table 1), suggesting that one of these two residues is responsible for the altered Fluo-4 transport seen in HB3 parasites. It has previously been demonstrated for the human homolog of PfMDR1 that the Y184F mutation does not influence drug specificity [12]. Furthermore, this residue only mildly increased drug resistance in comparison to other PfMDR1 mutation sites [4]. In contrast, amino acid position 1042 is thought to be part of the transporter binding pocket and forms electrostatic interactions with amodiaquine, CQ, MQ (only in the presence of asparagine), HF, vinblastine and vincristine [4, 10, 13]. Therefore, the role of PfMDR1 mutations at position 1042 was investigated in more detail.

To verify the influence of PfCRT on Fluo-4 transport, existing parasite clones containing either lysine or threonine at residue 76 in PfCRT were selected and compared. For this, stably transfected P. falciparum clones derived from the parental lines GC03 and 3BA6 were used, where GC03 harbours the K76 PfCRT genotype and is CQ sensitive, and 3BA6 contains the PfCRT K76T mutation that confers CQ resistance (Table 1). Both GC03 and 3BA6 contain the PfMDR1 sequence N86, F184, S1034, D1042, D1246, identical to the HB3 strain. Experiments verified that neither GC03 nor 3BA6 were able to accumulate Fluo-4 in the DV, as expected (Figure 2a, b).

Figure 2
figure 2

Fluo-4 fluorescence in P. falciparum clones with varying PfMDR1 and PfCRT mutations. Parasites were incubated with 5 µM Fluo-4 AM in Ringer’s solution for 50 min at 37°C, then washed with Ringer’s solution and transferred onto a microscope chamber. a Single images of P. falciparum clones. Scale bar 5 µm. b Mean Fluo-4 fluorescence ratio measured from the digestive vacuole (DV) and the cytosol. When treated with tariquidar (TQ), parasites were pre-incubated with 100 nM TQ for 10 min at 37°C before adding Fluo-4 AM for 50 min. Total n ≥ 80 for each strain, done in three independent experiments on different days. Error bars represent SEM. ***p < 0.0001. c Western blot of synchronized trophozoite stage parasites using anti-PfMDR1 antibodies. Anti-PfHSP70 was used as a loading control. PfMDR1 protein expression levels were normalized to PfHSP70 and measured in triplicate ± SEM.

As a control, parasite clones SDDGC03 and SDD3BA6, harbouring the same mutation as GC03 and 3BA6, revealed that the PfMDR1 N1042D mutation did not transport Fluo-4 into the DV (Figure 2a, b). Fluo-4 transport was only established with the introduction of asparagine at residue 1042 for both clones (SNDGC03 and SND3BA6). Transport was independent of the PfCRT K76T mutation, since Fluo-4 transport was equally impaired in GC03 and 3BA6 parental lines as well as the SDDGC03 and SDD3BA6 controls, and only seen in the clones SNDGC03 and SND3BA6 (Figure 2a, b). Transport specificity for PfMDR1 in clones SNDGC03 and SND3BA6 was confirmed through pre-incubation of the samples with the Pgp inhibitor TQ (Figure 2b; Additional file 1: Figure S1). These results indicate a pivotal role for asparagine at PfMDR1 residue 1042 in Fluo-4 transport, independent of PfCRT.

In addition to PfMDR1 polymorphisms, increased pfmdr1 gene copy number [8, 9, 26] and protein expression levels have also been suggested to play a role in drug resistance or susceptibility. To verify that Fluo-4 accumulation in the DV of clones SNDGC03 and SND3BA6 was not linked to increased PfMDR1 expression, PfMDR1 protein levels were analysed for these parasite clones (Figure 2c). No significant increase in PfMDR1 expression was detected for the GC03 and 3BA6 parental lines or their clones harbouring the D1042N substitution (p > 0.05). Therefore, Fluo-4 accumulation in the DV of these clones was only associated with the PfMDR1 mutation at residue 1042.

Effect of N1042D polymorphism on drug susceptibility

Since transport of Fluo-4 was dependent on specific PfMDR1 polymorphisms, it follows that the N1042D substitution in the PfMDR1 substrate-binding pocket may not only alter the binding and transport of Fluo-4 but also influence other substrates, such as anti-malarial drugs. To determine if the amino acid substitution at residue 1042 resulted in altered drug sensitivity or resistance of the parasite clones, growth inhibition assays were performed. A change in a given drug’s IC50 would indicate a role for PfMDR1 amino acid position 1042 in its transport. CQ IC50 values for CQS clones were found to be 47 ± 2.6 nM for GC03, 46 ± 2.2 nM for SDDGC03 and 49 ± 2.8 nM for SNDGC03, while CQ IC50 values for CQR clones were 262 ± 5.7 nM for 3BA6, 204 ± 9.4 nM for SDD3BA6 and 274 ± 5.6 nM for SND3BA6 (Table 2). This suggests that an amino acid mutation at PfMDR1 residue 1042 does not affect susceptibility or resistance to CQ. CQ resistance could be reversed in the CQR strains through the addition of 1 µM verapamil, while no significant difference was observed in CQS strains (p > 0.05), as expected. Similarly, CQS clones were more susceptible to quinacrine (QC) with IC50 values of 11 ± 0.3 nM for GC03, 13 ± 0.3 nM for SDDGC03 and 14 ± 0.4 nM for SNDGC03 compared to the higher QC IC50 values of 43 ± 3.4 nM for 3BA6, 41 ± 1.0 nM for SDD3BA6 and 42 ± 2.3 nM for SND3BA6. Here again, as for CQ, an amino acid mutation at PfMDR1 residue 1042 did not affect susceptibility or resistance to QC. The same was found for dihydroartemisinin (DHA). CQS clones were fourfold more resistant than CQR clones (4 ± 0.1 vs 1 ± 0.1 nM) and DHA resistance was not reversible by the addition of 1 µM verapamil. Therefore, the PfMDR1 N1042D mutation alone does not appear to be a major determinant for CQ, QC or DHA resistance in these parasites (Table 2).

Table 2 IC50 values of Plasmodium falciparum parasites used in this study

Interestingly, the N1042D mutation did provide changes in drug susceptibility for the anti-malarial drugs QN and MQ. A significant decrease in QN IC50 values was observed in the GC03 and 3BA6-derived clones SNDGC03 and SND3BA6 compared to the parental lines GC03 and 3BA6 or their recombinant controls SDDGC03 and SDD3BA6 (p < 0.005) (Table 2), which is in agreement with previous findings [10]. The opposite effect was observed for MQ. For the clones SNDGC03 and SND3BA6clones, the D1042N substitution led to an approximate threefold increase in resistance to MQ. MQ IC50 values increased from 6 ± 0.9 nM in GC03 to 19 ± 0.6 nM in SNDGC03, and from 5 ± 0.3 nM in 3BA6 to 14 ± 0.5 nM in SND3BA6. This was a significant increase in MQ IC50 values for the clones SNDGC03 and SND3BA6 compared to the parental lines (p = 0.0003 for GC03 vs SNDGC03; p = 0.0001 for 3BA6 vs SND3BA6). Similar increases in MQ resistance (approximately threefold) have been found in field isolates from Africa and Asia [27, 28]. A threefold increase in MQ resistance was also confirmed in vitro through allelic exchange of pfmdr1 [2]. Therefore, the PfMDR1 mutation crucial for MQ resistance in vitro is likely to play a role in increased MQ resistance in the field. Nevertheless, growth inhibition experiments do not fully elucidate altered transport of MQ by PfMDR1. For this reason, co-incubation of anti-malarial drugs with Fluo-4 can provide additional insight into the role of PfMDR1 transport in drug resistance.

Substrate competition of anti-malarial drugs with Fluo-4 to determine drug transport by PfMDR1

PfMDR1 is involved in the transport of substrates, including various anti-malarial drugs, from the cytosol into the DV [10, 11] but the role of PfMDR1 polymorphisms in drug resistance is not fully understood. Mutations at residue 86 has been suggested to allosterically influence the TMD11 drug binding site [4], while PfMDR1 residues 1034 and 1042 are located in a proposed binding pocket. Electrostatic interactions with both asparagine and aspartic acid at position 1042 have been demonstrated in silico for CQ, QN and MQ [4, 13]. While CQ was unable to form a hydrogen (H)-bond with either asparagine or aspartic acid at residue 1042, asparagine at residue 1042 was able to form a H-bond with MQ [4, 13]. No information is available on H-bond formation of QN with residue 1042.

Further insight into the importance of PfMDR1 in drug transport can be achieved through competition studies using two potentially competitive substrates, e.g., fixed amounts of Fluo-4 and increasing drug concentrations. For this purpose, two parasite strains were chosen to evaluate potential differences in anti-malarial drug transport in sensitive (3D7) and resistant (Dd2) parasites. 3D7 is sensitive to CQ, QN and MQ (Table 2), while Dd2 is resistant to these drugs. HB3 was not used, since it is not able to transport Fluo-4 into the DV. The measured changes of Fluo-4 accumulation in the DV in the presence of anti-malarial drugs were compared for 3D7 and Dd2 parasites.

For 3D7 parasites, increasing CQ concentration led to the strong decrease in Fluo-4 accumulation in the DV (Figure 3a). While TQ is a non-competitive inhibitor of MDR1 through inhibition of the ATPase activity [25], CQ is thought to bind to the transporter’s substrate-binding pocket and is likely a competitive inhibitor of Fluo-4. Pre-incubation of 3D7 parasites with 250 nM CQ resulted in decreased Fluo-4 accumulation in the DV, where only 5 ± 1.0% Fluo-4 fluorescence was measured when compared to non drug-treated parasites. In Dd2 parasites, pre-incubation with 250 nM CQ reduced Fluo-4 fluorescence in the DV by half (48 ± 3.0%). MQ reduced Fluo-4 transport into the DV to 45 ± 2.0% in 3D7 parasites and only 82 ± 6.0% in Dd2 parasites (Figure 3a). For MQ, a decrease in Fluo-4 transport in 3D7 compared to Dd2 parasites cannot be attributed to a PfMDR1 amino acid mutation at position 1042 alone since both strains harbour wild-type N1042. However, Dd2 parasites harbour an amino acid mutation at PfMDR1 residue 86, which was described to influence MQ sensitivity in African field isolates [24, 30, 31]. This N86Y mutation may alter MQ transport by PfMDR1, as shown in the experiments presented here. QN was less effective in reducing Fluo-4 accumulation than CQ or MQ and only decreased Fluo-4 fluorescence in the DV to 59 ± 2.9% in 3D7 parasites and not significantly in Dd2 parasites (92 ± 3.4%) (Figure 3a). The differences between 3D7 and Dd2 parasites at the level of Fluo-4 reduction may be linked to PfMDR1 residue 86, which is suggested to influence CQ susceptibility [29]. It is conceivable that N86Y can allosterically influence the drug binding site at TMD11, as suggested in [4]. Residue 86 was not examined in this study.

Figure 3
figure 3

Competition of Fluo-4 transport with anti-malarial drugs. a Parasites were pre-incubated with different concentrations of chloroquine (CQ), mefloquine (MQ) or quinine (QN), or left untreated before adding Fluo-4 AM. b Parasites were pre-incubated as described with 250 nM of CQ, MQ or QN. Experiments were done in triplicate 3 independent days. Error bars represent SEM. *P < 0.05; ***P < 0.0001.

Fluo-4 competition with anti-malarial drugs was also tested on the parasite clones SNDGC03 and SND3BA6. Using a single drug concentration of 250 nM, Fluo-4 accumulation was decreased in the DV of SNDGC03 and SND3BA6 parasites to 33 ± 1.9% and 39 ± 3.0% for CQ, 61 ± 3.1% and 65 ± 4.8% for MQ, and 58 ± 3.9% and 91 ± 4.4% for QN, respectively. Almost all tested anti-malarial drugs decreased Fluo-4 accumulation in the DV in both SNDGC03 and SND3BA6 clones, except for QN (Figure 3b). This again suggests that the influence of anti-malarial drugs on substrate transport by PfMDR1 is not only dependent on PfMDR1 polymorphisms but also on the varying genetic background of the parasite strain.

The influence of the N1042D substitution on Fluo-4 transport is explained in a proposed model (Figure 4). Fluo-4 is negatively charged and replacement of a neutral asparagine at residue 1042 with a negatively charged aspartic acid alters the local environment, making it less favourable to Fluo-4 transport. TQ does not influence Fluo-4 affinity for the substrate-binding pocket but prevents adenosine triphosphate (ATP) hydrolysis of PfMDR1. Since both NBDs act in concert [32], complete inhibition of ATPase activity by TQ can be achieved when one nucleotide-binding site is blocked.

Figure 4
figure 4

Model of Fluo-4 transport. The docking of substrates in the PfMDR1 binding pocket is influenced by the size, charge and polarity of local amino acids. Asparagine (N) and aspartic acid (D) are both polar, hydrophilic and small in volume. While N is an amide and uncharged, D is acidic and negatively charged. a For parasites containing the PfMDR1 N1042 polymorphism, the intrinsically negatively charged Fluo-4 gets transported from the cytoplasm into the digestive vacuole (DV) where it accumulates. b In the presence of PfMDR1 N1042D, which adds a negative charge to the binding pocket, Fluo-4 does not get transported by PfMDR1 and no Fluo-4 accumulation is detected in the DV. c Fluo-4 transport is abolished through the addition of the Pgp inhibitor tariquidar (TQ), which does not alter the substrate binding but prevents ATP hydrolysis and therefore the conformational change that is necessary for substrate dislocation.

Detailed transport kinetics for parasite strains of different origin will provide additional information on substrate transport via PfMDR1. Using Fluo-4 as a competitive substrate circumvents the issue of direct labelling of anti-malarial drugs with a fluorescent tag that may alter their transport properties. Therefore, Fluo-4 is a powerful tool to selectively study transport kinetics of PfMDR1 in intact P. falciparum-infected erythrocytes.

Conclusion

This study provides direct evidence for reduced PfMDR1-driven substrate transport through the single amino acid mutation N1042D, located in the substrate-binding pocket. The relevance of this amino acid mutation may have been underestimated and needs further investigation. Furthermore, it is now possible to test additional mutations within the binding pocket of PfMDR1 to verify their potential role in drug transport. Accordingly, using Fluo-4 in drug competition assays provides a powerful tool to better examine drug transport kinetics via PfMDR1. This newly acquired information can help elucidate the role of PfMDR1 in drug transport, which has remained controversial for decades. For example, while earlier investigations have suggested a link between MQ resistance and asparagine at PfMDR1 amino acid position 86 in isolates from Thailand and West Africa [30, 33], others have associated wild-type PfMDR1 and increased gene copy numbers with MQ resistance [3, 34]. More recent publications have found strong evidence for a role for N1042 in MQ resistance in Thai isolates [35, 36], suggesting that the importance of this mutation has been underestimated in previous field studies. The importance of N1042 for MQ resistance was similarly demonstrated in vitro through combined amino acid substitutions that generated a parasite strain harbouring the mutations S1024C, N1042D, D1246Y [2]. This newly generated strain was approximately threefold more susceptible to MQ exposure compared to the parental strain [2]. The results presented here support a potential role of PfMDR1 N1042 in parasite resistance to MQ. Further investigations will help elucidate the significance of this polymorphism on substrate transport.

References

  1. Sibley CH (2014) Understanding drug resistance in malaria parasites: basic science for public health. Mol Biochem Parasitol 195:107–114

    Article  CAS  PubMed  Google Scholar 

  2. Reed MB, Saliba KJ, Caruana SR, Kirk K, Cowman AF (2000) Pgh1 modulates sensitivity and resistance to multiple antimalarials in Plasmodium falciparum. Nature 403:906–909

    Article  CAS  PubMed  Google Scholar 

  3. Duraisingh MT, von Seidlein LV, Jepson A, Jones P, Sambou I, Pinder M et al (2000) Linkage disequilibrium between two chromosomally distinct loci associated with increased resistance to chloroquine in Plasmodium falciparum. Parasitology 121:1–7

    Article  CAS  PubMed  Google Scholar 

  4. Ferreira PE, Holmgren G, Veiga MI, Uhlen P, Kaneko A, Gil JP (2011) PfMDR1: mechanisms of transport modulation by functional polymorphisms. PLoS One 6:e23875

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  5. Cowman AF, Karcz S, Galatis D, Culvenor JG (1991) A P-glycoprotein homologue of Plasmodium falciparum is localized on the digestive vacuole. J Cell Biol 113:1033–1042

    Article  CAS  PubMed  Google Scholar 

  6. Rohrbach P, Sanchez CP, Hayton K, Friedrich O, Patel J, Sidhu AB et al (2006) Genetic linkage of pfmdr1 with food vacuolar solute import in Plasmodium falciparum. EMBO J 25:3000–3011

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  7. Pickard AL, Wongsrichanalai C, Purfield A, Kamwendo D, Emery K, Zalewski C et al (2003) Resistance to antimalarials in Southeast Asia and genetic polymorphisms in pfmdr1. Antimicrob Agents Chemother 47:2418–2423

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  8. Price RN, Uhlemann AC, Brockman A, McGready R, Ashley E, Phaipun L et al (2004) Mefloquine resistance in Plasmodium falciparum and increased pfmdr1 gene copy number. Lancet 364:438–447

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  9. Sidhu AB, Uhlemann AC, Valderramos SG, Valderramos JC, Krishna S, Fidock DA (2006) Decreasing pfmdr1 copy number in Plasmodium falciparum malaria heightens susceptibility to mefloquine, lumefantrine, halofantrine, quinine, and artemisinin. J Infect Dis 194:528–535

    Article  PubMed Central  PubMed  Google Scholar 

  10. Sidhu AB, Valderramos SG, Fidock DA (2005) pfmdr1 mutations contribute to quinine resistance and enhance mefloquine and artemisinin sensitivity in Plasmodium falciparum. Mol Microbiol 57:913–926

    Article  CAS  PubMed  Google Scholar 

  11. Sanchez CP, Rotmann A, Stein WD, Lanzer M (2008) Polymorphisms within PfMDR1 alter the substrate specificity for anti-malarial drugs in Plasmodium falciparum. Mol Microbiol 70:786–798

    CAS  PubMed  Google Scholar 

  12. Safa AR, Stern RK, Choi K, Agresti M, Tamai I, Mehta ND et al (1990) Molecular basis of preferential resistance to colchicine in multidrug-resistant human cells conferred by Gly-185—Val-185 substitution in P-glycoprotein. Proc Natl Acad Sci USA 87:7225–7229

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  13. Patel SK, George L-B, Prasanth Kumar S, Highland HN, Jasrai YT, Pandya HA et al (2013) A computational approach towards the understanding of Plasmodium falciparum multidrug resistance protein 1. ISRN. Bioinformatics 2013:15

    Google Scholar 

  14. Friedrich O, Reiling SJ, Wunderlich J, Rohrbach P (2014) Assessment of Plasmodium falciparum PfMDR1 transport rates using Fluo-4. J Cell Mol Med 18:1851–1862

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  15. Wellems TE, Panton LJ, Gluzman IY, do Rosario VE, Gwadz RW, Walker-Jonah A et al (1990) Chloroquine resistance not linked to mdr-like genes in a Plasmodium falciparum cross. Nature 345:253–255

  16. Trager W, Jensen JB (1976) Human malaria parasites in continuous culture. Science 193:673–675

    Article  CAS  PubMed  Google Scholar 

  17. Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp 41:95–98

    CAS  Google Scholar 

  18. Bacon DJ, Latour C, Lucas C, Colina O, Ringwald P, Picot S (2007) Comparison of a SYBR green I-based assay with a histidine-rich protein II enzyme-linked immunosorbent assay for in vitro antimalarial drug efficacy testing and application to clinical isolates. Antimicrob Agents Chemother 51:1172–1178

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  19. Le Nagard H, Vincent C, Mentre F, Le Bras J (2011) Online analysis of in vitro resistance to antimalarial drugs through nonlinear regression. Comput Methods Programs Biomed 104:10–18

    Article  PubMed  Google Scholar 

  20. Kaddouri H, Nakache S, Houze S, Mentre F, Le Bras J (2006) Assessment of the drug susceptibility of Plasmodium falciparum clinical isolates from africa by using a Plasmodium lactate dehydrogenase immunodetection assay and an inhibitory maximum effect model for precise measurement of the 50-percent inhibitory concentration. Antimicrob Agents Chemother 50:3343–3349

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  21. Lakshmanan V, Bray PG, Verdier-Pinard D, Johnson DJ, Horrocks P, Muhle RA et al (2005) A critical role for PfCRT K76T in Plasmodium falciparum verapamil-reversible chloroquine resistance. EMBO J 24:2294–2305

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  22. Chaiyaroj SC, Buranakiti A, Angkasekwinai P, Looressuwan S, Cowman AF (1999) Analysis of mefloquine resistance and amplification of pfmdr1 in multidrug-resistant Plasmodium falciparum isolates from Thailand. Am J Trop Med Hyg 61:780–783

    CAS  PubMed  Google Scholar 

  23. Mawili-Mboumba DP, Kun JF, Lell B, Kremsner PG, Ntoumi F (2002) Pfmdr1 alleles and response to ultralow-dose mefloquine treatment in Gabonese patients. Antimicrob Agents Chemother 46:166–170

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  24. Rason MA, Andrianantenaina HB, Ariey F, Raveloson A, Domarle O, Randrianarivelojosia M (2007) Prevalent pfmdr1 N86Y mutant Plasmodium falciparum in Madagascar despite absence of pfcrt mutant strains. Am J Trop Med Hyg 76:1079–1083

    CAS  PubMed  Google Scholar 

  25. Fox E, Bates SE (2007) Tariquidar (XR9576): a P-glycoprotein drug efflux pump inhibitor. Expert Rev Anticancer Ther 7:447–459

    Article  CAS  PubMed  Google Scholar 

  26. Witkowski B, Nicolau ML, Soh PN, Iriart X, Menard S, Alvarez M et al (2010) Plasmodium falciparum isolates with increased pfmdr1 copy number circulate in West Africa. Antimicrob Agents Chemother 54:3049–3051

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  27. Childs GE, Pang L, Wimonwattrawatee T, Pooyindee N, Nanakorn A, Limchitee S et al (1987) In vitro mefloquine resistance of Plasmodium falciparum isolated from the Burmese border region of Thailand. Southeast Asian J Trop Med Public Health 18:438–443

    CAS  PubMed  Google Scholar 

  28. Oduola AM, Omitowoju GO, Gerena L, Kyle DE, Milhous WK, Sowunmi A et al (1993) Reversal of mefloquine resistance with penfluridol in isolates of Plasmodium falciparum from south-west Nigeria. Trans R Soc Trop Med Hyg 87:81–83

    Article  CAS  PubMed  Google Scholar 

  29. Duraisingh M, Drakeley C, Muller O, Bailey R, Snounou G, Targett G et al (1997) Evidence for selection for the tyrosine-86 allele of the pfmdr1 gene of Plasmodium falciparum by chloroquine and amodiaquine. Parasitology 114:205–211

    Article  CAS  PubMed  Google Scholar 

  30. Duraisingh MT, Jones P, Sambou I, von Seidlein L, Pinder M, Warhurst DC (2000) The tyrosine-86 allele of the pfmdr1 gene of Plasmodium falciparum is associated with increased sensitivity to the anti-malarials mefloquine and artemisinin. Mol Biochem Parasitol 108:13–23

    Article  CAS  PubMed  Google Scholar 

  31. Eyase FL, Akala HM, Ingasia L, Cheruiyot A, Omondi A, Okudo C et al (2013) The role of pfmdr1 and pfcrt in changing chloroquine, amodiaquine, mefloquine and lumefantrine susceptibility in western-Kenya P. falciparum samples during 2008–2011. PLoS One 8:e64299

    Article  CAS  PubMed Central  PubMed  Google Scholar 

  32. Urbatsch IL, Sankaran B, Bhagat S, Senior AE (1995) Both P-glycoprotein nucleotide-binding sites are catalytically active. J Biol Chem 270:26956–26961

    Article  CAS  PubMed  Google Scholar 

  33. Price RN, Cassar C, Brockman A, Duraisingh M, van Vugt M, White NJ et al (1999) The pfmdr1 gene is associated with a multidrug-resistant phenotype in Plasmodium falciparum from the Western border of Thailand. Antimicrob Agents Chemother 43:2943–2949

    CAS  PubMed Central  PubMed  Google Scholar 

  34. Cowman AF, Crabb BS (2002) A parasite genome sheds light on an old enemy. Nat Biotechnol 20:1098–1099

    Article  CAS  PubMed  Google Scholar 

  35. Pillai DR, Hijar G, Montoya Y, Marouino W, Ruebush TK 2nd, Wongsrichanalai C et al (2003) Lack of prediction of mefloquine and mefloquine-artesunate treatment outcome by mutations in the Plasmodium falciparum multidrug resistance 1 (pfmdr1) gene for P. falciparum malaria in Peru. Am J Trop Med Hyg 68:107–110

    CAS  PubMed  Google Scholar 

  36. Anderson TJ, Nair S, Qin H, Singlam S, Brockman A, Paiphun L et al (2005) Are transporter genes other than the chloroquine resistance locus (pfcrt) and multidrug resistance gene (pfmdr) associated with antimalarial drug resistance? Antimicrob Agents Chemother 49:2180–2188

    Article  CAS  PubMed Central  PubMed  Google Scholar 

Download references

Authors’ contributions

SJR and PR designed the study; SJR performed experiments and quantified the data; SJR and PR wrote the manuscript. Both authors read and approved the final manuscript.

Acknowledgements

This study was supported in part by fellowships from the German Academic Exchange Service (DAAD) (SJR), the Robert P. Harpur Fellowship (SJR), the Centre for Host-Parasite Interaction (FRQNT) (SJR), and grants from the Natural Sciences and Engineering Research Council (NSERC) Discovery Grant (PR) and the Canada Foundation for Innovation (CFI) Leaders Opportunity Fund (PR).

Compliance with ethical guidelines

Competing interests The authors declare that they have no competing interests.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Petra Rohrbach.

Additional file

Additional file 1:

Figure S1. Inhibition of Fluo-4 transport through tariquidar. Plasmodium falciparum clones were pre-incubated with 100 nM tariquidar for 10 min at 37°C, then 5 µM Fluo-4 AM was added for 50 min. Fluo-4 accumulation in the digestive vacuole was diminished. Experiments were done in triplicate on different days. These images are supplementary to the data obtained for Figure 2. Scale bar, 5 µm.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Reiling, S.J., Rohrbach, P. Monitoring PfMDR1 transport in Plasmodium falciparum . Malar J 14, 270 (2015). https://0-doi-org.brum.beds.ac.uk/10.1186/s12936-015-0791-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12936-015-0791-3

Keywords